content
stringlengths 1
15.9M
|
---|
\section{INTRODUCTION}
This paper is based upon the research that Joe and I carried out around 1970. Motivated by the papers of Byers and Yang \cite{by} and Bloch \cite{bloch}, we demonstrated that the Bloch oscillations associated with a superconductor were present in a thin ring of an Ideal Fermi gas, albeit with a flux quantum determined by a single electron charge (thus, $hc/e$), in contrast with $hc/2e$ for the BCS superconductor. We also showed that there were analogous oscillations in the behavior of a rotating ring of an ideal Bose gas. None of our results were published. However, the results for the Ideal Fermi gas ultimately were and still are being developed much further for a real metal in the normal state (with important contributions made by Joe and his coworkers), including elastic scattering of electrons off impurities, electron scattering off magnetic impurities, electron-electron scattering, and inelastic electron-phonon scattering. See Joe's monograph \cite{joe1} for more details as well as the very recent paper by Joe and collaborators \cite{ham}. The essential predictions for the normal metal ring were ultimately observed in a number of beautiful experiments \cite{levy}. On the other hand, the work on the Bose gas was left in the bins of old notes. The purpose of this paper is to revive this problem and to honor my relationship with Joe.
One of the key steps in the formulation of this problem is, for me at least, based upon a talk by Gordon Baym presented at the 1967 Summer School held at St. Andrews, Scotland on "Mathematical Methods in Solid State and Superfluid Theory".\cite{andrews} Baym pointed out that if a fluid is placed in a rotating cylinder, the energies in the partition function must be those with respect to the rotating frame of reference. We will later see how this requirement leads to Bloch and Josephson oscillations in a ring of an Ideal Bose gas.
By coincidence, it was at this Summer School that I was fortunate to meet Joe. We had a marriage of like minds immediately. Joe was on his way to the US for a leave of absence, fresh from his work on demonstrating that while a one-dimensional Ising model with short range interactions does not exhibit a phase transition in the thermodynamic limit (that is, as the number of spins N goes to infinity), a finite chain of spins could exhibit long range order that is not significantly weaker than that obtained in the thermodynamic limit in a system with long range interactions . I had come to other corresponding conclusions about finite systems during my post-doc at Orsay that year. There I had heard Bernard Jancovi\c{c}i \cite{jan} give a talk on novel behavior of the susceptibility of a finite 2D harmonic lattice in spite of the accepted result that a 2D harmonic lattice was unstable. His analysis revealed that the long wavelength divergence of fluctuations that destroyed long range order were cut off in a finite system. That same year, I had also heard Vladimir Tkachenko \cite{tka} give a talk about his results demonstrating that a rotating lattice of vortices in a superfluid was unstable due to the diverging fluctuations in the vibrations of the vortices. I was led to note that the divergence vanished in a finite lattice, due to a cutoff in the long-wavelength fluctuations that was proportional to log-N.
Our meeting at St. Andrews led to many years of wonderful times together as friends as well as in research that reflected our shared view that theory must always strive to honestly reflect experimental conditions. I believe that this philosophy was the foundation for Joe's incredible ability to translate theory into a language that experimentalists could not only understand but also use to produce wonderful experimental results. Joe would often come up with great suggestions for new areas of research. However, for me it was even more delightful that whenever I had an idea, Joe was there to analyze it and lead us with an explosion of further developments, ways of understanding the phenomena and simplifying the analysis.
I had the pleasure of working with Joe and his colleagues on research that revealed that the study of finite systems could not only make some non-existence theorems based upon infinite systems of not too great importance practically, but also teach us much about systems in the thermodynamic limit and reveal novel behavior in real, finite systems.
We take all this for granted since the rise of nanoscopic physics in the '70s.
\section{Comparison of a Ring of an Ideal Fermi Gas with a Ring of an Ideal Bose Gas}
\subsection{Energy of the Quantum States}
Our ring has a radius R and cross-sectional dimensions that are so small that the first excited state with respect to either dimension has an energy much larger than the thermal energy kT. As a result, the ring is effectively one-dimensional. The rotational momentum is quantized in the inertial lab frame. Thus we have
\begin{equation}
p=n\frac{\hbar}{R}
\label{eq:mom}
\end{equation}
where n is any integer (including zero).
For both the Ideal Fermi gas, as well as the Ideal Bose gas (in the rotating frame of reference of a rotating ring), the energy of the states (simply kinetic energy) that is relevant in the partition function can be expressed as
\begin{equation}
E_{n}=\frac{p_{n}^2}{2m} =\frac{\hbar^{2}(n-\phi)^2}{2mR^2}
\label{eq:kin}
\end{equation}
where m is the particle mass and n is any integer from $-\infty$ to $+\infty$.
Note that for the Fermi case, this energy is in the lab frame, while for the Bose case, this energy is in the rotating frame.
In the case of the \textbf{Fermi} gas, the charged particles are in the presence of an external magnetic field parallel to the axis of the ring. The parameter $\phi$ is the ratio
\begin{equation}
\phi=\frac{\Phi}{\Phi_{0}}
\end{equation}
where $\Phi$ is the total magnetic flux through the ring and $\Phi_0$ is the flux quantum given by
\begin{equation}
\Phi_0=\frac{hc}{e}
\end{equation}
The total flux is a sum of the flux due to the external field and the flux produced by an electric current of the charged particles in the ring.
In the case of the \textbf{Bose} gas, the situation is a bit more complex. We have a ring of gas that is bounded by a wall that is rotating at a fixed angular velocity $\omega$. The rotational momentum is quantized in the inertial lab frame, and is thus given by equation (\ref{eq:mom}).
There is a corresponding quantum of velocity $v_{0}$ and of angular velocity $\omega_{0}$
\begin{equation}
v_{0}=\frac{\hbar}{mR} ~~~~~~~~~
\omega_{0}=\frac{\hbar}{mR^{2}}
\end{equation}
We assume that the system has come to thermodynamic equilibrium. As we mentioned above, the energy in the partition function is then the energy in the rotating frame. The rotational velocity in the rotating frame is given by
\begin{equation}
v_{rot}=v_{lab}-\omega R
\end{equation}
Thus the unit of quantum velocity in the rotational frame is
\begin{eqnarray}
v_{n}&=&n \frac{\hbar}{mR} - \omega R \\
\end{eqnarray}
Thus we obtain for the Bose gas equation (\ref{eq:kin}), with
\begin{equation}
\phi=\frac{\omega}{\omega_{0}}
\end{equation}
Note that the energy of a state can be written as
\begin{equation}
E_{n}=\frac{\hbar \omega_{0}}{2} (n-\phi)^2
\label{eq:kin2}
\end{equation}
Thus, $e_{0}\equiv \hbar \omega_{0}/2$ is the characteristic energy of the system.
\subsection{Electric Current in the Fermi Gas}
In the case of the Fermi gas, it is the electric current that exhibits Bloch oscillations as a function of the flux. It is given by
\begin{eqnarray}
I&=&\frac{e}{2\pi R}\sum_{n=-
\infty}^\infty v_{n}F_{n}\\
&=&\frac{e}{2\pi R}\sum_{n=-\infty}^\infty \frac{\hbar}{mR}(n-\phi)F_{n}
=\frac{e\hbar}{2\pi mR^2}\sum_{n=-\infty}^\infty (n-\phi)F_{n}
\label{eq:current}
\end{eqnarray}
where $F_{n}$ is the Fermi function
\begin{equation}
F_{n}=\frac{1}{exp[(E_{n}-\mu)/kT] +1}
\end{equation}
Here $\mu$ is the chemical potential.
From equation (\ref{eq:current}) we see that the current is a periodic function of the magnetic flux, with a period equal to the flux quantum $\Phi_{0}=hc/e$.
\subsection{Angular Momentum in the Bose Gas}
For the Bose gas we are interested in how the angular momentum $L$ depends upon the fixed angular velocity. In the lab frame it is given by
\begin{eqnarray}
L=\hbar\sum_{n=-\infty}^\infty n f_{n}
\end{eqnarray}
where $f_{n}$ is the Bose function
\begin{equation}
f_{n}=\frac{1}{exp[(E_{n}-\mu)/kT] -1}
\end{equation}
Here $\mu$ is the chemical potential.
We can separate the total angular momentum into two parts:
\begin{eqnarray}
L=L_{class}+L_{anom}
\end{eqnarray}
The first part is what we obtain in the classical limit:
\begin{eqnarray}
L_{class}= \mathcal{I} \omega =N\hbar \phi = \hbar\sum_{n=-\infty}^\infty \phi f_{n}
\end{eqnarray}
where $\mathcal{I}=NmR^{2}$ is the total moment of inertia.
The second part is the anomalous part - the angular momentum in the rotating frame - which would normally vanish since then the walls would carry the entire gas along with it.
\begin{eqnarray}
L_{anom}=\hbar\sum_{n=-\infty}^\infty (n-\phi)f_{n}
\label{eq:angmom}
\end{eqnarray}
From equation (\ref{eq:angmom}) we see that the anomalous angular momentum is a periodic function of the applied angular velocity, with a period equal to the quantum of angular velocity, $\omega_{0}=\hbar/{mR^{2}}$. Thus, the anomalous angular momentum corresponds to the above Bloch oscillations of the electric current.
\subsection{Crossover Temperatures for Oscillations}
It is clear that the discreteness of the energy levels is responsible for the observability of Bloch oscillations. As we raise the temperature, the discreteness becomes less relevant. Generally, the energy level spacing is given by
\begin{eqnarray}
\Delta E_{n}= \left[(n+1)^2-n^2\right]e_{0}=(2n+1)e_{0}
\end{eqnarray}
In the \textbf{classical regime}, the average energy per particle is $\sim kT$. The crossover temperature is determined by setting the characteristic energy level spacing equal to $kT$. The corresponding quantum number is $n=\sqrt{kT/e_{0}}$. Thus, the relevant energy level spacing is $~2e_{0}\sqrt{kT/e_{0}}=2\sqrt{kT e_{0}}$. And finally, we obtain the crossover temperature $T_{C}$ from the equation $\sqrt{kT e_{0}}=kT$:
\begin{eqnarray}
T_{C}\sim e_{0}/k
\end{eqnarray}
For \textbf{Fermions}, the characteristic energy is the Fermi energy $\epsilon_{F}$, which corresponds to a quantum number $n\sim N/2$. Thus the relevant energy level spacing is $\Delta E_{N/2} \sim Ne_{0}$. Again, the crossover temperature $T_{F}$ is determined by $\Delta E=kT$, so that it is given by
\begin{eqnarray}
T_{F}\sim Ne_{0}/k
\end{eqnarray}
In fact, it can be shown that the persistent current is given by \cite{chempot}
\begin{eqnarray}
I=N \sum_{p=1}^\infty \frac{(-1)^{Np}}{2 \pi p}\frac{2\pi ^2 pkT/Ne_{0}}{\sinh[2\pi ^2 pkT/Ne_{0}]}
\end{eqnarray}
Hence, the actual crossover temperature is better represented by
\begin{eqnarray}
T_{F}= Ne_{0}/(2 \pi ^{2} k)
\end{eqnarray}
For \textbf{Bosons}, the situation is entirely different. We know that bosons are attracted into the same state. The 3D Bose gas undergoes a Bose-Einstein condensation. While there is no condensation in 1D or 2D, there is nevertheless a relatively high occupation of the states with low energies - with quantum number n of order unity - as a result of the minus sign in the denominator of the distribution function.
We can learn a lot from an analysis of the situation at essentially absolute zero. In this case, all particles are in the ground state. Let $\alpha(\phi) \equiv -\mu (\phi)/kT$
Then we must have
\begin{equation}
f_{0}=\frac{1}{exp[e_{0} \phi^2/kT +\alpha] -1}=N
\end{equation}
For simplicity, we take $\phi=0$. Then,
\begin{equation}
\alpha =\ln(1+1/N) \sim 1/N
\end{equation}
For low temperatures, we still expect $\alpha \sim 1/N$.
We want $f_{n\neq0}$ to be of order unity for n on the order of unity but negligibly small for n not of order unity.
Since
\begin{equation}
f_{n}=\frac{1}{exp[e_{0} n^2/kT +\alpha] -1} \sim\frac{1}{exp[e_{0} n^2/kT + 1/N] -1}
\end{equation}
we must have $kT<Ne_{0}$. Thus, the crossover temperature $T_{B}$ for our Bose gas is the same as it is for the Fermi gas, namely
\begin{eqnarray}
T_{B}\sim Ne_{0}/k
\end{eqnarray}
It is more straightforward to show that for the low energy states to dominate, we must have
\begin{eqnarray}
f_{n} = \frac{1}{exp[(e_{0}(n-\phi)^2-\mu)/kT] -1} \cong \frac{1}{(e_{0}(n-\phi)^2-\mu)/kT}
\label{eq:approx}
\end{eqnarray}
for the dominant states.\\
We will define a characteristic temperature
\begin{eqnarray}
T_{B} \equiv Ne_{0}/2 \pi ^{2}k
\end{eqnarray}
The chemical potential is determined by the N, T and $\phi$ through the equation
\begin{equation}
N=\sum_{n=-\infty}^\infty f_{n} \cong \sum_{n=-\infty}^\infty \frac{1}{(e_{0}(n-\phi)^2-\mu)/kT}
\end{equation}
We will use the approximate expression for N from now on. Let us introduce the reduced \textbf{ reduced chemical potential} $m \equiv \mu /e_{0}$ and the \textbf{reduced temperature} $t \equiv T/ T_{B} $. Then it is straightforward to show that
\begin{eqnarray*}
& N\frac{e _{0}}{kT} & \equiv {\frac {{ 2 \pi }^{2}}{t}}
= \frac{1}{( {\phi}^{2}-m)} ~ + \\ & ~ & {\frac {\Psi \left( \sqrt {m}-\phi +1 \right) -\Psi \left( -\sqrt {m}-\phi+1
\right) +\Psi \left( \sqrt {m}+\phi+1 \right) -\Psi \left( -\sqrt {m}+\phi+
1 \right) }{2\sqrt {m}}}
\end{eqnarray*}
Here $\Psi(z)$ is the diGamma function \cite{GR}.
From this expression we see that the reduced chemical potential is expressible entirely in terms of the reduced angular velocity and the reduced temperature, thus confirming our choice of crossover temperature $T_{B}\equiv Ne_{0}/(2 \pi ^{2} k)$.\\
The \textbf{reduced anomalous angular momentum}, $\ell \equiv L_{anom}/N \hbar$ can be similarly expressed:
\begin{equation}
\ell=-\sqrt {m}+ \frac{t}{2 \pi ^{2}} \left( \Psi \left( \sqrt {m}+\phi+1 \right) -\Psi
\left( -\sqrt {m}-\phi+1 \right) - \frac{1}{ \sqrt{m}+\phi } \right)
\end{equation}
\section{Bloch Oscillations - Numerical Results}
Below we will summarize the numerical results we obtained using MAPLE. The procedure is to find the chemical potential given the number of particles N using the equation
\begin{eqnarray}
N=\sum_{n=-\infty}^\infty \frac{1}{exp[(e_{0}(n-\phi)^2-\mu)/kT] -1}
\label{eq:N}
\end{eqnarray}
In figure (\ref{fig:chem}) we see plots of the chemical potential as a function of the reduced angular velocity - $\phi=\omega / \omega_{0}$ - for two temperatures: $T_{B}/100$ (solid curve) and $10T_{B}$ (dashed curve). The chemical potential is expressed in units of $e_{0}$. Note that the chemical potential oscillates with great amplitude in conjunction with the Bose tendency to keep particles in the same state. This strong variation is in great contrast with the situation for Fermions, for which there is a much smaller, though sometimes important, variation that needs to be taken into account \cite{joe2} ~\cite{foot}.
\begin{figure}[htbp]
\begin{center}
\epsfig{file=ChemPot_AngVelCropped.eps, width=10cm}
\caption{Reduced Chemical Potential vs. Reduced Angular Velocity}
\label{fig:chem}
\end{center}
\end{figure}
In figure (\ref{fig:totmom}) we plot the total angular momentum in units of $N\hbar $ as a function of the reduced angular velocity $\phi$, for the temperature $T_{B}/100$ (solid curve) and for the classical regime(dash-dot curve). In figure (\ref{fig:mom}) we plot the anomalous angular momentum in units of $N\hbar $ as a function of the reduced angular velocity $\phi$, for two temperatures: $T_{B}/100$ (solid curve) and $10T_{B}$ (dashed curve).
\begin{figure}[h]
\begin{center}
\epsfig{file=TotAngMom_AngVel.eps, width=13cm}
\caption{Total Angular Momentum vs. Angular Velocity}
\label{fig:totmom}
\end{center}
\end{figure}
\begin{figure}[h]
\begin{center}
\epsfig{file=AngMom_AngVel.eps, width=13cm}
\caption{Anomalous Angular Momentum vs. Angular Velocity}
\label{fig:mom}
\end{center}
\end{figure}
\section{Josephson Oscillations}
As shown by Bloch \cite{bloch}, Josephson oscillations in a conductor can be explained in terms of the Bloch oscillations of the current in the presence of a time dependent magnetic flux. Generally,
\begin{eqnarray}
V= \frac{1}{c}\frac{\partial \Phi}{\partial t}
\end{eqnarray}
If the voltage is constant, the flux increases linearly in time, so that
\begin{eqnarray}
\phi= \frac{cVt}{\Phi_{0}} = \frac{eV}{h}t
\end{eqnarray}
Then, if the electrons remain in quasi-thermodynamic equilibrium, the current will oscillate with the Josephson frequency, given by
\begin{eqnarray}
f_{J}=\frac{eV}{\hbar}
\end{eqnarray}
\newpage
Now let us turn to the ring of an Ideal Bose gas. The corresponding experimental condition is to have a constant torque $\tau$ applied to the ring wall. We must remember that the resulting angular velocity of the wall is not the angular velocity of the gas. However, the angular velocity of the wall determines the state of state of the gas assuming, as above, that quasi-static thermodynamic equilibrium is maintained. We have with a constant torque
\begin{eqnarray}
L(\phi)=\tau t
\end{eqnarray}
Thus,
\begin{eqnarray}
\phi (t)= L^{-1} (\tau t)
\end{eqnarray}
where $L^{-1}$ is the inverse function of $L$.
To obtain the corresponding Josephson frequency we use the fact that $L_{anom}(\phi)$ is a periodic function of $\phi$:
\begin{eqnarray}
L_{anom}(\phi +1)=L_{anom}(\phi))
\end{eqnarray}
Since
\begin{eqnarray}
L_{anom}(\phi)=L(\phi)-L_{class}(\phi)=\tau t - N\hbar \phi
\end{eqnarray}
we easily find that the Josephson frequency is
\begin{eqnarray}
f_{J}=\frac{\tau}{N \hbar}
\end{eqnarray}
A plot of the angular velocity vs. time is shown in figure (\ref{fig:AngvelTime}) for a temperature $T_{B}/100$. The time axis is in units of the Josephson period and the angular velocity is in units of the quantum of angular velocity. The dash-dot curve is the classical result, for which the entire gas moves with the wall, so that the angular velocity is linear in time.
\begin{figure}[htbp]
\begin{center}
\epsfig{file=AngVel_Time.eps, width=10cm}
\caption{Angular Velocity vs. Time}
\label{fig:AngvelTime}
\end{center}
\end{figure}
Let us try to make sense of the graph. We will assume absolute zero, for which the jumps in the graph are discontinuous. First, we note that from zero to one time unit, $t_{J}=1/f_{J}$, the wall has one-half the quantum of angular velocity - that is $\omega_{_{0}} /2$ - whereas a single particle has an angular velocity that must an integral number of quanta. Furthermore, in the course of this time interval, the total change in angular momentum is $\Delta L = \tau t_{J}=N \hbar$, which corresponds to all of the particles having a single quantum of angular velocity. Therefore, during this time interval, the particles are continuously making a transition from the $n=0$ state to the $n=1$ state. At the end of the time interval, the entire gas is moving twice as fast as the walls! The subsequent behavior is then obvious and will not be discussed here.
\section{Proof of Bloch \& Josephson Oscillations in a Ring of Interacting Particles}
\textbf{THEOREM}: \emph{In the quantum regime, a ring of identical \textbf{interacting particles}, whether they be Bosons or Fermions, has an anomalous component of the angular momentum in response to a fixed angular velocity of the wall - a component that is periodic in the angular velocity, with a period $\omega_{0}$.}\\
The proof makes use of the technique used by Byers and Yang \cite{by} in proving that in the presence of a uniform magnetic field, magnetic flux is quantized in a hollow superconducting cylinder. \cite{periodiccurrent}
\\
\textbf{Lemma 1}: \\
The energy eigenvalues in the rotating frame are periodic functions of the angular velocity, with a period equal to the quantum of angular velocity.\\
\textbf{ Proof:}\\
The coordinates of the particles can be taken to be the angles $\{\theta_{j} \}$, with $j=1...N$. The Schroedinger equation in the lab frame is given by
\begin{eqnarray}
\widehat{H}\Psi=E\Psi
\end{eqnarray}
where the Hamiltonian operator $\widehat{H}$ in the inertial lab frame is given by
\begin{eqnarray}
\widehat{H}=-\frac{\hbar^{2}}{2mR^{2}}\sum_{j=1}^N\frac{\partial^{2}}{\partial\theta_{j}^{2}} + V(\{\theta_{j}\})
\end{eqnarray}
In the rotating frame of reference, the Schroedinger equation is given by
\begin{eqnarray}
\widehat{H}_{rot}\Psi_{rot}=E_{rot}\Psi_{rot}
\end{eqnarray}
where the Hamiltonian operator $\widehat{H}_{rot}$ is given by
\begin{eqnarray}
\widehat{H}_{rot}=\frac{\hbar^{2}}{2mR^{2}}\sum_{j=1}^N
(-i\frac{\partial}{\partial\theta_{j}} - \phi)^{2}
+ V(\{\theta_{j}\})
\end{eqnarray}
Note that the energy eigenvalue is a function of $\phi$: $E_{rot}=E_{rot}(\phi)$. Also, the boundary condition on the wave function is
\begin{eqnarray}
\Psi_{rot}(\theta_{i}+2\pi)=\Psi_{rot}(\theta_{i})
\end{eqnarray}
for any $\theta_{i}$.\\
Now let us define the function $\Psi_{rot}'$:
\begin{eqnarray}
\Psi_{rot}=e^{i\phi \sum_{j}\theta_{j}} \Psi_{rot}'
\end{eqnarray}
This new function satisfies the equation (note that the Hamiltonian corresponds to the inertial frame and is independent of $\phi$):
\begin{eqnarray}
\widehat{H}\Psi_{rot}'=E_{rot}(\phi) \Psi_{rot}'
\end{eqnarray}
where the energy eigenvalue depends upon $\phi$ because the boundary condition on $\Psi_{rot}'$ is
\begin{eqnarray}
\Psi_{rot}'(\theta_{i}+2\pi)= e^{-2 \pi i \phi} \Psi_{rot}'(\theta_{i})
\end{eqnarray}
for a given $\theta_{i}$, with all other $\theta_{j}$ kept fixed.\\
We then note that if $\phi$ changes by unity $\phi \rightarrow \phi +1$, the boundary condition doesn't change. Since the boundary condition determines the specific solution to the Schroedinger equation and hence the eigenvalue, we see that the eigenvalues, with subscript label $s$, satisfy $E_{rot,s}(\phi +1) = E_{rot,s}(\phi)$.\\
\textbf{QED}\\
\textbf{Lemma 2:}\\
The partition function in the rotating frame is a periodic function of the angular velocity, with a period equal to the quantum of angular velocity.\\
\textbf{Proof:}\\
This result follows automatically from the fact that the partition function in the rotating frame is given by
\begin{eqnarray}
Z_{rot}=\sum_{s}e^{- E_{rot,s}/kT}
\end{eqnarray}\\
\textbf{Lemma 3:}\\
The angular momentum in the rotating frame, $L_{rot}\equiv L_{anom}$ is given by
\begin{eqnarray}
L_{anom}= kT \frac{\partial \ln Z_{rot}}{\partial \omega}=-\frac{\partial F_{rot}}{\partial \omega}
\end{eqnarray}
where $F_{rot}$ is the free energy in the rotating frame.\cite{cfcurrent}\\
\textbf{Proof:}\\
The partition function in the rotating frame can be expressed as
\begin{eqnarray}
Z_{rot}=Tr~ e^{- \widehat{H}_{rot}/kT}
\end{eqnarray}
while the Hamiltonian in the rotating frame can be expressed as
\begin{eqnarray}
\widehat{H}_{rot}=\frac{\widehat{L}_{rot}^{2}}{2I}
+ V(\{\theta_{j}\})
\end{eqnarray}
where
\begin{eqnarray}
\widehat{L}_{rot}= \widehat{L} - I\omega =\hbar \sum_{j} (-i\frac{\partial}{\partial \theta_{j}} ) -I\omega
\end{eqnarray}
It easily follows that
\begin{eqnarray}
L_{anom}=\frac{1}{Z_{rot}}Tr~ e^{- \widehat{H}_{rot}/kT}\widehat{L}_{rot} = kT \frac{\partial \ln Z_{rot}}{\partial \omega}
\end{eqnarray}
The original theorem follows trivially since the periodicity of the partition function is passed on to the angular momentum. In addition, since the energy eigenvalues can be shown to be \emph{even} functions of the angular velocity, the angular momentum is an \emph{odd} function of the angular velocity. And finally, we should note that while we have demonstrated the periodicity of the angular velocity, there is no guarantee that its amplitude is non-vanishing.
\section{Summary and Discussion}
We have shown that a ring of an Ideal Bose gas exhibits both Bloch and Josephson oscillations at a low temperature. Let us consider a concrete numerical example to assess the feasibility (albeit at present remote) of observing these oscillations. We will assume a spin-polarized condensate of $N=10,000$ atoms of hydrogen as a Boson with low mass -- $1.7\times 10^{-24}gm$ and a ring radius of $0.1mm$. We obtain the following:
\begin{center}
$\omega_{0}=5.9$ rad/sec,~~~$e_{0}=2.9\times10^{-27}$ ergs, ~~~ $T_{B}=30~nK$.
\end{center}
It is difficult to imagine how one could observe this anomalous behavior with current experimental techniques. While there has been great progress in confinement of BE condensates, the anomalies presented here require the presence of a confining toroidal wall that can be controllably rotated.\\
Josephson oscillations of an electric current in a ring are the natural expansion of the long-known oscillations (usually referred to as "Bloch oscillations") in the velocity of electrons in a periodic lattice that are driven by an electric field $\mathcal{E}$. Here, as long as electrons remain in one band (Zener tunneling being absent) the electrons move from one end of a band to a zone edge, where they are reflected to the opposite edge. The frequency is given by $f=e\mathcal{E}a/h$, where $a$ is the lattice spacing. In the case of ring, the lattice spacing of a periodic lattice is replaced by the circumference of the ring. The potential V in the Josephson frequency $f_{J}=eV/h$ is related to the electric field through $\mathcal{E}a=V$. Similarly, the anomalous behavior presented in this paper has its counterpart in motion of atoms in a periodic lattice. In fact, T. Salger et al \cite{salger} have recently presented evidence of Bloch oscillations of a BE condensate Rb atoms moving in a periodic optical lattice. Here, the optical lattice is accelerated so as to produce an effective external force.
\newpage
|
\section{Introduction}
Recently there have been considerable interests in the investigations of
topological ordered states, and the quantum phase transitions between them.
Since topological ordered phases usually do not exhibit conventional
symmetry breaking, these phase transitions naturally can not be described by
the conventional Landau-Ginzburg paradigm. Despite of the lack of local
order parameters, tremendous progresses have been made in characterizing
topological ordered states. Properties such as ground state degeneracy,
quasiparticle statistics, existence of edge states, topological entanglement
entropy,\cite{Kitaev2,Xiaogang} and entanglement spectrum\cite{HuiLi} have
been proposed and used to distinguish different topological ordered states.
In contrast, the study of topological phase transitions is still in its
infancy, and progresses are in demand.
One dimensional quantum spin chains have been a subject of interests for
many years. It started with the famous Haldane conjecture,\cite{Haldane}
followed by the Affleck-Kennedy-Lieb-Tasaki (AKLT) construction of the
valence bond solid (VBS) states and their associated parent Hamiltonians
\cite{AKLT} Moreover, the SU(2) AKLT model has been generated by introducing
q-deformed SU(2) group,\cite{zittartz} supersymmetry,\cite{arovas} and
higher symmetric groups, such as SU(n),\cite{affleck-greiter} SP(n),\cit
{rachel}, and SO(n).\cite{Tu2} More recently a systematic method for
constructing translational invariant VBS state for the general Lie group has
been proposed.\cite{Tu}
For the $S=1$ VBS state, there is an appealing physical picture where each
spin-1 is decomposed into two ``virtual''\ spin-1/2's. Across each valence
bond, two neighboring virtual spins pair into a singlet. den Nijs and
Rommelse proposed a nonlocal string order parameter (SOP) which revealed a
hidden ``diluted antiferromagnetic order''.\cite{Nijs} Kennedy and Tasaki
found an unitary transformation that turns the nonlocal SOP to a local
ferromagnetic order parameter associated with a hidden $Z_{2}\otimes Z_{2}$
symmetry.\cite{Kennedy} However, it is extremely difficult to generalize
such a description to the cases of higher quantum integer spin chains.
In this paper, we present a model for the $S=2$ chain which exhibits two
distinct VBS states in different parameter regimes. For one of the states,
each spin-$2$ is decomposed into two virtual spin-$1$'s, and for the other
it is decomposed into two spin-$3/2$'s. These virtual spins then pair up
across every nearest neighbor bonds. In the following, we shall refer to
these two states as VBS$_{1}$ and VBS$_{3/2}$ states. For an open chain, the
ground states have 9- and 16-fold degeneracies in these two cases,
respectively. Hence these two VBS states have different topological order.
Interestingly, there is a continuous quantum phase transition between these
two VBS states. By analyzing the relation between the von Neumann
entanglement entropy and the spin-spin correlation length,\cit
{tag,pollmann-moore} we deduce the central charge associated with the
critical conformal field theory to be two. We further conjecture that the
underlying critical field theory may be described by the level-four $SU(2)$
Wess-Zumino-Witten model.
The paper is organized as follows. In Sec. II, we review the properties of
VBS$_{1}$ and VBS$_{3/2}$ states. In Sec. III, the quantum phase transition
for the above $S=2$ spin model is explored using the infinite time evolving
decimation method.\cite{Vidal} The entanglement spectrum around the phase
transitions is studied and the central charges for the underlying conformal
field theory at the critical line are determined. A summary is given in Sec.
IV.
\section{Two distinct VBS states of a spin-2 chain}
The model Hamiltonian of the spin-2 chain is proposed as
\begin{eqnarray}
H &=&\sum_{i}\left[ J_{2}P_{2}(i,i+1)+J_{3}P_{3}(i,i+1)+P_{4}(i,i+1)\right]
\notag \\
&=&\sum_{i}\left[ \frac{189J_{3}-400J_{2}+30}{420}\left( \mathbf{S}_{i
\mathbf{S}_{i+1}\right) \right. \notag \\
&&\text{ \ \ \ }-\frac{40J_{2}+7J_{3}-9}{360}(\mathbf{S}_{i}\mathbf{S
_{i+1})^{2} \notag \\
&&\text{ \ \ \ }+\frac{10J_{2}-5J_{3}+1}{180}(\mathbf{S}_{i}\mathbf{S
_{i+1})^{3} \notag \\
&&\text{ \ \ \ }\left. +\frac{20J_{2}-7J_{3}+1}{2520}(\mathbf{S}_{i}\mathbf{
}_{i+1})^{4}\right] . \label{eq:model}
\end{eqnarray
where $P_{T}(i,i+1)$ is the SU(2) symmetric operator that projects the spin
states associated with sites $i$ and $i+1$ into the total spin-$T$
multiplet. The coupling constants $J_{2}$ and $J_{3}$ are all positive. In
order to make the paper self-contained, we first review the VBS$_{1}$ and VB
$_{3/2}$ states, respectively.
\subsection{The VBS$_{1}$ state - AKLT state}
To construct this state, we view each spin-2 as a symmetric product of two
virtual spin-$1$'s. In the VBS$_{1}$ state, two neighboring virtual spin-1
form a singlet. The direct product of the spin-2 multiplets on neighboring
sites can be decomposed into a direct sum of the total spin $S=0,1,2,3,4$
multiplets. Due to the singlet pairing of the neighboring virtual spin-1's,
the total spin $S=3,4$ multiplets can not be generated. For $J_{2}=0$, the
model Hamiltonian only penalizes the $S=3,4$ two-spin states. Hence VBS$_{1}$
state is the unique ground state of Eq.(\ref{eq:model}). In an open chain,
the unpaired virtual spin-$1$'s at the two ends are free, and they give rise
to the $3\times 3=9$ fold ground state degeneracy.\cite{AKLT,Kennedy}
In order to write down the ground state wave function, we can use the
Schwinger boson representation, and the spin-$2$ operators can be expressed
as
\begin{equation}
S_{i}^{+}=a_{i}^{\dag }b_{i},\ S_{i}^{-}=b_{i}^{\dag }a_{i},\
S_{i}^{z}=(a_{i}^{\dag }a_{i}-b_{i}^{\dag }b_{i})/2,
\end{equation
with a local constraint $a_{i}^{\dag }a_{i}+b_{i}^{\dag }b=4$. Then the $S=2$
AKLT VBS ground states can be expressed in a simple form,\cite{Arovas}
\begin{equation}
|\Psi _{\mathrm{AKLT}}\rangle =\prod_{i}(a_{i}^{\dag }b_{i+1}^{\dag
}-b_{i}^{\dag }a_{i+1}^{\dag })^{2}|vac\rangle \label{EqAKLT}
\end{equation
where each $\left( a_{i}^{\dag }b_{i+1}^{\dag }-b_{i}^{\dag }a_{i+1}^{\dag
}\right) $ creates a singlet bond composed of two spin-$1/2$ between $i$ and
$i+1$ sites. Furthermore, by re-arranging the creation operators in Eq.(\re
{EqAKLT}) and combining operators with the same site operators together,
|\Psi _{\mathrm{AKLT}}\rangle $ can be written in a matrix product state
form straightforwardly,
\begin{equation}
|\Psi _{\mathrm{AKLT}}\rangle =\sum_{i_{1},i_{2},\cdots ,i_{N}=-2}^{2
\mathrm{Tr}(A^{[i_{1}]}A^{[i_{2}]}\cdots A^{[i_{N}]})|i_{1}i_{2}\cdots
i_{N}\rangle ,
\end{equation
where $\left\{ A^{[m]}\right\} $ with $m=\pm 2,\pm 1,0$ are $3\times 3$
matrixes
\begin{gather}
A^{[-2]}=\left(
\begin{array}{ccc}
0 & 0 & 0 \\
0 & 0 & 0 \\
2\sqrt{6} & 0 &
\end{array
\right) ,\text{ }A^{[-1]}=\left(
\begin{array}{ccc}
0 & 0 & 0 \\
-2\sqrt{3} & 0 & 0 \\
0 & 2\sqrt{3} &
\end{array
\right) , \notag \\
A^{[0]}=\left(
\begin{array}{ccc}
2 & 0 & 0 \\
0 & -4 & 0 \\
0 & 0 &
\end{array
\right) ,\text{ }A^{[1]}=\left(
\begin{array}{ccc}
0 & 2\sqrt{3} & 0 \\
0 & 0 & -2\sqrt{3} \\
0 & 0 &
\end{array
\right) , \notag \\
A^{[2]}=\left(
\begin{array}{ccc}
0 & 0 & 2\sqrt{6} \\
0 & 0 & 0 \\
0 & 0 &
\end{array
\right) .
\end{gather}
\subsection{The VBS$_{3/2}$ State - SO(5) symmetric state}
Instead of splitting a spin-$2$ into two virtual spin-$1$'s, one can also
split it into two spin-$3/2$'s. In the following, we shall first view the VB
$_{3/2}$ state as the AKLT state of a larger symmetry group, which is
equivalent to the $SO(5)$ symmetric matrix product state in a two-leg
electronic ladder\cite{sczhang}. Afterwards we will rephrase everything in
terms of the physical spin $SU(2)$. As pointed out in Ref.\cite{Congjun},
one can view the $\pm 3/2,\pm 1/2$ states of a spin-3/2 as the four states
of the spinor representation of $SO(5)$. Similarly one can regard the $\pm
2,\pm 1,0$ states of spin-$2$ as the five-dimensional vector irreducible
representation (IR) of $SO(5)$. Analogous to decomposing a spin-$1$ vector
IR of $SU(2)$ into two virtual spin-$1/2$'s spinor IR of $SU(2)$, we can
view the vector IR as the symmetric component of the tensor product of two
virtual spinor IR's, i.e.,
\begin{equation}
\underline{4}\otimes \underline{4}=\underline{1}\oplus \underline{5}\oplus
\underline{10}. \label{decom1}
\end{equation
The numerals are the dimensions of the $SO(5)$ IR's. The tensor product of
two $\underline{5}$'s on adjacent sites decomposes into
\begin{equation}
\underline{5}\otimes \underline{5}=\underline{1}\oplus \underline{10}\oplus
\underline{14}. \label{decom2}
\end{equation
Comparing expressions of Eq.(\ref{decom1}) and Eq.(\ref{decom2}), we view
Eq.~(\ref{decom1}) as the tensor product of two neighboring virtual spins
after their respective partners have form $SO(5)$ singlet with other virtual
spins. Then one can find that $SO(5)$ singlet $\underline{1}$ and the
antisymmetric $\underline{10}$ appear in the decomposition but the symmetric
$\underline{14}$ is absent. Therefore, if $H=\sum_{i}P_{14}(i,i+1)$, the VBS
_{3/2}$ state where neighboring virtual $\underline{4}$'s pair into $SO(5)$
singlet will be the ground state. In an open chain, due to the unpaired free
$SO(5)$ spinors at two ends, the ground states are $4\times 4=16$ fold
degenerate. A clear and detailed argument of this degeneracy is given is
section III(B). Furthermore, the projection operator $P_{\underline{14
}(i,i+1)$ can be expressed in terms of the $SO(5)$ generators\cite{Tu2}
\begin{equation}
P_{\underline{14}}(i,j)=\frac{1}{2}\sum_{1\leq a<b\leq
5}L_{i}^{ab}L_{j}^{ab}+\frac{1}{10}(\sum_{1\leq a<b\leq
b}L_{i}^{ab}L_{j}^{ab})^{2}+\frac{1}{5}.
\end{equation
Because the physical spin is $SU(2)$, which is a subgroup of $SO(5)$, each
IR of $SO(5)$ must decompose into an integral number of $SU(2)$ multiplets.
Thus the $\underline{14}$ discussed above must be expressible as the direct
sum of $SU(2)$ IR obtained by decomposing the direct product of two $S=2$
multiplets. Since the 14-dimensional IR is symmetric upon the exchange of
site indices, it must only contain even-spin $SU(2)$ multiplets. A simple
calculation shows that $\underline{14}\rightarrow S=2\oplus S=4.$
Consequently, $\sum_{i}P_{14}(i,i+1)$ reduces to Eq.(\ref{eq:model}) with
J_{3}=0$, which is first given by Tu, Zhang, and Xiang.\cite{Tu2}
Moreover, the $SO(5)$ generators can be represented by $L^{ab}=\sum_{\alpha
,\beta }\psi _{\alpha }^{\dag }\Gamma _{\alpha \beta }^{ab}\psi _{\beta }$,
where $\psi _{j,\alpha }^{\dag }$ creates a spin-$3/2$ fermion with spin
index $\alpha =\pm 3/2,\pm 1/2$, $\Gamma ^{a},(a=1,\cdots ,5)$ are the
4-dimensional Dirac $\Gamma $ matrices, and $\Gamma ^{ab}=\frac{i}{2}[\Gamma
^{a},\Gamma ^{b}]$. By using the above representations, the VBS$_{3/2}$
state can be written as\cite{Tu2}
\begin{equation}
|\mathrm{VBS}_{3/2}\rangle =\prod_{j}\mathcal{P}_{S=2}(j)(\sum_{\alpha \beta
}\psi _{j,\alpha }^{\dag }\mathcal{R}_{\alpha \beta }\psi _{j+1,\beta
}^{\dag })|\text{vac}\rangle
\end{equation
where $\mathcal{P}_{S=2}(j)$ is the spin-quintet projector and $\sum_{\alpha
\beta }\psi _{j,\alpha }^{\dag }\mathcal{R}_{\alpha \beta }\psi _{j+1,\beta
}^{\dag }$ is an $SO(5)$ invariant valence bond singlet creation operator.
\mathcal{R}$ is the $SO(5)$ invariant matrix. This VBS$_{3/2}$ state can
also be expressed as a matrix product states (MPS)
\begin{equation}
|\mathrm{VBS}_{3/2}\rangle =\sum_{i_{1},..,i_{N}=-2}^{2}\mathrm{Tr
(B^{[i_{1}]}B^{[i_{2}]}\cdots B^{[i_{N}]})|i_{1}i_{2}\cdots i_{N}\rangle ,
\end{equation
where $\{B^{[m]}\}$ with $m=0,\pm 1,\pm 2$ are given by the following
4\times 4$ matrice
\begin{gather}
B^{[-2]}=\left(
\begin{array}{cccc}
0 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 \\
\sqrt{2} & 0 & 0 & 0 \\
0 & \sqrt{2} & 0 &
\end{array
\right) ,\text{ }B^{[-1]}=\left(
\begin{array}{cccc}
0 & 0 & 0 & 0 \\
-\sqrt{2} & 0 & 0 & 0 \\
0 & 0 & 0 & 0 \\
0 & 0 & \sqrt{2} &
\end{array
\right) , \notag \\
B^{[0]}=\left(
\begin{array}{cccc}
1 & 0 & 0 & 0 \\
0 & -1 & 0 & 0 \\
0 & 0 & -1 & 0 \\
0 & 0 & 0 &
\end{array
\right) ,\text{ }B^{[1]}=\left(
\begin{array}{cccc}
0 & \sqrt{2} & 0 & 0 \\
0 & 0 & 0 & 0 \\
0 & 0 & 0 & -\sqrt{2} \\
0 & 0 & 0 &
\end{array
\right) \notag \\
B^{[2]}=\left(
\begin{array}{cccc}
0 & 0 & \sqrt{2} & 0 \\
0 & 0 & 0 & \sqrt{2} \\
0 & 0 & 0 & 0 \\
0 & 0 & 0 &
\end{array
\right) .
\end{gather}
\section{Continuous quantum phase transitions}
In the model Hamiltonian Eq.(\ref{eq:model}), for $J_{2}=0$, the ground
state of the model is the VBS$_{1}$ state,\cite{AKLT} while $J_{3}=0$ it
corresponds to the VBS$_{3/2}$ state.\cite{Tu2} When both $J_{2}$ and $J_{3}$
are non-zero, the model is no longer exactly solvable. Then, we expect that
a quantum phase transition may be reached by adjusting the value of
J_{3}/J_{2}$. To our knowledge, this is one of the few microscopic models,
exhibiting continuous quantum phase transitions between two distinct
topological ordered phases.
\subsection{Ground state phase diagram}
In any one-dimensional quantum spin systems, the corresponding ground states
can always be simulated by the wave functions in the MPS form, as the area
law can be easily satisfied. To study the ground state properties of the
general Hamiltonian Eq.(\ref{eq:model}), we thus propose a MPS with a finite
local matrix dimension to approximate the ground state $|\psi _{g}\rangle $,
\begin{equation}
|\Psi _{g}\rangle =\sum_{\cdots m_{i}m_{i+1}\cdots }\mathrm{Tr}(\cdots
\Gamma ^{m_{i}}\Lambda \Gamma ^{m_{i+1}}\Lambda \cdots )|\cdots
m_{i}m_{i+1}\cdots \rangle , \label{gr}
\end{equation
where $m_{i}=-2,-1,0,1,2$ is the spin quantum number of $S_{i}^{z}$,
\left\{ \Gamma ^{m_{i}}\right\} $ is a set of $D\times D$ dimensional
matrices with bond dimension $D$ corresponding to number of states kept in
density matrix renormalization group, and $\Lambda $ is also a $D\times D$
dimensional non-negative diagonal matrix with its matrix elements $\lambda
_{\alpha }$, satisfying the normalization condition $\sum_{\alpha }\lambda
_{\alpha }^{2}=1$. The trace gives the superposition coefficients of the
Hilbert space basis. The local matrices $\left\{ \Gamma ^{m}\right\} $ and
\Lambda $ are set to be identical on different sites due to the translation
invariance.\cite{TransInvar}
The spirit of the infinite time evolving block decimation algorithm (iTEBDA)
\cite{Vidal} is to do the following evolution in imaginary time
\begin{equation}
\lim_{\tau \rightarrow \infty }\frac{\exp (-H\tau )|\Psi _{0}\rangle }
\left\| \exp (-H\tau )|\Psi _{0}\rangle \right\| }=\lim_{N\rightarrow \infty
}\frac{\left( \exp (-H\varepsilon )\right) ^{N}|\Psi _{0}\rangle }{\left\|
\left( \exp (-H\varepsilon )\right) ^{N}|\Psi _{0}\rangle \right\| },
\end{equation
where $|\Psi _{0}\rangle $ is an arbitrary random initial state having
nonzero overlap with $|\psi _{g}\rangle $, $\tau $ is imaginary time, and
\varepsilon $ is a small interval satisfying $\varepsilon N=\tau $. It is a
projection method with high efficiency: as $\tau $ increases, the weight of
ground state in evolved state $\exp (-H\tau )|\Psi _{0}\rangle $ grows
exponentially. Next we split the Hamiltonian into two non-commutative parts,
\begin{equation}
H=\sum_{i\mathrm{=odd}}\hat{h}(i,i+1)+\sum_{i=\mathrm{even}}\hat{h
(i,i+1)=H_{odd}+H_{even},
\end{equation
where all the local two-body bond operators $\hat{h}(i,i+1)$ commute with
one another in each part $H_{odd}$ or $H_{even}$. Then Suzuki-Trotter
decomposition is used to divide the respective evolution in sequential
order,
\begin{equation}
\exp (-H\varepsilon )|\Psi _{0}\rangle =\exp (-\varepsilon H_{odd})\exp
(-\varepsilon H_{even})+O(\varepsilon ^{2})
\end{equation
Owing to the fact that $\exp (-\varepsilon H_{odd})$ and $\exp (-\varepsilon
H_{even})$ compose of commutating evolution operators, all the bond
evolution in each part can be implemented simultaneously, which is
compatible with the translation invariance of MPS. After each evolution
step, the domain of local matrices increases by at least one site and the
number of different local matrices increases by a factor at least $5$. Then
a singular value decomposition and bond dimension truncation are introduced
to keep the MPS in the form given by Eq.(\ref{gr}) and bond dimension fixed
in each evolution step, the detailed can be found in Ref. \cite{Vidal}.
Now we show how to calculate the ground state energy density. First we
construct the transfer matrix
\begin{equation}
G=\sum_{m=-2}^{2}\Gamma ^{m}\Lambda \otimes \left( \Gamma ^{m}\right) ^{\ast
}\Lambda \label{transfer}
\end{equation
and compute its dominant eigenvalue $\eta _{1}$, as well as the associated
right and left eigenvectors $|r_{1}\rangle $ and $|l_{1}\rangle $. Then the
matrix used to compute the energy expectation value is constructed as
\begin{equation}
G_{E}=\sum_{p,q,s,t}\langle p,q|\hat{h}|s,t\rangle \Gamma ^{p}\Lambda \Gamma
^{q}\Lambda \otimes (\Gamma ^{s})^{\ast }\Lambda (\Gamma ^{t})^{\ast
}\Lambda . \label{energytransfer}
\end{equation
where $\langle p,q|$ and $|s,t\rangle $ are the wave functions defined by
local Hilbert space of two adjacent sites and with the spin quantum numbers
p,q,s,t=-2,-1,0,1,$ $2$. The ground state energy per site $E_{g}$ can be
calculated by
\begin{equation}
E_{g}=\lim_{N\rightarrow \infty }\frac{tr\left( G^{N-2}G_{E}\right) }
tr\left( G^{N}\right) }=\frac{\langle l_{1}|G_{E}|r_{1}\rangle }{\eta
_{1}^{2}\langle l_{1}|r_{1}\rangle }.
\end{equation}
For a fixed $J_{2}$, we calculate $E_{g}$ as a function of $J_{3}$. The
quantity $E_{g}$ and its first order derivative with respect to $J_{3}$ are
finite and continuous, which are displayed in FIG.\ref{energy} and FIG.\re
{firstderiv}. However, the second derivative of $E_{g}$ with respect to
J_{3}$ exhibits divergence as $J_{3}$ is tuned to a certain critical value.
This is similar to the specific heat divergence in classical phase
transitions. Such a behavior is shown in FIG.\ref{secderiv} for several
typical values of $J_{2}$. We thus conclude that the system undergoes a
second-order phase transition at zero temperature. By determining the
positions of the critical points, we thus derive the zero temperature phase
diagram in FIG.\ref{phasediag}.
\begin{figure}[tbp]
\centering \includegraphics[width=3in]{energy.eps}
\caption{(Color online) Ground state energy density varies as a function of
J_{3}$ for fixed $J_{2}$. The local matrix dimension $D$ is set to be $300$
in the calculation.}
\label{energy}
\end{figure}
\begin{figure}[tbp]
\centering \includegraphics[width=3in]{firstderiv.eps}
\caption{(Color online) First derivative of the ground state energy density
varies as a function of $J_{3}$ for fixed $J_{2}$.}
\label{firstderiv}
\end{figure}
\begin{figure}[tbp]
\includegraphics[width=3in]{secondderiv.eps}
\caption{(Color online) The second order derivative of the ground state
energy density varies as a function of $J_{3}$ for fixed $J_{2}$. }
\label{secderiv}
\end{figure}
\begin{figure}[tbp]
\includegraphics[width=3in]{phasediagram.eps}
\caption{Ground state phase diagram of the model Hamiltonian Eq.(\re
{eq:model}).}
\label{phasediag}
\end{figure}
Moreover, the calculations of the spin-spin correlation length and
entanglement entropy also show a singular behavior and provide further
evidence of the second order phase transition of the model. In FIG.\re
{correlengthentangleentropy}, the numerical results of both spin-spin
correlation length and entanglement entropy are depicted as a function of
J_{3}$ for fixed $J_{2}$. At the critical point, the spin-spin correlation
length is divergent, and the entanglemnt entropy shows a cusp. Away from the
critical point, an extrapolation of correlation length and entanglement
entropy can be calculated as $D$ goes to infinity, indicating that both of
them saturate to finite values. Due to the finite spin-spin correlation
length and the form of the model Hamiltonian in terms of projection
operators, it is straightforward to prove the existence of excitation ga
\cite{hastings} and to identify two phases separated by the critical line in
the $J_{2}-J_{3}$ phase diagram as VBS$_{1}$ and VBS$_{3/2}$ states,
respectively. The\emph{\ }computational methods of entanglement entropy and
correlation length are explained in detail in the following sections.
\begin{figure}[tbp]
\centering \includegraphics[width=3in]{correlengthentangleentropy.eps}
\caption{(Color online) (a) The spin-spin correlation length varies as a
function of $J_{3}$ for two typical values of $J_{2}$. (b) The entanglement
entropy varies as a function of $J_{3}$ for two typical values of $J_{2}$.
The local matrix dimension $D$ is fixed at $300$ in this calculation.}
\label{correlengthentangleentropy}
\end{figure}
\subsection{Entanglement spectrum across the transition}
Li and Haldane\cite{HuiLi} have recently proposed that entanglement spectrum
(ES), i.e., the minus logarithms of the eigenvalues of a reduced density
matrix, can be used to characterize topological order. If there is an
entanglement gap separating the low-lying ES and the upper parts, then one
can find a one-to-one correspondence between the low-lying ES and the low
energy spectrum of individual edge excitations. In particular, the lowest
level of the entanglement spectrum for a topological ordered state should be
degenerate. When the state changes from VBS$_{3/2}$ to VBS$_{1}$ by tuning
the ratio of $J_{3}/J_{2}$, how does the topological order changes in this
process, especially when crossing the phase transition point? We will try to
use entanglement spectrum as a probe to partially answer this question.
If the MPS in Eq.(\ref{gr}) is in ``canonical form'', then upon dividing the
system into left and right parts the ground state wave function should
become
\begin{equation}
|\Psi _{g}\rangle =\sum_{\alpha }\lambda _{\alpha }|\Phi _{\alpha
}^{L}\rangle |\Phi _{\alpha }^{R}\rangle , \label{partition}
\end{equation
where $\{|\Phi _{\alpha }^{L}\rangle ,\alpha =1,2,\cdots ,D\}$ and $\{|\Phi
_{\alpha }^{R}\rangle ,\alpha =1,2,\cdots ,D\}$ are orthogonal basis states
of the left and right semi-infinite chain,
\begin{equation}
\langle \Phi _{\alpha }^{L}|\Phi _{\beta }^{L}\rangle =\langle \Phi _{\alpha
}^{R}|\Phi _{\beta }^{R}\rangle =\delta _{\alpha \beta }
\label{canonical_global}
\end{equation
It can be shown that the canonical condition Eq.(\ref{canonical_global})
imposes the following constraint\cite{Perez} on the $\Gamma ^{m}$ and
\Lambda $ in Eq.(\ref{gr}),
\begin{equation}
\sum_{m}\Gamma ^{m}\Lambda ^{2}(\Gamma ^{m})^{\dag }=\sum_{m}(\Gamma
^{m})^{\dag }\Lambda ^{2}\Gamma ^{m}=\mathbf{I}_{D^{2}\times D^{2}}.
\label{canonical_local}
\end{equation
In general, a MPS has a gauge freedom, because the local matrixes can be the
same up to a similarity transformation. In canonical form, it is extremely
easy and straightforward to write down the entanglement spectra $P_{\alpha
}=-\log (\lambda _{\alpha }^{2})$ . In our calculation, we perform the
canonical transformation\cite{Orus} explicitly at the end of iTEBDA and
obtain the MPS in its canonical form.
\begin{figure}[tbp]
\centering \includegraphics[width=3.2in]{entanglespectrum.eps}
\caption{(Color online) The thirty-sixth lowest values of the entanglement
spectra for the VBS$_{3/2}$ and VBS$_{1}$ phases on two sides of the
critical points. (a) $(J_{2},J_{3})=(0.1,0.029)$, (b)
(J_{2},J_{3})=(1.0,0.25)$, (c) $(J_{2},J_{3})=(3.0,0.53)$. A light gray
vertical line is put on the critical point to separate the two phases. Each
entanglement spectrum is represented by a small cross, and the degenerate
spectra are spatially staggered a little bit in horizontal direction to
distinguish and count them. In this calculations, the local matrix dimension
$D$ is set to be $300$.}
\label{EntangleSpectra}
\end{figure}
The 36 lowest values of ES are plotted in FIG.\ref{EntangleSpectra} for
J_{2}=0.1,1.0$ and $3.0$. It can be clearly seen that, for the VBS$_{3/2}$
state, the lowest entanglement eigenvalue is four-fold degenerate in one to
one correspondence with the four-fold degenerate edge states; while for the
VBS$_{1}$ state, the lowest eigenvalue of the ES is three-fold degenerate.
Above this degenerate eigenvalues there exists a large gap, so the
degenerate levels are protected topologically. So such a calculation of the
ES can be used to confirm that both VBS$_{1}$ and VBS$_{3/2}$ are really
topological ordered states, as well as that the ground state degeneracy of a
long enough open chain are really $16$ and $9$ respectively. When
approaching the critical points, the degeneracies are gradually lifted but
the gap still survives. However, in the vicinity of quantum critical points
(J_{2},J_{3})=(0.1,0.029),(1.0,0.25),(3.0,0.54)$, the degeneracies in the ES
no longer exist and the large gaps between the degenerate lowest level and
the higher levels are no longer present. Due to the expected finite-size
effect near the critical region, so far we can not simply conclude that the
topological order is destroyed completely at the critical points. Recently a
partition with a very non-local real space cut has been proposed\cit
{Arovas2}, and some new light has shed on using ES to detect non-local
orders in gapless spin chains. However, it still needs further investigation
to clarify this question and will be done in future works.
\subsection{Central charge on the critical line}
At the critical points, the system should be described by conformal
invariant quantum field theories. For such theories the central charge
encodes information about the universality class. According to conformal
field theory, the von Neumann entanglement entropy should diverge
logarithmically with the correlation length\cite{Calabrese},
\begin{equation}
S_{e}=\frac{c}{6}\ln (\xi )+S_{0},
\end{equation
where $S_{e}$ is the entanglement entropy between two semi-infinite parts of
a whole chain, $c$ is the central charge, $\xi $ is spin-spin correlation
length in units of lattice spacing, and $S_{0}$ is a non-universal constant.
For a MPS in canonical form, entanglement entropy can be calculated easil
\cite{tag,pollmann-moore}:
\begin{equation}
S_{e}=-\sum_{\alpha }\lambda _{\alpha }^{2}\ln \lambda _{\alpha }^{2}
\label{entangleentropy}
\end{equation
where $\lambda _{\alpha }$ are the coefficients in Eq.(\ref{partition}).
The spin correlation length can be deduced by two points spin-spin
correlation function as
\begin{widetext}
\begin{eqnarray}
\label{correlation} \lim_{\left\vert i-j\right\vert \rightarrow
\infty }\lim_{N\rightarrow \infty }\langle S_{i}^{z}S_{j}^{z}\rangle
&=&\lim_{\left\vert i-j\right\vert \rightarrow \infty
}\lim_{N\rightarrow \infty }\frac{tr\left( G^{N-\left\vert
i-j\right\vert -1}G_{z}G^{\left\vert i-j\right\vert
-1}G_{z}\right) }{tr\left( G^{N}\right) } \\
&=&\lim_{\left\vert i-j\right\vert \rightarrow \infty
}\frac{\left\vert \langle l_{1}|G_{z}|r_{1}\rangle \right\vert
^{2}}{\eta _{1}^{2}\langle l_{1}|r_{1}\rangle }+\frac{\langle
l_{1}|G_{z}|r_{2}\rangle \langle
l_{2}|G_{z}|r_{1}\rangle }{\eta _{1}\eta _{2}\langle l_{1}|r_{1}\rangle
\left( \frac{\eta _{2}}{\eta _{1}}\right) ^{\left\vert
i-j\right\vert } \nonumber
\end{eqnarray}
\end{widetext}where transfer matrix $G$, eigenvalues $\eta _{1}$, and
eigenvectors $\langle l_{1}|$ and $|r_{1}\rangle $ have the same definition
as in Eq.(\ref{transfer}). $\eta _{2}$ is the second largest magnitude
eigenvalue of $G$ and $\langle l_{2}|$ and $|r_{2}\rangle $ are
corresponding left and right eigenvectors. Similar to $G_{E}$ in Eq.(\re
{energytransfer}), $G_{z}$ is defined by
\begin{equation}
G_{z}=\sum_{p,q}\langle p|\hat{S}^{z}|q\rangle \Gamma ^{p}\Lambda \otimes
(\Gamma ^{q})^{\ast }\Lambda .
\end{equation
Generally, for a state without spin long-range order, $\langle
l_{1}|G_{z}|r_{1}\rangle =0$, the two points correlation function can be
written as
\begin{equation}
\lim_{\left| i-j\right| \rightarrow \infty }\lim_{N\rightarrow \infty
}\langle S_{i}^{z}S_{j}^{z}\rangle ~~\sim ~~e^{-\frac{|i-j|}{\xi }},
\end{equation
where the correlation length $\xi =1/\log (|\eta _{1}/\eta _{2}|)$.
\begin{figure}[tbp]
\centering \includegraphics[width=3.4in]{centralcharge.eps}
\caption{(Color online) The scaling behavior of the entanglement entropy
S_{e}$ as the correlation length $\protect\xi $ in the vicinity of different
critical points. (a) $(0.1,0.029)$, (b) $(1.0,0.25)$, (c) $(2.0,0.42)$, (d)
(3.0,0.54)$. The black solid line is as a reference with a central charge
c=2$ and the red squares are numerical results. In our calculation, the
local matrix dimension increases from $12$ to $600$.}
\label{CentralCharge}
\end{figure}
Using Eq.(\ref{entangleentropy}) we have calculated $S_{e}$ and $\xi $ in
the vicinity of several different critical points on the phase boundary of
FIG.\ref{phasediag}. The associated scaling relation between $S_{e}$ and
\xi $ are shown in FIG.\ref{CentralCharge}. Although there may be some
deviations at small $\xi $, $S_{e}$ tends to lie on the $c=2$ line for large
$\xi $. The fact that all the central charges are approximately equal to $2$
implies a single fixed point governing the critical behavior of the entire
phase transition line. So far the conformal field theory with $c>1$ can not
be classified systematically, and therefore to determine the corresponding
conformal field theory of a fixed line with $c=2$ might be worth attempting.
Here we present some conjectures deduced from the conformal field theory
kinematics. According to the central charge value and the constituents of
the VBS$_{1}$ and VBS$_{3/2}$ states, the level-four SU(2)
Wess-Zumino-Witten model is the most possible effective field theory and the
conformal weight of the primary field given by\cite{Zamolodchikov} $\Delta
^{(j)}=j(j+1)/6$ with $j=0,1/2,1,3/2,2$. Actually, the level-four $SU(2)$
Wess-Zumino-Witten model can be regarded as the effective critical field
theory of the following spin $S=2$ antiferromagnetic Takhtajan-Babujian mode
\cite{Alcaraz}:
\begin{eqnarray}
H &=&J\sum_{i}\left[ -\frac{1}{4}+\frac{13}{48}(\mathbf{S}_{i}\mathbf{S
_{i+1})+\frac{43}{864}(\mathbf{S}_{i}\mathbf{S}_{i+1})^{2}\right. \notag \\
&&\text{ \ \ \ \ }\left. -\frac{5}{432}(\mathbf{S}_{i}\mathbf{S}_{i+1})^{3}
\frac{1}{288}(\mathbf{S}_{i}\mathbf{S}_{i+1})^{4}\right] ,
\end{eqnarray
which can be written in terms of the projection operators a
\begin{eqnarray}
H &=&J\sum_{i}\left[ J_{1}P_{1}(i,i+1)+J_{2}P_{2}(i,i+1)\right. \notag \\
&&+J_{3}P_{3}(i,i+1)+P_{4}(i,i+1)
\end{eqnarray
with $J_{1}=\frac{12}{25}$, $J_{2}=\frac{18}{25}$ and $J_{3}=\frac{22}{25}$.
Compared to the model Hamiltonian Eq.(\ref{eq:model}), there appears an
additional interaction term $P_{1}(i,i+1)$ with the largest interaction
strength $J_{1}=\frac{12}{25}$. By calculating the entanglement spectrum, we
find that this critical point just lies on the boundary between the AKLT and
a dimerization phases. For the dimerization phase, the entanglement spectrum
show an even-odd difference from the topological ordered phase, i.e., if the
bipartition is done at a bond connecting left even and right odd sites, the
lowest spectrum is $5$ fold degenerate, otherwise the lowest spectrum is
non-degenerate and has a big gap with the upper part. This even-odd
difference indicates the existence of dimerization phase. The relation
between this critical point and the $SO(5)$-AKLT critical line can be
understood as follows. When we fix $J_{1}=\frac{12}{25}$ and $J_{2}=\frac{1
}{25}$, $J_{3}$ varies from $0$ to $\frac{22}{25}$, the system evolves from
the $SO(5)$ symmetric phase to the dimerization phase, and then the AKLT
phase for $J_{3}>\frac{22}{25}$. Moreover, for the fixed value of $J_{2}$
the dimerization region shrinks and finally disappears when $J_{1}$ is
decreased, which is compatible with our ground state phase diagram.
Therefore, we expect that there exists a crossover flow from the fixed line
of the transition between the $SO(5)$-AKLT phases to the $S=2$
antiferromagnetic Takhtajan-Babujian model.
\section{Conclusion}
In summary, we propose a one-dimensional spin-2 Hamiltonian, which exhibits
two topologically distinct VBS states in different solvable limits. By using
the infinite time evolving block decimation algorithms, we have studied the
quantum phase transition between them and determined the central charge to
be $c=2$. Of course, continuous phase transition between topological phases
characterized by different number of edge states is known. For example, by
tuning the coefficient of the topological term in the $SO(3)/SO(2)$
non-linear $\sigma $ model, it is possible to induce phase transition
between VBS states associated with \textit{different spin values}. The
transition studied in this paper is very different. It takes place between
two topologically distinct VBS states associated with the \textit{same} spin
value. We are not aware of any previous study of this type of phase
transition.
\begin{acknowledgments}
The authors are grateful to Dr. Hong-Hao Tu for stimulating discussions and
earlier collaborations. We acknowledge the support of NSF of China and the
National Program for Basic Research of MOST-China. DHL was supported by DOE
grant number DE-AC02-05CH11231.
\end{acknowledgments}
\textit{Note Added} After we submitted the original version of this
manuscript for publication, a paper\cite{jiang} concerning with the similar
issue by different numerical density matrix renormalization group method on
a finite length of chain appeared on the archive, where the authors claimed
the existence of dimerization phase separated the VBS$_{1}$ and VBS$_{3}/2$
phases in the ground state phase diagram. However, if the ground state
energy density and its second-order derivative do not show any singularity
as the coupling parameters approach to the boundary of the ''dimerization
phase'' on both sides, no quantum phase transition can occur, and the claim
of dimerization phase existence is not reliable.
|
\section{Introduction}
Frequency standards have achieved an unprecedented success in
experimental demonstrating accuracies of $4\times10^{-16}$ with a
cesium microwave fountain clock ~\cite{Santarelli} and
$1.9\times10^{-17}$ with an ion optical clock~\cite{Schneider,
Rosenband}. For optical frequency standards based on neutral atoms,
in order to effectively increase the interrogation time, Katori
proposed to utilize optical lattice trap formed with a magic
wavelength trapping laser~\cite{Katori,Takamoto}. This clever
technique greatly enhanced established high-accuracy optical
frequency standard with neutral Sr atom to an accuracy of $1
\times10^{-16}$ ~\cite{Ye,Blatt,Ludlow}. Different optical clock
schemes based on Ca~\cite{Wilpers}, Yb~\cite{Barbaer,Barber} atoms
trapped with magic wavelength lasers have been proposed.
Optical trap with a far off-resonant laser is a very useful tool for
the confinement of cold atoms. Nevertheless, for the precision of
clock transitions in frequency standards, light shift due to
trapping laser has to be avoided. Thus the wavelength of the
trapping laser should be tuned to a region where the light shift for
the clock transition is eliminated, that means the light shifts of
the two clock transition states cancel each other. The wavelength
$\lambda$ is called magic wavelength~\cite{Katori,Takamoto}.
Recently, cesium primary frequency standard with atoms trapped in an
optical lattice with a magic wavelength was
suggested~\cite{Flambaum,CPL}, and possible magic wavelengths for
clock transitions in aluminium and gallium atoms were also
calculated~\cite{Beloy}.
In contrast to the above mentioned magic wavelengths for optical
clock transitions and microwave clock transitions, here we
investigate magic wavelengths for terahertz clock transitions.
Absolute frequency standards in the terahertz domain with fine
structure transition lines of the Mg and Ca metastable triplet
states were first proposed in 1972 by Strumia~\cite{Strumia}. After
more than twenty years of continuing improvement, a frequency
standard based on the $^{3}P_{1} - ^{3}P_{0}$ Mg transition and
thermal atoms in a beam has reached an uncertainty of
$1\times10^{-12}$ ~\cite{Godone, Levi}. However, these potential
terahertz transitions for high-resolution clock references have
never been experimentally investigated with laser cooled or laser
trapped atoms.
In this paper, we present our most recent calculation of trapping
laser magic wavelengths for Sr, Ca and Mg atoms, considering
different possible clock transitions between metastable triplet
states $^{3}P$. Accurate terahertz clocks could then be built based
on such atoms which are cooled and trapped in an optical lattice.
\section{Theoretical description}
\begin{figure}[t] \centering
\includegraphics[width=10cm]{fig1.eps}
\caption{(Color online) Simplified diagram of the lowest energy
levels for alkaline-earth atoms and some possible laser couplings. }
\end{figure}
For alkaline-earth atoms, two valence electrons result in two series
of atomic energy levels as the electron spins can be parallel
(triplet states) or anti parallel (singlet state). The energy
diagram can be simplified as shown in Fig. 1. The ground state is
$^{1}S_{0}$, and the lowest excited states $nsnp$ are $^{1}P_{1}$
and $^{3}P_{J}$ which can be divided into three fine structure
sublevels $^{3}P_{2}$,$^{3}P_{1}$,$^{3}P_{0}$. For the $^{3}P_{J}$
state, transitions to higher states can be divided into three groups
: $^{3}P_{J}-^{3}S_{1}$,$^{3}P_{J}-^{3}P_{J}$ and
$^{3}P_{J}-^{3}D_{J}$.
By second-order perturbation theory, the energy shift
$U_{i}(\omega,p,m_i)$ of atomic state $|i\rangle$ with energy
$E_{i}$ and Zeeman sublevel $m_{i}$, which is induced by a trapping
laser field with frequency $\nu$=$\omega/2\pi$, polarization $p$,
and irradiance intensity $I$ can be expressed as
$U_i(\omega,p,m_i)$=$-\alpha_i(\omega,p,m_i)I/2\epsilon_0 c$ with
the induced polarizability $\alpha_i$.
The polarizability can be calculated by summing up the contributions
from all dipole interactions between the fine structure state
$|i\rangle$ and $|k\rangle$ with the Einstein coefficient $A_{Jki}$
(spontaneous emission rate for $E_k$$>$$E_i$), Zeeman sublevels
$m_i, m_k$ and transition frequency
$\nu_{Jki}$=$\omega_{Jki}/2\pi$~\cite{grimm, Carsten},
\begin{eqnarray}
\alpha_i=6\pi c^3\epsilon_0
\sum\limits_{k,m_k}\frac{A_{Jki}(2J_k+1)}{\omega^2_{Jki}(\omega^2_{Jki}-\omega^2)}\left(\begin{array}{ccc}
J_i & 1 & J_k\\m_i &p &-m_k\end{array}\right)^2
\end{eqnarray}
where
\begin{eqnarray}
A_{Jki} = \frac{e^2}{4\pi\epsilon_0}\frac{{4\omega
_{Jki}^3}}{{3\hbar {c^3}}}\frac{1}{{2{J_k} + 1}}|\left\langle
{{\beta_k}{J_k}} \left\| D \right\| {{\beta _i}{J_i}}
\right\rangle|^2
\end{eqnarray}
Here e is the electron charge, $\hbar\omega_{Jki}$ is the energy
difference between fine structure states $|k\rangle$ and
$|i\rangle$, $\beta$ denotes other quantum numbers of the state, and
${\left\langle {{\beta_k}{J_k}} \left\| D \right\| {{\beta_i}{J_i}}
\right\rangle } $ is the dipole reduced matrix element. The
expression in large parentheses in Eq.(1) denotes a $3J$ symbol
which describes the selection rules and relative strength of the
transition depending on the involved angular momenta $J$, the
projection $m$, and the polarization $p$.
If we know $\omega_{Jki}$ and $A_{Jki}$ in Eq.(1), we can get the
polarizability $\alpha_i$. However, typically the literature gives
the total transition rate $A_T$ from a given excited state to the
fine structure manifold states below. So we need establish the
relation
\begin{tablehere
\begin{scriptsize}
\begin{widetext
\caption{Sr element: Transition Wavenumbers (WN)($cm^{-1}$)
corresponding to $\omega_{Jki}$, Einstein Coefficients for fine
structure states $A_J(\times10^6s^{-1})$ and Total
$A_T(\times10^6s^{-1})$ for fine structure states manifold,
Correction Factors $\boldsymbol{\zeta}$. The Wavenumber data
originate from~\cite{EMoore}.}
\begin{tabular}{|c|c|c|c|c|}\hline
\hline{$$}&\multicolumn{3}{c|}{{$5s^{2}$$^{1}S_{0}$}}&{$$}\\\hline
$\textbf{\textrm{UpperState}}$&{WN}&{${A_{J}}$}&{$\boldsymbol{\zeta}$}&\raisebox{0ex}[0pt]{${A_{T}}$}\\\hline
$5s5p^{1}P_{1}$ &21698.48 &190.01 &1.000 &190.01$^{b}$\\
$5s6p^{1}P_{1}$ &34098.44 &1.87 &1.000 &1.87$^{d}$\\
$5s7p^{1}P_{1}$ &38906.90 &5.32 &1.000 &5.32$^{d}$\\
$5s8p^{1}P_{1}$ &41172.15 &14.9 &1.000 &14.9$^{a}$\\
$4d5p^{1}P_{1}$ &41184.47 &12 &1.000 &12$^{a}$\\
$5s9p^{1}P_{1}$ &42462.36 &11.6 &1.000 &11.6$^{a}$\\
$5s10p^{1}P_{1}$ &43327.94 &7.6 &1.000 &7.6$^{a}$\\
$5s11p^{1}P_{1}$ &43938.26 &4.88 &1.000
&4.88$^{a}$\\\hline
\end{tabular
\\
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|c|}\hline
\hline{$$}&\multicolumn{3}{c|}{{$5s5p^{3}P_{0}$}}
&\multicolumn{3}{c|}{{$5s5p^{3}P_{1}$}}
&\multicolumn{3}{c|}{{$5s5p^{3}P_{2}$}} &{$$}\\\hline
$\textbf{\textrm{UpperState}}$&{WN}
&{${A_{J}}$}&{$\boldsymbol{\zeta}$}&{WN}&{${A_{J}}$}&{$\boldsymbol{\zeta}$}&{WN}
&{${A_{J}}$}&{$\boldsymbol{\zeta}$}&\raisebox{0ex}[0pt]{${A_{T}}$}\\\hline
$5s6s^{3}S_{1}$ &14721.275&10.226&1.0828 &14534.444 &29.526 &1.0421 &14140.232 &45.314 &0.9596&85$^{c}$\\
$5s7s^{3}S_{1}$ &23107.193&1.402 &1.0517 &22920.362 &4.106 &1.0264 &22526.15 &6.495 &0.9743&12$^{e}$\\
$5s8s^{3}S_{1}$ &26443.920&0.954 &1.0450 &26257.089 &2.803 &1.0230 &25862.877 &4.464 &0.9776&8.22$^{a}$\\
$5s9s^{3}S_{1}$ &28133.680&0.525 &1.0422 &27946.849 &1.543 &1.0216 &27552.637 &2.464 &0.9790&4.53$^{a}$\\
$5s10s^{3}S_{1}$ &29110.080&0.320 &1.0408 &28923.249 &0.943
&1.0208 &28529.037 &1.508 &0.9797&2.77$^{a}$\\\hline
$5p^{2}$$^{3}P_{0}$&------- &0.000 &---- &20689.119 &117.64 &0.9803 &------- &0.000 &---- &120$^{e}$\\
$5p^{2}$$^{3}P_{1}$&21082.618&41.492&1.0373 &20895.787 &30.297 &1.0099 &20501.575 &47.690 &0.9538&120$^{e}$\\
$5p^{2}$$^{3}P_{2}$&------- &0.000 &---- &21170.317 &31.509
&1.0503 &20776.105 &89.343 &0.9927&120$^{e}$\\\hline
$5s4d^{3}D_{1}$ &3841.536 &0.290 &1.2660 &3654.705 &0.187 &1.0901 &3260.493 &0.009 &0.7740&0.412$^{f}$\\
$5s4d^{3}D_{2}$ &------- &0.000 &---- &3714.444 &0.354 &1.1444 &3320.232 &0.084 &0.8174&0.412$^{f}$\\
$5s4d^{3}D_{3}$ &------- &0.000 &---- &------- &0.000 &---- &3420.704 &0.368 &0.8938&0.412$^{f}$\\
$5s5d^{3}D_{1}$ &20689.423&35.732&1.0544 &20502.592 &26.080 &1.0261 &20108.38 &1.640 &0.9681&61$^{g}$\\
$5s5d^{3}D_{2}$ &------- &0.000 &---- &20517.664 &47.049 &1.0284 &20123.452 &14.796 &0.9702&61$^{g}$\\
$5s5d^{3}D_{3}$ &------- &0.000 &---- &------- &0.000 &---- &20146.492 &59.390 &0.9736&61$^{g}$\\
$5s6d^{3}D_{1}$ &25368.383&14.303&1.0457 &25181.552 &10.492 &1.0228 &24787.34 &0.667 &0.9755&24.62$^{f}$\\
$5s6d^{3}D_{2}$ &------- &0.000 &---- &25186.488 &18.897 &1.0234 &24792.276 &6.008 &0.9761&24.62$^{f}$\\
$5s6d^{3}D_{3}$ &------- &0.000 &---- &------- &0.000 &---- &24804.591 &24.069 &0.9776&24.62$^{f}$\\
$5s7d^{3}D_{1}$ &27546.88 &8.223 &1.0424 &27360.049 &6.043 &1.0213 &26965.837 &0.386 &0.9778&14.2$^{a}$\\
$5s7d^{3}D_{2}$ &------- &0.000 &---- &27364.969 &10.883 &1.0219 &26970.757 &3.473 &0.9783&14.2$^{a}$\\
$5s7d^{3}D_{3}$ &------- &0.000 &---- &------- &0.000 &---- &26976.337 &13.902 &0.9790&14.2$^{a}$\\
$5s8d^{3}D_{1}$ &28749.18 &4.920 &1.0407 &28562.349 &3.619 &1.0206 &28168.137 &0.231 &0.9789&8.51$^{a}$\\
$5s8d^{3}D_{2}$ &------- &0.000 &---- &28565.959 &6.517 &1.0210 &28171.747 &2.083 &0.9793&8.51$^{a}$\\
$5s8d^{3}D_{3}$ &------- &0.000 &---- &------- &0.000 &---- &28176.207 &8.337 &0.9797&8.51$^{a}$\\
$5s9d^{3}D_{1}$ &29490.28 &3.184 &1.0400 &29303.449 &2.342 &1.0203 &28909.237 &0.150 &0.9797&5.51$^{a}$\\
$5s9d^{3}D_{2}$ &------- &0.000 &---- &29303.449 &4.216 &1.0203 &28909.237 &1.350 &0.9797&5.51$^{a}$\\
$5s9d^{3}D_{3}$ &------- &0.000 &---- &------- &0.000 &----
&28914.037 &5.401 &0.9802&5.51$^{a}$\\\hline
\multicolumn{11}{|l|}{$^{a}$ ~\cite{SrA}, $^{b}$ ~\cite{Srb}, $^{c}$
~\cite{Src}, $^{d}$ ~\cite{Srd}, $^{e}$ ~\cite{Sre}, $^{f}$
~\cite{Srf}, $^{g}$ ~\cite{Srg}.}\\\hline
\end{tabular}
\end{widetext}
\end{scriptsize}
\end{tablehere}
\begin{tablehere
\begin{scriptsize}
\begin{widetext}
\caption{Ca element: Transition wavenumbers (WN)($cm^{-1}$)
corresponding to $\omega_{Jki}$, Einstein Coefficients for fine
structure states $A_J(\times10^6s^{-1})$ and Total
$A_T(\times10^6s^{-1})$ for fine structure states manifold,
Correction Factors $\boldsymbol{\zeta}$. Besides the updated values
listed in Ref.~\cite{Carsten} for $A_T$, the others originate
from~\cite{NIST}.}
\begin{tabular}{|c|c|c|c|c|}\hline
\hline{$$}&\multicolumn{3}{c|}{{$4s^{2}$$^{1}S_{0}$}}&{$$}\\\hline
$\textbf{\textrm{UpperState}}$&{WN}&{${A_{J}}$}&{$\boldsymbol{\zeta}$}&\raisebox{0ex}[0pt]{${A_{T}}$}\\\hline
$4s4p^{1}P_{1}$ &23652.304&218 &1.000 &218 \\
$4s5p^{1}P_{1}$ &36731.615&0.27 &1.000 &0.27\\
$4s6p^{1}P_{1}$ &41679.008&16.7 &1.000 &16.7\\
$4snp^{1}P_{1}$ &43933.477&30.1 &1.000 &30.1\\
$4s7p^{1}P_{1}$ &45425.358&15.3 &1.000 &15.3\\
$4s8p^{1}P_{1}$ &46479.813&6.1 &1.000 &6.1\\
\hline
\end{tabular
\\
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|c|}\hline
\hline{$$}&\multicolumn{3}{c|}{{$4s4p^{3}P_{0}$}}
&\multicolumn{3}{c|}{{$4s4p^{3}P_{1}$}}
&\multicolumn{3}{c|}{{$4s4p^{3}P_{2}$}} &{$$}\\\hline
$\textbf{\textrm{UpperState}}$&{WN}
&{${A_{J}}$}&{$\boldsymbol{\zeta}$}&{WN}&{${A_{J}}$}&{$\boldsymbol{\zeta}$}&{WN}
&{${A_{J}}$}&{$\boldsymbol{\zeta}$}&\raisebox{0ex}[0pt]{${A_{T}}$}\\\hline
$4s5s^{3}S_{1}$ &16381.594&9.855 &1.0195 &16329.432 &29.278 &1.0096 &16223.552 &47.845 &0.9899 &87$^{a}$\\
$4s6s^{3}S_{1}$ &25316.340&3.488 &1.0126 &25264.178 &10.397 &1.0062 &25158.298 &17.110 &0.9935 &31\\
$4s7s^{3}S_{1}$ &28822.866&1.573 &1.0110 &28770.704 &4.692 &1.0054 &28664.824 &7.733 &0.9943 &14\\
$4s8s^{3}S_{1}$ &30580.783&1.033 &1.0102 &30528.621 &3.083 &1.0052 &30422.741 &5.084 &0.9947 &9.2\\
$4s9s^{3}S_{1}$ &31590.382&0.606 &1.0099 &31538.220 &1.809 &1.0050 &31432.340 &2.985 &0.9950 &5.4\\
$4s10s^{3}S_{1}$ &32224.147&0.370 &1.0097 &32171.985 &1.105
&1.0049 &32066.105 &1.824 &0.9951 &3.3\\\hline
$4p^{2}$$^{3}P_{0}$&------ &0.000 &---- &23207.480 &179.046 &0.9947 &----- &0.000 &---- &180\\
$4p^{2}$$^{3}P_{1}$&23306.907&60.474&1.0079 &23254.745 &45.045 &1.0010 &23148.865 &74.033 &0.9871 &180\\
$4p^{2}$$^{3}P_{2}$&------ &0.000 &---- &23341.495 &45.563 &1.0125 &23235.615 &134.784 &0.9984 &180\\
$3d^{2}$$^{3}P_{0}$&------ &0.000 &---- &33314.030 &110.231 &1.0021 &----- &0.000 &---- &110\\
$3d^{2}$$^{3}P_{1}$&33379.722&36.971&1.0083 &33327.560 &27.594 &1.0034 &33221.680 &45.540 &0.9936 &110\\
$3d^{2}$$^{3}P_{2}$&------ &0.000 &---- &33353.459 &27.662
&1.0059 &33247.579 &82.178 &0.9961 &110\\\hline
$4s4d^{3}D_{1}$ &22590.296&48.981&1.0134 &22538.134 &36.471 &1.0061 &22432.254 &2.396 &0.9916 &87\\
$4s4d^{3}D_{2}$ &---------&0.000 &------ &22541.804 &65.687 &1.0067 &22435.924 &21.580 &0.9922 &87\\
$4s4d^{3}D_{3}$ &---------&0.000 &------ &--------- &0.000 &------ &22441.506 &86.400 &0.9931 &87\\
$4s5d^{3}D_{1}$ &27585.101&20.786&1.0112 &27532.939 &15.498 &1.0053 &27427.059 &1.021 &0.9934 &37\\
$4s5d^{3}D_{2}$ &---------&0.000 &------ &27534.653 &27.905 &1.0056 &27428.773 &9.192 &0.9937 &37\\
$4s5d^{3}D_{3}$ &---------&0.000 &------ &--------- &0.000 &------ &27431.444 &36.782 &0.9941 &37\\
$4s6d^{3}D_{1}$ &29891.172&13.472&1.0104 &29839.010 &10.049 &1.0049 &29733.130 &0.663 &0.9940 &24\\
$4s6d^{3}D_{2}$ &---------&0.000 &------ &29840.356 &18.094 &1.0052 &29734.476 &5.965 &0.9942 &24\\
$4s6d^{3}D_{3}$ &---------&0.000 &------ &--------- &0.000 &------ &29736.431 &23.868 &0.9945 &24\\
$4s7d^{3}D_{1}$ &31144.072&8.417 &1.0100 &31091.910 &6.279 &1.0047 &30986.030 &0.414 &0.9942 &15\\
$4s7d^{3}D_{2}$ &---------&0.000 &------ &31093.586 &11.306 &1.0050 &30987.706 &3.729 &0.9944 &15\\
$4s7d^{3}D_{3}$ &---------&0.000 &------ &--------- &0.000 &------ &30990.116 &14.922 &0.9948 &15\\
$4s8d^{3}D_{1}$ &31878.324&4.318 &1.0094 &31826.162 &3.221 &1.0041 &31720.282 &0.213 &0.9940 &7.7\\
$4s8d^{3}D_{2}$ &---------&0.000 &------ &31829.944 &5.803 &1.0048 &31724.064 &1.915 &0.9946 &7.7\\
$4s8d^{3}D_{3}$ &---------&0.000 &------ &--------- &0.000 &------ &31729.298 &7.663 &0.9952 &7.7\\
$4s3d^{3}D_{1}$ &5177.459 &0.502 &1.0503 &5125.297 &0.365 &1.0188 &5019.417 &0.023 &0.9570 &0.86$^{a}$\\
$4s3d^{3}D_{2}$ &---------&0.000 &------ &5139.197 &0.663 &1.0272 &5033.317 &0.207 &0.9650 &0.86$^{a}$\\
$4s3d^{3}D_{3}$ &---------&0.000 &------ &--------- &0.000
&------ &5055.057 &0.841 &0.9775 &0.86$^{a}$\\\hline
\multicolumn{11}{|l|}{\rule[1.5mm]{0mm}{3mm}\rule[-1mm]{0mm}{3mm}$^{a}$~\cite{Carsten}.}\\\hline
\end{tabular}
\end{widetext}
\end{scriptsize}
\end{tablehere}
between $A_{Tki}$ and $A_{Jki}$. We know $A_{Tki}$ can be
expressed as:
\begin{eqnarray}
A_{Tki} = \frac{e^2}{4\pi\epsilon_0}\frac{{4\omega
_{Tki}^3}}{{3\hbar {c^3}}} \frac{1}{{2{L_k} + 1}} |\left\langle
{{\beta _k}{L_k}}\left\| D \right\| {{\beta _i}{L_i}} \right\rangle|
^2
\end{eqnarray}
Here $\hbar\omega_{Tki}$ is the energy difference between two fine
structure manifold states $|k\rangle$ and $|i\rangle$. Using the
formula:
\begin{eqnarray}
\begin{split}
\left\langle {{\beta_k}{J_k}} \left\| D \right\|
{{\beta_i}{J_i}} \right\rangle
&= {( - 1)^{{L_k} + {S_k} + {J_i} + 1}}{\delta _{{S_k}{S_i}}}\sqrt {(2{J_k} + 1)(2{J_i} + 1)}\\
&\times \left\{ {\begin{array}{*{20}{c}}
{{J_k}} & 1 & {{J_i}} \\
{{L_i}} & {{S_i}} & {{L_k}} \\
\end{array}} \right\}\left\langle {{\beta _k}{L_k}} \left\| D \right\| {{\beta _i}{L_i}} \right\rangle\\
\end{split}
\end{eqnarray}
and combining Eq.(2) and (3), we can get
\begin{eqnarray}
A_{Jki} = A_{Tki}\times \zeta (\omega _{ki} ){\widetilde {R_{ki}} }
\end{eqnarray}
Here
\begin{eqnarray}
\zeta (\omega _{ki} )={\omega _{Jki}^3}/\omega _{Tki}^3
\end{eqnarray}
is the energy dependent correction~\cite{Martin}, reflecting the alteration on the transition rate due to the effects
such as the orbit-spin interaction and the spin-spin interaction
which causes the fine structure splitting.
And
\begin{eqnarray}
\begin{split}
\widetilde {R_{ki}}&=(2{L_k} + 1)(2J_i + 1) \times {\left\{
{\begin{array}{*{20}{c}}
J_k & 1 & J_i \\
L_i & S_i & L_k \\
\end{array}} \right\}^2}
\end{split}
\end{eqnarray}
gives the fraction of the coupling strength between an excited state
$|k\rangle$ and a lower state $|i\rangle$. Since the total
transition rate $A_{Tki}$ is usually available in the literature,
this geometric ratio tells us how to scale the interaction for a
particular fine structure state of interest.
To calculate the wavelength dependent polarizability, we combine
Eq.(1) with Eq.(5), and use the known transition frequencies and
spontaneous emission rates in the literature. This light
polarizability is very sensitive to the Einstein coefficient.
However, theoretical and experimental values of magic wavelength for
the optical clock transition obtained in the past can be used to
confirm our calculation.
In this paper, we use this method to calculate the light shift for
the terahertz clock transition from $^{3}P_{0}$ to $^{3}P_{1},m=0$
levels, and from $^{3}P_{1},m=0$ to $^{3}P_{2},m=0$ levels for boson
isotopes with the nuclear spin $I=0$. After calculating magic
wavelengths for Sr and Ca optical clock transitions and comparing
them to the experimental values, we calculate the polarizability of
terahertz transition with data collection mainly from
Ref.~\cite{NIST, EMoore,SrA,Srb,Src,Srd,Sre,Srf,Srg}.
\section{Calculation of Magic wavelength }
\subsection{Strontium}
Using the method above, for Strontium, we first calculate the magic
wavelengths of two optical lattice clock transitions with the datas
listed in Table I and compare the results with experimental values.
Then, we calculate the crossing points for terahertz clock
transitions where the difference of polarizability is zero. Table I
shows Transition Wavenumbers, Einstein Coefficients and Correction
Factors for the $5s^{2}$~$^{1}S_{0}$, $5s5p~^{3}P_{0}$,
$5s5p~^{3}P_{1}$ and $5s5p~^{3}P_{2}$ states for Sr element. For
Einstein Coefficient $A_{Tki}$, first we choose the available
updated experimental values in Ref.~\cite{Srb,Srd,Srg}, then we use
updated theoretical data in Ref.~\cite{Src,Srf}, and for the rest we
mainly use theoretical values in Ref.~\cite{SrA}.
According to our calculation, the crossing point for the $^{1}S_{0}$
to $^{3}P_{0}$ transition occurs at 813.1 nm, while the crossing
point for the $^{1}S_{0}$ to $^{3}P_{1}(m_{J}={\pm}1)$ transition
with linear polarized light takes place at 915.4 nm. Both of those
results are in agreement with the experimental values of 813.428(1)
nm~\cite{Martin58, Martin44, Martin41, Martin42} and 914(1)
nm~\cite{Src}. This confirms our calculation procedure.
\begin{figure}[ht]
\includegraphics [width=9cm]{newfig2}
\caption{(Color online) The wavelength dependence of the difference of atomic
polarizability for Sr element around 400 nm.}
\end{figure}
\begin{figure}[ht]
\includegraphics[width=9cm]{newfig3}
\caption{(Color online) The wavelength dependence of the difference of atomic
polarizability for Sr element around 1650 nm.}
\end{figure}
Fig. 2 and Fig. 3 display the wavelength dependence of the atomic
polarizability difference $\Delta\alpha$ for Sr with trapping laser
wavelength around 400 nm and 1650 nm, respectively. The result is
scaled by a factor of 1/($4\pi \varepsilon _0 a_0^3$) and the
polarizability is given in atomic unit. In Fig.2, for linear
polarized light, $\Delta\alpha$ between $^3P_1$ and $^3P_0$ and
$\Delta\alpha$ between $^3P_2$ and $^3P_1$ are given in solid and
dash dotted lines, respectively. In Fig.3, $\Delta\alpha$ between
$^3P_1$ and $^3P_0$ for linear polarized light and $\Delta\alpha$
between $^3P_2$ and $^3P_1$ for circular polarized light are
presented. The cross markers are the crossing points where
$\Delta\alpha$ is zero. From Fig.2 and 3, we can know that the magic
wavelength for $^{3}P_{0}$ to $^{3}P_{1},m=0$ with linear polarized
light are 381.2 nm, 413.6 nm, 419.3 nm, 1714 nm and 3336 nm, while
for $^{3}P_{1},m=0$ to $^{3}P_{2},m=0$ are 384.5 nm,441.9 nm and
511.0 nm. On the other hand, for $m=0$ and circular polarization of
light, the magic wavelengths for $^{3}P_{0}$ to $^{3}P_{1},m=0$
transition are 511.8 nm and 662.8 nm, while for $^{3}P_{1},m=0 $ to
$^{3}P_{2},m=0$, the magic wavelengths are 717.7 nm and 1591 nm.
\subsection{Calcium}
We calculate the polarizabilities using the data in Table II with
the same method. Table II shows Transition Wavenumbers, Einstein
Coefficients and Correction Factors for the $4s^{2}~$$^{1}S_{0}$,
$4s4p~^{3}P_{0}$, $4s4p~^{3}P_{1}$ and $4s4p~^{3}P_{2}$ states for
Ca. For Einstein Coefficient, we use the updated theoretical values
according to Ref.~\cite{Carsten}, and others are from the data
listed in Ref.~\cite{NIST}. In order to check the accuracy of our
calculation and the data used, we get the magic wavelength 799.2 nm
for the $^{1}S_{0}, m=0$ to $^{3}P_{1}, m=0$ optical transition with
circularly polarized trapping light, which agrees well with the
experimental value 800.8( 22 ) nm~\cite{Carsten}.
The wavelength dependence of the atomic polarizability difference
$\Delta\alpha$ around 350 nm and 1350 nm are shown with atomic unit
in Fig.4 and 5, respectively. The crossing points where the
$\Delta\alpha$ is zero are marked by cross. The magic wavelengths
for linear polarization occur at 1361 nm and 2066 nm for clock
transition $^{3}P_{0}-^{3}P_{1}$, while for the transition between
\begin{tablehere
\begin{scriptsize}
\begin{widetext} \caption{Mg element: Transition wavenumbers (WN)($cm^{-1}$)
corresponding to $\omega_{Jki}$, Einstein Coefficients for fine
structure states $A_J(\times10^6s^{-1})$ and Total
$A_T(\times10^6s^{-1})$ for fine structure states manifold,
Correction Factors $\boldsymbol{\zeta}$ for Mg element. The
Wavenumber and $A_T$ data originate from~\cite{NIST}.}
\begin{tabular}{|c|c|c|c|c|}\hline
\hline{$$}&\multicolumn{3}{c|}{{$3s^{2}$$^{1}S_{0}$}}&{$$}\\\hline
$\textbf{\textrm{UpperState}}$&{WN}&{${A_{J}}$}&{$\boldsymbol{\zeta}$}&\raisebox{0ex}[0pt]{${A_{T}}$}\\\hline
$3s3p^{1}P_{1}$ &35051.264&491 &1.000 &491\\
$3s4p^{1}P_{1}$ &49346.729&61.2 &1.000 &61.2\\
$3s5p^{1}P_{1}$ &54706.536&16.0 &1.000 &16.0\\
$3s6p^{1}P_{1}$ &57214.990&6.62 &1.000 &6.62\\
$3s7p^{1}P_{1}$ &58580.230&3.28 &1.000 &3.28\\
$3s8p^{1}P_{1}$ &59403.180&1.88 &1.000 &1.88\\
\hline
\end{tabular
\\
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|c|}\hline
\hline{$$}&\multicolumn{3}{c|}{{$3s3p^{3}P_{0}$}}
&\multicolumn{3}{c|}{{$3s3p^{3}P_{1}$}}
&\multicolumn{3}{c|}{{$3s3p^{3}P_{2}$}} &{$$}\\\hline
$\textbf{\textrm{UpperState}}$&{WN}
&{${A_{J}}$}&{$\boldsymbol{\zeta}$}&{WN}&{${A_{J}}$}&{$\boldsymbol{\zeta}$}&{WN}
&{${A_{J}}$}&{$\boldsymbol{\zeta}$}&\raisebox{0ex}[0pt]{${A_{T}}$}\\\hline
$3s4s^{3}S_{1}$ &19346.998&11.293 &1.0063 &19326.939 &33.774 &1.0032 &19286.225 &55.932 &0.9968 &101\\
$3s5s^{3}S_{1}$ &30022.121&3.380 &1.0041 &30002.062 &10.120 &1.0020 &29961.348 &16.800 &0.9980 &30.3\\
$3s6s^{3}S_{1}$ &34041.395&1.372 &1.0036 &34021.336 &4.107
&1.0018 &33980.622 &6.821 &0.9982 &12.3\\\hline
$3p^{2}$$^{3}P_{0}$&------- &0.000 &---- &35942.306 &537.085 &0.9983 &------- &0.000 &---- &538\\
$3p^{2}$$^{3}P_{1}$&35982.995&179.638 &1.0017 &35962.936 &134.500 &1.0000 &35922.222 &223.405 &0.9966 &538\\
$3p^{2}$$^{3}P_{2}$&------- &0.000 &---- &36003.476 &134.957
&1.0034 &35962.762 &403.500 &1.0000 &538\\\hline
$3s3d^{3}D_{1}$ &26106.653&89.865 &1.0047 &26086.594 &67.244 &1.0024 &26045.880 &4.462 &0.9977 &161\\
$3s3d^{3}D_{2}$ &------- &0.000 &---- &26086.563 &121.028 &1.0023 &26045.849 &40.157 &0.9977 &161\\
$3s3d^{3}D_{3}$ &------- &0.000 &---- &------- &0.000 &---- &26045.867 &160.630 &0.9977 &161\\\
$3s4d^{3}D_{1}$ &32341.930&28.106 &1.0038 &32321.871 &21.040 &1.0019 &32281.157 &1.397 &0.9981 &50.4\\
$3s4d^{3}D_{2}$ &------- &0.000 &---- &32321.830 &37.872 &1.0019 &32281.116 &12.576 &0.9981 &50.4\\
$3s4d^{3}D_{3}$ &------- &0.000 &---- &------- &0.000 &---- &32281.078 &50.304 &0.9981 &50.4\\
$3s5d^{3}D_{1}$ &35117.866&13.101 &1.0035 &35097.807 &9.808 &1.0017 &35057.093 &0.652 &0.9983 &23.5\\
$3s5d^{3}D_{2}$ &------- &0.000 &---- &35097.784 &17.655 &1.0017 &35057.070 &5.865 &0.9983 &23.5\\
$3s5d^{3}D_{3}$ &------- &0.000 &---- &------- &0.000 &---- &35057.040 &23.460 &0.9983 &23.5\\
$3s6d^{3}D_{1}$ &36592.473&6.967 &1.0033 &36572.414 &5.217 &1.0017 &36531.700 &0.347 &0.9983 &12.5\\
$3s6d^{3}D_{2}$ &------- &0.000 &---- &36572.381 &9.391 &1.0017 &36531.660 &3.120 &0.9983 &12.5\\
$3s6d^{3}D_{3}$ &------- &0.000 &---- &------- &0.000
&---- &36531.657 &12.479 &0.9983 &12.5\\\hline
\end{tabular}
\end{widetext}
\end{scriptsize}
\end{tablehere}
level $^{3}P_{1}, m=0$ and $^{3}P_{2},m=0 $ at 312.2 nm, 316.2 nm,
325.4 nm, 344.0 nm and 393.4 nm.
The laser polarization have no effect on the polarizability for the
ground state ( $J=0$ ) because the ac Stark shift is identical with
any polarizations. It is also true for $^{3}P_{0}$ state. However,
the influence of circular polarized laser light is worth study for
other states. For $m=0$, we can obtain the magic wavelengths for the
$^{3}P_{0}$ to $^{3}P_{1}$ clock transition at 301.0 nm and 310.0
nm, while for $^{3}P_{1}$ to $^{3}P_{2}$ one finds 1318 nm and 2254
nm.
\begin{figure} \centering
\includegraphics[width=9cm]{newfig4}
\caption{(Color online) Wavelength dependence of the difference between excited state and ground state atomic
polarizability around 350 nm for Ca element.}
\end{figure}
\begin{figure} \centering
\includegraphics[width=9cm]{newfig5}
\caption{(Color online) Wavelength dependence of the difference between excited state and ground state atomic
polarizability around 1350 nm for Ca element.}
\end{figure}
\subsection{Magnesium}
With the completion of the NIST database, the atomic polarizability
of the Mg triplet states in the presence of linear and circular
polarized light can also be calculated. Table III presents
Transition Wavenumbers, Einstein Coefficients and Correction Factors
for the $3s^{2}~$$^{1}S_{0}$, $3s3p~^{3}P_{0}$, $3s3p~^{3}P_{1}$ and
$3s3p~^{3}P_{2}$ states for Mg element. Using the data presented in
Table III, the magic wavelengths of $^{3}P_{0}$ to $^{3}P_{1},m=0$
transition with linear polarization are 335.6 nm and 399.5 nm. The
magic wavelengths are 308.6 nm, 336.5 nm, 406.1 nm for the
transition between $^{3}P_{1},m=0$ and $^{3}P_{2},m=0$.
For circular polarization of light, the magic wavelengths for the
transition $^{3}P_{0}$ to $^{3}P_{1},m=0$ are 307.7 nm, 336.4 nm,
407.8 nm. However, we can not find any magic wavelength for circular
laser between level $^{3}P_{1},m=0 $ and $^{3}P_{2},m=0$.
For Mg atoms, several optical transitions between the energy levels
of terahertz clock transition states and other levels exist, such as
456.5 nm ( ${3s^{2}}~{^1S} - 3s3p~^{3}P$ ), 383.6 nm ( $3s3p~^{3}P -
3s3d~^{3}D$ ), 309.6 nm ( $3s3p~^{3}P - 3s4d~^{3}D$ ), 333.2 nm (
$3s3p~^{3}P - 3s5s~ ^{3}S$ ) and 517.4 nm ( $3s3p~^{3}P -
3s4s~^{3}S$ ). Hence, not all the magic wavelengths are good enough
for clock transition, because the slope of the light shift
difference with the wavelength is too large (shown in the final
table IV). To some extent, a possible magic wavelength near 400 nm
is shown in Fig. 6 with the atomic unit. In Fig.6, $\Delta\alpha$
for $^3P_1-^3P_0$ transition and $^3P_2-^3P_1$ transition with
different polarization are given. The cross markers reflect the
crossing points where the atomic polarizability difference is zero.
\begin{figure} \centering
\includegraphics[width=9cm]{newfig6}
\caption{(Color online) The wavelength dependence of the difference of atomic
Polarizability around 400 nm for Mg element.}
\end{figure}
\section{Discussions and Conclusions}
\begin{table}[ht]
\begin{flushleft}
\caption{Magic wavelengths for terahertz region. L1 is the linear
laser for the $^{3}P_{0}$ to $^{3}P_{1}$ clock transition while C1
is for the circular laser, L2 is the linear laser for the
$^{3}P_{1}$ to $^{3}P_{2}$ clock transition while C2 is for the
circular laser. $\kappa$ is the slope of the shift difference of two
level of clock transition levels at the corresponding trapping laser
wavelength with unit $Hz/nm$, the sign denotes the direction of the
change with the shift for the high level minus the lower level. The
data are given in the reasonable experiment condition with input
power 150 $mW$ focused to a waist of 65 $\mu m$ as the light
intensity $1.1301*10^3 W/cm^2$.}
\end{flushleft}
\centering
\begin{tabular}{|c|c|c|c|c|c|c|}\hline\hline
&\multicolumn{2}{c|}{$Mg$}&\multicolumn{2}{c|}
{\rule[1.5em]{2.5mm}{0mm}\rule[-0.8em]{2.5mm}{0mm}$Ca$\rule[1.5em]{5mm}{0mm}}&\multicolumn{2}{c|} {$Sr$}\\\cline{2-7}
\raisebox{1.5ex}[0pt]{$\boldsymbol{^{3}P_{0,1,2}}$}&\rule[1.5em]{4mm}{0mm}$\lambda$\rule[1.5em]{4mm}{0mm}&$\kappa$&
\rule[1.5em]{4mm}{0mm}$\lambda$\rule[1.5em]{4mm}{0mm}&$\kappa$&\rule[1.5em]{4mm}{0mm}$\lambda$\rule[1.5em]{4mm}{0mm}&$\kappa$\\
&$(nm)$\rule[-1em]{0mm}{0mm}&($Hz/nm$)&$(nm)$&($Hz/nm$)&$(nm)$&($Hz/nm$)\\
\hline
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm}$L1$ &335.6&-1125&1361&-3.201&381.2&-615.7\\
\cline{2-7}\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm}&399.5&-103.5&2066&54.94&413.6&32.22\\ \cline{2-7}
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm} &&&&&419.3&-31.4377\\ \cline{2-7}
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm} &&&&&1714&-5.590\\ \cline{2-7}
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm} &&&&&3336&9.122\\
\hline
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm}$L 2$ &308.6&-26574&312.2&8319&384.5&1792\\ \cline{2-7}
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm}
&336.5&1905&316.2&5879&441.9&5759\\\cline{2-7}
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm}
&406.1&220.7&325.4&1641&511&1697.63\\ \cline{2-7}
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm} &&&344.0&542.9&&\\\cline{2-7}
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm} &&&393.4&1787&&\\\cline{2-7}
\hline
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm}$C1$ &307.7&-3174&301.0&8610&511.8&252.8\\ \cline{2-7}
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm} &336.4&469.9&310.0&-4072&662.8&-1068\\ \cline{2-7}
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm} &407.8&54.76&&&&\\
\hline
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm}$C 2$ &&&1318&-3.365&717.7&2692\\ \cline{2-7}
\rule[1mm]{0mm}{3mm}\rule[-1.5mm]{0mm}{4mm} &&&2254&13.39&1591&-5.551\\ \hline
\end{tabular}
\end{table}
In summary, we have calculated magic wavelengths for terahertz clock
transitions for alkaline-earth atoms. The calculation results are
presented in Table IV along with the slopes of the difference of
polarizabilities at corresponding magic wavelengths. Depending on
the calculation and the current laser development, we recommend
1714nm and 1591nm for Sr terahertz clock, 1361nm and 1318nm for Ca
terahertz clock, 399.5nm and 407.8nm for Mg terahertz clock, because
the difference of polarizabilities have small slopes at these magic
wavelengths, where we ignore the effect of highly excited states and
continuum states which can only make little contribution to the
wavelength dependent polarizabilities at terahertz region.
In this paper, we are only focusing on the study of possible magic
wavelengths of trapping laser for these terahertz clock transitions
of Sr, Ca and Mg atoms. These terahertz clock transitions were first
proposed as early as 1972~\cite{Strumia}, and recently have been
proposed to be applied in active optical clock~\cite{Chen}. These
clock transitions of alkaline-earth atoms correspond to a 0.6 THz to
11.8 THz frequency region. After the successful developments of
microwave fountain frequency standards, optical clocks with trapped
ions and optical lattice trapped neutral atoms, it is interesting to
study clock transitions at terahertz wavelengths. The advantages and
disadvantages of terahertz magic atomic clock will be discussed
elsewhere. The wavelength range studied in this paper (from 500
$\mu$m to 25 $\mu$m) corresponding to THz frequency standards fills
the gap between microwaves and optical waves.
We thank V. Thibault for critical reading our manuscript, and J. M. Li for his helpful discussions.
We appreciate the anonymous referee for the useful suggestions. This work is partially supported by the state Key Development Program for
Basic Research of China (No.2005CB724503, 2006CB921401,
2006CB921402) and NSFC (No.10874008, 10934010, 60490280 and
10874009). This work is also supported by PKIP of CAS (KJCX2.YW.W10) and Open Research Found of
State Key Laboratory of Precision Spectroscopy (East China
Normal University).\\
|
\section{Introduction}
Since its introduction in 1997, WLAN standard IEEE 802.11 received tremendous attention from both researchers and industry. Its widespread commercial
adoption attributes to its self configuring nature
of operation which offers easy and cheap deployment and maintenance.
In addition to providing wireless last mile coverage for the Internet, the standard has an inherent potential for integration of smart home components using cheap wireless devices.
Although wireless communication technology is also available as GSM, CDMA,
and Bluetooth, etc., designing wireless networks with these technologies imposes certain challenges for everyday use applications with high resource constraints.
For example, Bluetooth suffers from low bandwidth and short transmission range. It is attractive for personal area networks (PAN) but can not
cover large house or support community networking.
On the other hand, GSM and CDMA provide higher bandwidth and greater coverage. But the cost of installation and infrastructure maintenance makes the service cost very high rendering them inapplicable for everyday communications.
The newly introduced WiMAX, which is the industry counterpart of IEEE 802.16, overcame large scale cellular network deployment through long coverage only to offer another costly solution.
On the other hand, compared to other standards, IEEE 802.11 based WLAN offers moderate bandwidth with moderate coverage but does not require large scale infrastructure or costly devices and operates in unlicensed bands.
Due to the ad hoc nature of the standard, major
infrastructure deployment is not needed which makes it attractive for diverse
wireless applications including Ad Hoc, sensor, and mesh networks.
Only a small number of APs (Access Point) can serve Internet connectivity to a large network. Stations can join
and leave the premise of any AP without restriction.
Minimal scale infrastructure makes the operations cheap which can be availed for any kind of communication.
But the absence of infrastructure has its toll on performance.
Firstly, there is less control over the network.
Due to the inherent nature of wireless communication, new connections can
not be restricted compared to a slot based wired router.
Thus, network overloading is hard to protect against. Secondly, at each hop (closer to the AP), the number of transmissions becomes higher and a critical zone is formed around the AP limiting channel bandwidth even more.
The formation of critical zone limits the achievable throughput of
the network.
These are the reasons why a careful capacity study should be undertaken before designing and deploying such a network.
IEEE 802.11 defines two medium access methods, namely, Distributed Coordination Function (DCF) and Point Coordination Function (PCF).
DCF is more popular in literature and industry due to its ad-hoc nature of operation and inter-operability.
To arbitrate access to the channel, DCF uses a Binary Exponential Backoff (BEB) algorithm which can be either $\infty$-retry type or $m$-retry type.
If the sender keeps trying to send a packet until it succeeds, the backoff
algorithm is termed $\infty$-retry BEB where a packet can suffer repeated failures and hog the channel for long time.
$m$-retry BEB drops a packet after $m$ failures giving the next packet a
greater chance of reaching destination in time so that delay sensitive applications perform better.
In our current work, we analyze $m$-retry BEB based DCF.
A number of works studied performance of WLANs under
various conditions.
Bianchi~\cite{Bianchi2000} used a Markov model analysis
to investigate saturation throughput of a $\infty$-retry limit BEB based DCF assuming ideal channel and ignoring capture effect.
Ziouva and Antonakopoulos~\cite{Ziouva2002} presented a similar study with modified
assumptions on idle channel condition.
Bianchi's model was later extended by Wu \textit{et al.}~\cite{Wu2002} to model
$m$-retry BEB which is further extended by Chatzimisios \textit{et al.}~\cite{Chatzimisios2005,Chatzimisios2004}
to incorporate the effect of imperfect channel condition.
Although real networks operate under unsaturated load for a considerable amount of time~\cite{Daneshgaran2008}, all the above works investigated throughput under saturation condition only.
But the study of unsaturated network for both finite and infinite retry
limit BEB is of great importance in practical terms.
Barowski \textit{et al.}~\cite{Barowski2005} and Daneshgaran \textit{et al.}~\cite{Daneshgaran2008} calculated throughput under unsaturated load for an $\infty$-retry limit BEB.
Additionally, capture effect is incorporated in~\cite{Daneshgaran2008}.
Liaw \textit{et al.}~\cite{Liaw2005} presented unsaturated throughput for $m$-retry limit but the model did not consider capture effect or imperfect channel.
Most commercial 802.11 products support power capture today which brings a positive impact on overall network performance.
But to the best of the authors' knowledge, no existing work in the current literature presents an analytical model of $m$-retry BEB based DCF considering power capture.
Moreover, queuing analysis is mostly ignored which plays a crucial role in determining capacity of delay and loss sensitive applications\cite{Siddique2008,Siddique2009}.
No existing study presents a comprehensive
model and theoretical performance analyses of 802.11 WLANs considering all four factors concomitantly, namely, \setcounter{romanCounter}{1}(\roman{romanCounter})
unsaturated traffic \setcounter{romanCounter}{2}(\roman{romanCounter})~finite retry limit, \setcounter{romanCounter}{3}(\roman{romanCounter})~imperfect channel, and \setcounter{romanCounter}{4}(\roman{romanCounter})~power capture.
In this paper, we propose a simplistic Markov model considering all the
above factors and investigate the performance of an unsaturated IEEE 802.11 compliant WLAN in details, both theoretically and through simulation.
Consideration of imperfect channel and capture effect makes our model more suitable to reflect real world scenario closely.
Our model can accommodate both saturated and unsaturated traffic and is equally
applicable for both basic and RTS/CTS type handshake.
Common performance measures are expressed in terms of network configuration
which makes the model useful in network design and planning.
Medium access delay and queuing loss are also estimated which will
be very useful in devising techniques to ensure QoS of delay sensitive applications
over 802.11 networks.
\section{Analytical Model}
\label{section:Mathematical Model}
We develop a Markov model for DCF mechanism employing $m$-retry BEB to analyze
performance of a homogeneous WLAN.
The model considers effects of imperfect channel and power capture and is applicable to both Basic and RTS/CTS type DCF.
It considers unsaturated traffic but can also model saturated traffic, as shown later.
\subsection{The Markov Model}
Let $n$ be the number of contending stations, $m$ the retry limit, $m'$ the number of retry stages and $p_E$ the packet error rate in an imperfect channel.
Generally, $p_E$ depends on the modulation scheme and device parameters such as transmission rate, noise, and header/data length, etc.
The model considers channel error as packet error rate and thus it can be used with any modulation scheme.
If $P\{Event\}$ is the probability of $Event$ then we define the following necessary notations and their relations for the most common probabilities
which are used in the model.
\begin{equation*}\begin{split}
\tau &{=} P\{\text{Random\ node\ transmits\ in\ random\ slot}\},\\
p_B &{=} P \{\text{Channel\ is\ busy}\}{=}1{-}(1{-}\tau)^n,\\
p_C &{=} P \{\text{Collision} | \text{Transmission}\}{=}{{(1 {-} (1 {-} \tau )^{n {-} 1} )} \mathord{\left/
{\vphantom {{\left(1 {-} \left(1 {-} \tau \right)^{n {-} 1} \right)} {p_B }}} \right.
\kern-\nulldelimiterspace} {p_B }}
.\\
\end{split}\end{equation*}
Capture probability $p_p$ is the probability of a collided frame being received correctly
by power capture i.e., a frame with higher power can be received even when more than one frame collided.
Assuming an 11 chip Barker sequence being used for code spreading, the spreading factor
$s$ is 11.
The receiver uses a co-relator to de-spread the original signal. The inverse of processing gain for correlation receiver, denoted by $g$, is given by
$g{=}2(3s)^{{-}1}$.
If power capture in Rayleigh fading channel is enabled, the capture probability
for a simultaneous $i$ interfering packets is given by \cite{Hadzi-velkov2002},
\begin{equation}
p_p{=}\sum_{i=1}^{n{-}1}{\binom{n}{i+1}}\tau^{i+1}(1{-}\tau)^{n-i-1}(1{+}zg)^{-i}
\end{equation}
where $z$ is the capture threshold.
$p_F$ is the probability that a transmission failed, i.e., an ACK is not received
and the packet requires a retransmission.
This can happen due to a reception with error, a collision without capture, or both.
$p_F$ can be defined as,
\begin{equation}\label{eqn:p_E}
p_F{=}(p_C{-}p_p){+}p_Ep_p{+}(p_E{-}p_Ep_C)
\end{equation}
$p_S$ is the probability that a transmitted frame is received correctly,
i.e., it is acknowledged and $p_S{=}1{-}p_F$.
We define $p_Q$ to be the probability that the queue is non-empty.
$p_Q$ is a function of the packet arrival rate $\lambda$ and channel access delay $d_C$. Assuming a $M/M/1/l_Q$ queue of length $l_Q$ with Markovian
arrival and departure, $p_Q$ can be shown to be
\begin{equation}
p_Q{=}\frac{\lambda d_C {-} (\lambda d_C)^{l_Q{+}1}}{1{-}(\lambda d_C)^{l_Q{+}1}}.
\end{equation}
Here $d_C$ is the time required to serve each packet which is given by
the interval between the time when a packet becomes the head of the queue
and the time when it is acknowledged and removed from the queue.
\begin{figure}[!t]\centering\includegraphics[width=0.33\textwidth]{hmm}
\caption{Markov model.}\label{figure:siddi02}\end{figure}
The Markov model we used to model DCF with $m$-retry BEB is shown in Fig.~\ref{figure:siddi02}.
The states (ellipses) in a row are in the same retry stage while retry stages
increase from top to bottom.
The columns denote different counter values increasing from left to right.
The state $E$ at the top left corner of the figure represents an empty queue scenario.
This state is the key to modeling unsaturated network.
When the queue is empty, the system stays in state $E$.
Although our model resembles those presented in~\cite{Chatzimisios2004} and \cite{Daneshgaran2008}, it differs in a number of ways.
\cite{Daneshgaran2008}~modeled $\infty$-retry BEB where a packet is retransmitted
until success.
Therefore, the outgoing arcs from state ($m'$,$0$) with probability $p_F$
are recursive to $m'$-th retry stage and retry stages $m'{+}1,m'{+}2,...,m$
are not present~in\cite{Daneshgaran2008}.
Our model being $m$-retry BEB, retry stage is reset irrespective of success
and the outgoing arc from state $(m,0)$ neither is recursive,
nor depends on success or failure of the last transmission.
On the other hand, \cite{Chatzimisios2004} modeled saturation traffic.
Therefore, when the retry stage is reset, the Markov model always goes to the $0$-th retry stage and empty queue state $E$ is not present.
In our model, if a transmission is successful or the retry limit is exceeded, $E$ becomes the new state if the queue is empty.
Although \cite{Liaw2005} investigated unsaturated load for $m$-retry BEB,
the effects of imperfect channel and power capture are ignored.
We define the Markov model states as a bi-dimensional process $\left\{s(t),b(t)\right\}$ where $s(t)$ is the retry stage and
$b(t)$ is the counter value at time $t$.
The current backoff window $W_i$ at retry stage $i$ is defined as,
\begin{equation*}
W_i{=}\begin{cases}
2^iW
&\text{for } 0\leqslant i\leqslant m',\\
2^{m'}W
&\text{for } i>m'.\\
\end{cases}\end{equation*}
\subsection{One Step Transition Probabilities}
We denote the one step transition probability from state $(i,k)$ to $(i',k')$ with $P\{i',k'|i,k\}$ are defined as,
\begin{equation*}\setlength{\arraycolsep}{0pt}
\begin{array}{l}
P\{i,k|i,k{+}1\}{=}1 \text{\ for } \ 0{\leqslant} i{\leqslant} m\text{ and }1{\leqslant} k{\leqslant} W_{i}{-}2,\\
P\{i,k|i{-}1,0\}{=}
\frac{p_F}{W_i} \text{ for } 1{\leqslant} i{\leqslant} m\text{ and }\ 0{\leqslant} k{\leqslant} W_i{-}1,\\
P\{0,k|i,0\}{=}\frac{p_Q(1{-}p_F)}{W}\text{ for } 0{\leqslant} i{\leqslant} m{-}1,0{\leqslant} k{\leqslant} W{-}1,\\
P\{0,k|m,0\}{=}\frac{p_Q}{W} \text{\ for\ } 0{\leqslant} k{\leqslant} W{-}1,\\
P\{E|i,0\}=(1-p_Q)(1-p_F) \text{\ for\ } \ 0{\leqslant} i{\leqslant} m{-}1,\\
P\{E|m,0\}{=}(1-p_Q) \text{ for } 0{\leqslant} i{\leqslant} m{-}1,\\
P\{0,k|E\}{=}\frac{1}{W} \text{\ for\ } 0{\leqslant} k{\leqslant} W{-}1.
\end{array}
\end{equation*}
\subsection{Stationary Probabilities of the Markov States}
We define $b_{i,k}$ as the stationary probability distribution of any state $(i,k)$
which is given by,
\begin{equation*}
b_{i,k}{=}\lim\limits_{t\to \infty}P\{s(t){=}i,c(t){=}k\},i\in[0,m],k\in[0,2^iW].
\end{equation*}
Let the steady state probability of the system being in state $E$ be denoted by $b_E$.
As long as
the retry limit is not exceeded, the retry stage increments after each failed
transmission and
\begin{equation*}
b_{i,0}{=}b_{i{-}1,0}p_{F}{=}p_{F}^ib_{0,0} \text{ for } 1{\leqslant} i{\leqslant} m .
\end{equation*}
Due to chain regularity,
\begin{equation*}
b_{i,k}{=}\frac{W_i{-}k}{W_i}
\begin{cases}
p_Q(1{-}p_{F})\sum\limits^{m{-}1}_{j{=}0}b_{j,0}{+}p_{Q}b_{m,0}
& \text{for } i{=}0,\\
p_{F}b_{i{-}1,0}
&\text{for } 1{\leqslant} i{\leqslant} m.\\
\end{cases}\end{equation*}
The steady state probability of entering and leaving any state is equal.
Imposing this condition on state $E$ we obtain,
\begin{equation}
\label{equation:b_E}
\setlength{\arraycolsep}{0pt}
b_E{=}\frac{1}{p_Q}\left\{(1{-}p_Q)(1{-}p_{F})\sum\limits_{i{=}0}^{m{-}1}b_{i,0}{+}(1{-}p_Q)b_{m,0}\right\}.
\end{equation}
Imposing the same condition on state $\{0,0\}$ gives,
\begin{equation}\label{equation:b_E,simple}
p_{F}b_{0,0}
{=}(1{-}p_{F})\sum\limits_{i{=}0}^{m}b_{i,0}{+}p_{F}b_{m,0}.
\end{equation}
Sum of all states, where a packet is transmitted, gives,
\begin{equation}
\label{equation:sum of b_i,0}
\sum\limits_{i{=}0}^{m{-}1}b_{i,0}{=}b_{0,0}\sum\limits_{i{=}0}^{m{-}1}p^{i}{=}\frac{1{-}p^{m}}{1{-}p_F}b_{0,0}.
\end{equation}
From (\ref{equation:b_E}) \& (\ref{equation:sum of b_i,0}), we derive
\begin{equation*}
b_E
{=}\frac{1{-}p_Q}{p_Q}b_{0,0}.\\
\end{equation*}
\begin{figure*}[!t]
\normalsize
\setcounter{tempCounter}{\value{equation}}
\setcounter{equation}{7}
\begin{equation}
\label{equation:b_{0,0}}
b_{0,0}{=}\frac{2p_Q(1{-}p_F)(1{-}2p_F)}{
\begin{split}
p_QW(1{-}p_F)\left(1{-}(2p_F)^{m'{+}1}\right){+}p_Q(1{-}2p_F)(1{{-}}p_F^{m'{+}1}){+}p_Q
p_F^{m'{+}1}(1{-}2p_F)(2^{m'}W{+}1)(1{-}p_F^{m{-}m'})\\
\:{+}2(1{-}p_F)(1{-}2p_F)(1{-}p_Q)(1{-}p_F^m){+}2p_F^m(1{-}p_F)(1{-}2p_F)(1{-}p_Q)
\end{split}}.
\end{equation}
\setcounter{equation}{\value{tempCounter}}
\hrulefill
\end{figure*}
The sum of probabilities of being in every state should be equal to $1$ i.e.,$\sum\limits_{i{=}0}^{m}\sum\limits_{k{=}0}^{W_i{-}1}{b_{i,k}}{+}b_E{=}1
$ which gives (\ref{equation:b_{0,0}}).
Under saturation condition, a node will always have some packet to send.
Therefore, $p_Q{\to}1$
which gives $b_{0,0}$ for saturation model as shown in~\cite{Chatzimisios2004}.
A packet is transmitted only from the states $b_{i,0}$ where $i{\in}[0,m]$. Since $\tau$ denotes
the probability of transmission by a random node at a random slot,
\setcounter{equation}{8}
\begin{equation}
\label{eqn:tau}
\tau{=}\sum\limits_{i{=}0}^{m}b_{i,0}{=}b_{0,0}\sum\limits_{i{=}0}^{m}p_F^{i}{=}\frac{1{-}p_F^{m{+}1}}{1{-}p_F}b_{0,0}.
\end{equation}
Let $t_{\sigma}$, $t_{sifs}$, and $t_{difs}$ be the length of an idle slot, SIFS, and DIFS period.
$t_{\delta}$ is the propagation delay.
Similarly, $t_D$, $t_A$, $t_R$, and $t_C$ are the transmission time of data frame, ACK frame, RTS frame, and CTS frame, respectively.
A successful transmission is detected by the reception of an ACK packet.
We define $t_S$ as the time a sender has to wait before receiving an ACK
for the last packet and remove it from the queue. Taking into account of all the delay elements, $t_S$ can be defined as
\begin{equation*}\setlength{\arraycolsep}{0pt}
t_S {=}
\begin{cases}
t_{difs} {+} t_D {+} t_{\delta} {+} t_{sifs} {+} t_A {+} t_{\delta}
&\text{for Basic,}\\
\begin{split}
t_{difs} {+} t_R {+} t_{\delta} {+} t_{sifs} {+} t_{C} {+} t_{\delta} {+}
t_{sifs}\\
{+} t_D {+} t_{\delta} {+} t_{sifs} {+} t_A {+} t_{\delta}\\
\end{split} &
\text{for RTS/CTS}.\\
\end{cases}
\end{equation*}
When a transmission fails, the sender can detect neither collision nor erroneous
reception.
Therefore, it has to wait for a time duration $t_F$ before taking the transmission to be failed, which can be defined as
\begin{equation*}\setlength{\arraycolsep}{0pt}
t_F {=}\begin{cases}
t_{difs} {+} t_D {+} t_{\delta} {+} t_{sifs} {+} t_A {+} t_{\delta}
&\text{for Basic,}\\
t_{difs} {+} t_R {+} t_{\delta} {+} t_{sifs} {+} t_{C} {+} t_{\delta}
&\text{for RTS/CTS.}\\
\end{cases}
\end{equation*}
Using $p_F$ and $\tau$ defined as in (\ref{eqn:p_E}) and (\ref{eqn:tau}),
respectively, we calculate the expected slot length $t_{slot}$ as,
\begin{equation*}\setlength{\arraycolsep}{0pt}
t_{slot} {=}(1 {-} p_B) t_{\sigma} {+} p_B (1 {-} p_S) t_F {+} p_B p_S p_E t_F {+} p_B p_S (1 {-} p_E) t_S.
\end{equation*}
\begin{figure*}\centering
\subfloat[Effect of payload, $l_D$]{\includegraphics[width=0.33\textwidth]{packetSize}\label{fig:packetSize}}
\hfill
\subfloat[Effect of packet error rate, $p_E$]{\includegraphics[width=0.33\textwidth]{P_e-2}\label{fig:P_e-2}}
\hfill
\subfloat[Effect of number of stations, $n$]{\includegraphics[width=0.33\textwidth]{nn}\label{fig:nn}}
\caption{Simulation and analytical throughput for different (a) payload, (b) packet error rate,
and (c) number of stations.}
\label{fig:sim}
\end{figure*}
\subsection{Performance Measures}
Throughput $\psi$ is the number of payload bits that the MAC layer can transmit per second which is given by,
\begin{equation*}
\psi{=}\frac{p_Bp_S(l_D+l_I+l_U)}{t_{Slot}}.\\
\end{equation*}
where $l_D$, $l_I$, and $l_U$ are length of application data, IP header, and UDP/TCP header.
A packet is dropped when it suffers from more than $m$ number of failed
transmissions. Therefore, network packet loss $e_N$ is the probability that
a packet suffers from at least $m{+}1$ failed transmissions and is given by~\cite{Wu2002},
\begin{equation*}
e_N {=} p_F^{m{+}1}.
\end{equation*}
If queue length is not very large, queuing loss $e_Q$ can play an important
role for loss sensitive applications. For the $M/M/1/l_Q$ queue described earlier, $e_Q$ is given by,
\begin{equation*}
e_Q{=}\begin{cases}
\frac{(\lambda d_C)^{l_Q}{-}(\lambda d_C)^{l_Q{+}1}}{1-(\lambda d_C)^{l_Q{+}1}}&\text{when } \lambda d_C {\neq} 1,\\
\frac{1}{l_Q{+}1}
& \text{otherwise.}
\end{cases}
\end{equation*}
Channel access delay $d_C$ is defined to be the length of the period starting when a packet becomes the head of the queue (HoQ) and ending when an ACK frame
confirms its successful reception.
Delay faced by the dropped packets (because of exceeding retry limit) does not
contribute to it.
The probability that a packet is successfully transmitted is $1{-}e_N{=}1{-}p_F^{m+1}$.
A packet may face delay at each retry stage and the probability that it reaches
stage $i$ and is not dropped (i.e., it is successful) is given by $\frac{p_F^i{-}p_F^{m{+}1}}{1{-}p_F^{m{+}1}}$.
At retry stage $i$, a packet can face $\frac{W_i}{2}$ number of slots on average.
In total, the packet has to wait for $n_{slot}$ number of slots as HoQ which is given by,
\begin{equation*}
n_{slot} {=} \sum_{i{=}0}^{m}\frac{W_i {+} 1}{2} \frac{p_F^i{-}p_F^{m{+}1}}{1{-}p_F^{m+1}}.
\end{equation*}
Using the above expression for number of slots, the channel access delay can
be estimated as~\cite{Wu2002},
\begin{equation*}
d_C=n_{slot}t_{slot}.
\end{equation*}
\section{Simulation and Results}
\label{section:Simulation and Results}
We performed all mathematical analyses in Matlab 2007a and
modeled the simulations in Network Simulator 2 (NS-2.28) which is widely
used by network researchers.
We found a very agreeing match, as shown later, between the simulations and our theoretical datasets which validates our model.
NS-2's original tracing mechanism being slow and inadequate to trace a packet
at each network layer, a separate tracing mechanism is developed and used in this work.
To carefully study the effect of network load and packet arrival rate, a
separate Poisson agent is also developed along with a corresponding AP (Null) agent.
These classes were made open source and posted in NS-2 official mail group for comments and public use.
We tested the simulations for IEEE 802.11B with DSSS based physical layer but the method is applicable
to any 802.11 variant.
The parameters used in the simulation are shown in Table~\ref{table:simulation parameters}.
\begin{table}[!t]\renewcommand{\arraystretch}{1.3}\caption{Simulation parameters}\label{table:simulation parameters}\centering
\begin{tabular}{|c|c|c|c|}
\hline
Parameter & Value &\ Parameter & Value\\
\hline
SIFS & $10\mu s$ &
Idle slot & $20\mu s$\\
DIFS & $50\mu s$ &
Propagation delay & $1\mu s$\\
PLCP header & $144b$ &
Preamble & $48b$\\
Data rate & $1$ Mbps&
Basic rate & $1$ Mbps\\
PLCP rate & $1$ Mbps&
Run length& 2000s \\
\hline
\end{tabular}
\end{table}
Fig.~\ref{fig:sim}\subref{fig:packetSize} shows throughput $\psi$ as a function of packet arrival rate $\lambda$ for different payload sizes.
We used a wide range of payload sizes ($100{\sim}5000$ bytes).
The simulation results match mathematical data closely.
The throughput increases almost linearly with increasing arrival
rate up to a certain point, beyond which it does not increase any further.
The first region indicates unsaturated condition of the network where $\lambda {\ll} \frac{1}{d_C}$ while the second region indicates near saturation or saturation region where $\lambda {\geq} \frac{1}{d_C}$.
The smooth transition between the two regions is where network switches into saturation and $\lambda{\approx}\frac{1}{d_C}$ (e.g., $\lambda{=}20{\sim}30$ packets/sec for $l_D$=500B).
This transition is clearly a function of the payload and it is achieved at a lower $\lambda$ when the payload is larger.
\begin{figure*}\centering
\subfloat[Effect of initial collision window, $W$]{\includegraphics[width=0.33\textwidth]{A1}\label{fig:A1}}
\hfill
\subfloat[Effect of retry limit, $m$] {\includegraphics[width=0.33\textwidth]{A3A}\label{fig:A3A}}
\hfill
\subfloat[Effect of number of stations, $n$]{\includegraphics[width=0.33\textwidth]{A4A}\label{fig:A4A}}
\caption{Effect of (a) initial collision window, (b) retry limit, and (c) number
of stations on throughput and delay.}
\label{fig:math}
\end{figure*}
\begin{figure*}\centering
\subfloat[Effect of packet arrival rate, $\lambda$]{\includegraphics[width=0.33\textwidth]{B1}\label{fig:B1}}
\hfill
\subfloat[Effect of initial collision window, $W$]{\includegraphics[width=0.33\textwidth]{B4}\label{fig:B4}}
\hfill
\subfloat[Effect of retry limit, $m$]{\includegraphics[width=0.33\textwidth]{B7}\label{fig:B7}}
\caption{Effect of a) arrival rate, b) initial contention window, and c) retry limit on probability of collision $p_C$, probability of transmission $\tau$, probability of non-empty queue
$p_Q$, network loss $e_N$, and queuing loss $e_Q$. }
\label{fig:queue}
\end{figure*}
For a larger payload, the transmission time becomes longer, making the transmissions more susceptible to collisions and errors.
Thus the channel access time becomes longer saturating the network even with a smaller packet arrival rate.
A longer channel access time incurs relatively high degradation in throughput since the channel is kept busy for each packet in the queue.
However, the payload contributes even more to the total throughput and a greater payload still gives a greater throughput although channel access time is longer in this case.
This is the reason the rate of increase of throughput in the unsaturated region is higher for longer packets and the same level of throughput can be achieved by longer packets at a lower packet arrival rate.
For the rest of this study, we use $l_D=1000$ bytes as a representative data
payload size.
Fig.~\ref{fig:sim}\subref{fig:P_e-2} shows change in throughput for different packet error rates.
These results show that although the impact of packet error rate is considerably high in saturated region, its impact is negligible in unsaturated region.
In simulation, errors are introduced uniformly in the received packets.
Since the number of transmissions in the unsaturated region is quite low, less number of retries is necessary and hence the impact is found insignificant.
On the other hand, in the saturated region, a large number of transmissions and retransmissions take place; as a result the impact of erroneous packets on overall throughput becomes greater.
Fig.~\ref{fig:sim}\subref{fig:nn} presents results for different numbers of stations in the network.
Since the payload is kept constant to 1000 bytes, the saturation throughput is same in all cases which is 0.9 Mbps.
Below the saturation point, i.e., in the unsaturated region, the rate of increase in throughput is different for different network sizes.
Both the packet arrival rate and the number of nodes define the load on the network since the number of generated packets depends on them.
The number of nodes also means the number of competitors for the channel.
Therefore, even if the number of generated packets in unit time (and payload) is same, a greater number of nodes may lead to a greater number of collisions, which in turn will result in a higher channel access time and a lower throughput.
The results demonstrated the same phenomenon where, every thing else being unchanged, a higher number of nodes achieved saturation even with a much lower packet arrival rate.
While a $4$-node network did not achieve saturation until $\lambda{=}60$, a $15$-node network is saturated even at $\lambda{=}9$.
Therefore, network size plays an important role in attaining saturation level of a network, its impact being much higher than even packet arrival rate.
Since our model matches very closely with simulation results (as shown above),
in the rest of this paper we present analysis based on theoretical data derived
from the model.
Fig.~\ref{fig:math}\subref{fig:A1} presents the throughput and channel access delay as a function
of initial contention window $W$ for two packet error rates for the parameters
shown in figure.
Although the most common values of $W$ are $16$ and $32$, a contention window range of $[1{\ldots}100]$ is investigated to demonstrate its effect.
A low $W$ reduces backoff time which in turn increases collisions and decreases
throughput.
However, decrease in the idle period reduces overall channel access time.
Although number of stations and packet arrival rate are kept constant, $d_C$
is initially low but grows quickly
until $W{=}20$, e.g., $d_C{=}0.054$s and $0.065$s at $W{=}10$ and $20$ for $p_E{=}0$.
Increase in channel access delay becomes minimal for $W{>}70$.
Throughput shows a similar trend but apparently the effect of channel error
on throughput is
more aggressive compared to that of $W$.
Fig.~\ref{fig:math}\subref{fig:A3A} shows throughput and channel access delay for a range of retry limits in presence of channel errors.
Delay is found to be lower for a lower $m$.
When $m{=}0$, a packet will be dropped even after the first failure in its transmission.
With higher $m$, a higher number of failed transmissions attempts are tolerated.
Therefore, the HoQ packet can hog the channel for a longer period and channel access delay becomes higher.
On the contrary to $\infty$-limit BEB, where delay upper limit is high, $m$-retry
BEB keeps the channel access delay within a defined margin.
In delay sensitive applications (e.g., VoIP, video conference, etc.), a late packet is as good as a dropped packet.
Playing a late packet will only increase jitter at the receiver end.
Therefore, it is better to drop a late packet when it can help other packets to reach in time~\cite{Siddique2008}.
For the parameters shown in the figure, $d_C$ is found to be $66.9$ms for $m{=}0$, which reaches $67.2$ms for $m{=}3$ and
remains unchanged for higher values of $m$.
On the other hand, throughput is higher for lower $m$ (e.g., $0.6279$ mbps
for $m{=}1$), drops exponentially
with increasing $m$, and becomes nearly constant for $m{\geq}4$.
Fig.~\ref{fig:math}\subref{fig:A4A} elaborates throughput and channel access delay as a function of the number of stations for two packet arrival rates.
With $\lambda{=}2$, throughput is 260.027 kbps at $n{=}1$ which rises very quickly to 892.397 kbps at $n{=}5$ and remains unchanged for higher number
of nodes.
We omit discussion on higher number of stations since the network would then operate in saturated region while we mainly investigate the unsaturated condition in this
paper.
Interested readers may consult \cite{Chatzimisios2004} and the references therein for more discussions on the saturated condition.
With $\lambda{=}6$ the throughput is much higher initially (699.483 kbps at $n{=}1$) and quickly reaches 891.681 kbps at $n{=}3$.
As $n$ increases, throughput remains the same irrespective of $\lambda$.
For $\lambda{=}2$, channel access delay is initially very low ($d_C{=}462\mu s$) at $n{=}1$ which increases slowly at the beginning until $n{=}3$ ($d_C{=}200 \mu s$). Apparently, the offered network is too small to provide sufficient contention. After that, it increases almost linearly with $n$.
A steady state system is defined by transmission probability and collision probability. In Fig.~\ref{fig:queue}\subref{fig:B1}$\sim$~\ref{fig:queue}\subref{fig:B7}, collision probability $p_C$, transmission probability $\tau$, non-empty queue probability $p_Q$, network loss $e_N$, and queue loss $e_Q$ are shown for varying packet arrival rate, initial collision window, and retry limit.
With increase in packet arrival rate, as shown in Fig.~\ref{fig:queue}\subref{fig:B1}, the probability that the queue is non-empty rises very quickly and becomes
$1$ for $\lambda{>}20$.
With the queue always non-empty, further increase in packet arrival rate increases the number of packets in the queue.
When a newly generated packet finds the queue full, it will be dropped.
Therefore, soon queue loss will overwhelm the system and loss becomes very high.
That is why queuing loss $e_Q$ starts growing very quickly from $\lambda{=}20$.
But notably transmission probability $\tau$ remains unchanged after some initial increase up to $\lambda{=}20$.
Since the number of stations is constant, increase in $\lambda$ does not have additional impact once the nodes become nearly saturated.
Fig.~\ref{fig:queue}\subref{fig:B4}, on the other hand, elaborates the same performance measures for different values of contention window $W$.
A higher contention window gives greater spread of counter values over contenders ensuring less collision.
At the same time, nodes will wait longer in idle state and channel access time increases.
We find transmission probability to decrease with increasing collision window.
Probability of a non-empty queue $p_Q$ rises quickly and becomes $1$ as before but queuing loss is much smaller in this case.
Since the packet arrival rate and the number of stations are constant, queuing loss
$e_Q$ does not increase after reaching its highest value of 0.2931.
Transmission probability, collision probability, and network loss decrease
initially with increase in $W$ since nodes tend to spend more time idly and frequency of transmissions become lower.
Fig.~\ref{fig:queue}\subref{fig:B7} presents the above mentioned parameters for variation of retry limit.
At a lower $m$, a packet will be dropped after a smaller number of failures.
As a result, number of transmissions will be less and network loss will be higher.
However, it also means that channel access delay will have a lower upper-limit.
We find that transmission probability decreases initially and remains constant when the network configuration can transmit the offered load.
Collision probability and queuing loss follow a similar pattern of different magnitudes.
Since network loss is given by $p^{m+1}_F$, $e_N$ continues to decrease exponentially with increasing $m$ in the constant $p_C$ region.
Probability of the queue being non-empty is $1$ when packet arrival rate is greater than or equal to servicing rate of the queue i.e., $\lambda{\geqslant} \frac{1}{d_C}$ or $\lambda d_C {\geqslant} 1$.
Channel access delay decreases with increase in $m$ as discussed.
In particular, $d_C{=}0.209$ at $m{=}1$ and $d_C{=}0.0707$ at $m{=}10$.
Therefore, albeit $\lambda$ (${=}20$) is kept constant, $\lambda d_C {\geqslant} 1$ for the whole region and $p_Q{=}1$ for this configuration.
\section{Conclusion}
\label{section:Conclusion}
This paper introduces a Markov model analysis for IEEE 802.11 $m$-retry BEB
based DCF under unsaturated condition.
We investigated
the most widely used network performance measures in an imperfect channel
with provision for power capture.
The performance measures are presented in terms of
network parameters and will be of great assistance in designing and
assessing IEEE 802.11 based wireless networks.
WLANs tend to suffer from severe bottlenecks
formed around the AP under high load which can cause call drop, call
rejection~\cite{Siddique2008} for voice/video calls, and degradation of
VoIP voice quality.
Therefore, these networks should be carefully designed considering
expected load and network parameters.
This model considers both network load and configuration, and is equally
applicable for both saturated and unsaturated studies and, therefore,
can play a significant role in network planning.
\bibliographystyle{IEEEtran}
|
\section{Introduction and preliminary results}\label{sec:intro}
{\it Steganography} is a scientific discipline within the field known as {\it data hiding}, concerned with hiding information into a commonly used media, in such a way that no one apart from the sender and the intended recipient can detect the presence of embedded data. A comprehensive overview of the core principles and the mathematical methods that can be used for data hiding can be found in \cite{moulin}.
An interesting steganographic method is known as {\it matrix encoding}, introduced by Crandall~\cite{cra} and analyzed by Bierbrauer et al.~\cite{bier}. Matrix encoding requires the sender and the recipient to agree in advance on a parity check matrix $H$, and the secret message is then extracted by the recipient as the syndrome (with respect to $H$) of the received cover object. This method was made popular by Westfeld~\cite{west}, who incorporated a specific implementation using Hamming codes in his F5 algorithm, which can embed $t$ bits of message in $2^t - 1$ cover symbols by changing, at most, one of them.
\medskip
There are two parameters which help to evaluate the performance of a
steganographic method over a cover message of $N$ symbols: the \textit{average distortion} $D=\frac{R_a}{N}$, where $R_a$
is the expected number of changes over uniformly distributed messages; and the
\textit{embedding rate} $E= \frac{t}{N}$, which is the amount of bits that can be
hidden in a cover message. In general, for the same embedding rate a method is better when the
average distortion is smaller. Following the terminology used by Fridrich et al.~\cite{fri}, the pair $(D,E)$ will be called {\it $CI$-rate}.
Furthermore, as Willems et al. in~\cite{WiMa}, we will also assume that a discrete source produces a sequence $\mathbf{x}=(x_1,\ldots,x_N)$, where $N$ is the block length, each $x_i \in \aleph =\{0,1,\ldots,2^B-1\}$, and $B\in \{8,12,16\}$ depends on whether the source is a grayscale digital image, or a CD audio, etc. The message $\mathbf{s} \in \{1,\ldots,M\}$ we want to hide into a host sequence $\mathbf{x}$ produces a composite sequence $\mathbf{y} = f(\mathbf{x},\mathbf{s})$, where $\mathbf{y} =(y_1,\ldots y_N)$ and each $y_i\in \aleph$. The composite sequence $\mathbf{y}$ is obtained from distorting $\mathbf{x}$, and the distortion will be assumed to be a squared-error distortion (see~\cite{WiMa}). In these conditions, if information is only carried by the least significant bit (LSB) of each $x_i$, the appropriate solution comes from using binary Hamming codes~\cite{west}, improved using product Hamming codes~\cite{riri}. For larger magnitude of changes,
but limited to $1$, that is, $y_i=x_i +c$, where $c\in \{0,+1,-1\}$, the situation is called ``$\pm 1$-steganography", and the information is carried by the two least significant bits. It is known that the embedding becomes statistically detectable
rather quickly with the increasing amplitude of embedding
changes. Therefore, our interest goes to avoid changes of amplitude greater than one. With this assumption, our steganographic scheme will be compared with the upper bound from~\cite{WiMa} for the embedding rate in ``$\pm 1$-steganography", given by $H(D)+D$, where $H(D)$ is the binary entropy function $H(D)= -D\log_2(D) - (1-D)\log_2(1-D)$ and $0\leq D\leq 2/3$ is the average distortion. A main purpose of steganography is designing schemes in order to approach this upper bound.
In most of the previous papers, ``$\pm 1$-steganography" has involved a ternary coding problem. Willems et al.~\cite{WiMa} proposed a schemed based on ternary Hamming and Golay codes, which were proved to be optimal. Fridrich and Lison\v{e}k~\cite{fri} proposed a method based on rainbow colouring graphs which, for some values, outperformed the scheme obtained by direct sum of ternary Hamming codes with the same average distortion.
However, both methods from~\cite{WiMa} and~\cite{fri} show a problem when dealing with extreme grayscale values, since they suggest making a change of magnitude greater than one in order to avoid having to apply the change $x_i-1$ and $x_i+1$ to a host sequence of value $x_i=0$ and $x_i=2^B-1$, respectively. Note that the kind of change they propose would obviously introduce larger distortion and therefore make the embedding more statistically detectable.
In this paper we also consider the $\pm 1$-steganography. Our new method is based on perfect $\Z_2\Z_4$-linear codes which, although they are not linear, they have a representation using a parity check matrix that makes them as efficient as the Hamming codes. As we will show, this new method not only performs better than the one obtained by direct sum of ternary Hamming codes from~\cite{WiMa}, but it also deals better with the extreme grayscale values, because the magnitude of embedding changes is under no circumstances greater than one.
To make this paper self-contained, we review in Section~\ref{sec:additivecodes} a few elementary concepts on perfect $\Z_2\Z_4$-linear codes, relevant for our study. The new steganographic method is presented in Section~\ref{sec:stegoz2z4}, whereas an improvement to better deal with the extreme grayscale values problem is given in Section~\ref{sec:anomalies}. Finally, the paper is concluded in Section~\ref{sec:conclusions}.
\section{Perfect $\Z_2\Z_4$-linear codes}\label{sec:additivecodes}
In general, any non-empty subgroup ${\cal C}$ of ${\mathbb{Z}}_2^\alpha \times {\mathbb{Z}}_4^\beta$ is a {\it $\Z_2\Z_4$-additive code}, where ${\mathbb{Z}}_2^\alpha$ denotes the set of all binary
vectors of length $\alpha$ and ${\mathbb{Z}}_4^\beta$ is the set of all $\beta$-tuples in ${\mathbb{Z}}_4$. Let $C=\Phi({\cal C})$, where
$\Phi: {\mathbb{Z}}_2^{\alpha}\times{\mathbb{Z}}_4^{\beta} \longrightarrow {\mathbb{Z}}_2^{n}$ is given by the map $$\Phi(u_1, \ldots, u_{\alpha} | v_1, \ldots, v_{\beta})=(u_1, \ldots, u_{\alpha} | \phi(v_1), \ldots, \phi(v_{\beta})),$$ where $\phi(0)=(0,0)$, $\phi(1)=(0,1)$, $\phi(2)=(1,1)$, and $\phi(3)=(1,0)$ is the usual Gray map from ${\mathbb{Z}}_4$ onto ${\mathbb{Z}}_2^2$.
A $\Z_2\Z_4$-additive code ${\cal C}$ is also isomorphic to an abelian structure like
${\mathbb{Z}}_2^{\gamma}\times {\mathbb{Z}}_4^{\delta}$. Therefore, ${\cal C}$ has $|{\cal C}|=2^\gamma 4^\delta $ codewords, where $2^{\gamma+\delta}$ of them are of order two. We call such code ${\cal C}$ a
{\it $\Z_2\Z_4$-additive code of type $(\alpha,\beta;\gamma,\delta)$} and its
binary image $C$ is a {\it $\Z_2\Z_4$-linear code of type
$(\alpha,\beta;\gamma,\delta)$}. Note that the Lee distance of a $\Z_2\Z_4$-additive code ${\cal C}$ coincides with the Hamming distance of the $\Z_2\Z_4$-linear code $C=\Phi({\cal C})$, and that the binary code $C$ does not have to be linear.
The {\it ${\mathbb{Z}}_2{\mathbb{Z}}_4$-additive dual code} of ${\cal C}$, denoted by ${\cal
C}^\perp$, is defined as the set of vectors in ${\mathbb{Z}}_2^\alpha \times {\mathbb{Z}}_4^\beta$ that are orthogonal to every codeword in ${\cal C}$, being the definition of inner product in ${\mathbb{Z}}_2^{\alpha}\times {\mathbb{Z}}_4^{\beta}$ the following:
\begin{equation} \label{inner}
\langle u,v \rangle=2(\sum_{i=1}^{\alpha}
u_iv_i)+\sum_{j=\alpha+1}^{\alpha+\beta}
u_jv_j\in {\mathbb{Z}}_4, \end{equation}
where $u,v\in {\mathbb{Z}}_2^{\alpha}\times {\mathbb{Z}}_4^{\beta}$ and computations are made considering the zeros and ones in the $\alpha$ binary coordinates as quaternary zeros and ones, respectively.
The binary code $C_\perp=\Phi({\cal C}^\perp)$, of length $n=\alpha+2\beta$, is called the {\it
${\mathbb{Z}}_2{\mathbb{Z}}_4$-dual code} of $C$.
A $\Z_2\Z_4$-additive code ${\cal C}$ is said to be {\it perfect} if code $C=\Phi({\cal C})$ is a perfect $\Z_2\Z_4$-linear code, that is all vectors in ${\mathbb{Z}}_2^{n}$ are within distance one from a codeword and the distance between two codewords is, at least, three.
For any $m\geq 2$ and each $\delta$ $\in$ $\{0,\ldots,\lfloor \frac{m}{2}\rfloor \}$ there exists a perfect $\Z_2\Z_4$-linear code $C$ of binary length $n=2^m-1$, such that its $\Z_2\Z_4$-dual code is of type $(\alpha,\beta;\gamma,\delta)$, where $\alpha=2^{m-\delta}-1$, $\beta=2^{m-1}-2^{m-\delta-1}$ and $\gamma=m-2\delta$ (note that the binary length can be computed as $n=\alpha+2\beta$). The above result is due to~\cite{br} and it allows us to write the parity check matrix $H$ of any $\Z_2\Z_4$-additive perfect code for a given value of $\delta$. Matrix $H$ can be represented taking all possible vectors in ${\mathbb{Z}}_2^\gamma \times {\mathbb{Z}}_4^\delta$, up to sign changes, as columns. In this representation, there are $\alpha$ columns which correspond to the binary part of vectors in ${\cal C}$, and $\beta$ columns of order four which correspond to the quaternary part. We agree on a representation of the $\alpha$ binary coordinates as coordinates in $\{0,2\} \in {\mathbb{Z}}_4$.
\section{Steganography based on perfect ${\mathbb{Z}}_2{\mathbb{Z}}_4$-linear codes}\label{sec:stegoz2z4}
Take a perfect $\Z_2\Z_4$-linear code and consider its $\Z_2\Z_4$-dual, which is of type $(\alpha,\beta;\gamma,\delta)$. As stated in the previous section, this gives us a parity check matrix $H$ which has $\gamma$ rows of order two and $\delta$ rows of order four.
For instance, for $m=4$ and according to~\cite{br}, there are three different $\Z_2\Z_4$-additive perfect codes of binary length $n=2^4-1=15$ which correspond to the possible values of $\delta\in \{0,\ldots,\lfloor \frac{m}{2}\rfloor \} = \{0,1,2\}$. For $\delta=0$, the corresponding $\Z_2\Z_4$-additive perfect code is the usual binary Hamming code, while for $\delta=2$ the $\Z_2\Z_4$-additive perfect code has parameters $\alpha=3$, $\beta=6$, $\gamma=0$, $\delta=2$ and the following parity check matrix:
\begin{equation}\label{pcm}
H=\left ( \begin{array}{ccc|cccccc}
2&0&2 & 0&1&1&1&1&2\\
2&2&0 & 1&0&1&2&3&1
\end{array}\right ).
\end{equation}
Let $\mathbf{h}_i$, for $i \in \{1,\ldots,\alpha+\beta \}$, denote the $i$-th column vector of $H$.
Note that the all twos vector ${\mathbf{2}}$ is always one of the columns in $H$ and, for the sake of simplicity, it will be written as the column $\mathbf{h}_1$. We group the remaining first $\alpha$ columns in $H$ in such a way that, for any $2\leq i \leq (\alpha+1)/2$, the column vector $\mathbf{h}_{2i}$ is paired up with its complementary column vector $\bar{\mathbf{h}}_{2i} = \mathbf{h}_{2i+1}$, where $\bar{\mathbf{h}}_{2i}=\mathbf{h}_{2i}+{\mathbf{2}}$.
\medskip
To use these perfect $\Z_2\Z_4$-additive codes in steganography take $N=2^{m-1}=\frac{\alpha+1}{2}+\beta$ and let $x=(x_1,\ldots,x_N)$ be an $N$-length source of grayscale symbols such that $x_i \in \aleph=\{0,1,\ldots,2^B-1\}$, where, for instance, $B=8$ for grayscale images.
We assume that a grayscale symbol $x_i$ is represented as a binary vector $(v_{7i},\ldots,v_{1i},v_{0i})$ such that
\begin{equation}\label{representation}
x_i=\sum_{j=0}^{B/2-1}\phi^{-1}(v_{(2j+1)i} \;,\; v_{(2j)i}) \cdot 4^{j},
\end{equation}
where $\phi^{-1}()$ is the inverse of Gray map. We will use the two least significant bits (LSBs), $v_{1i},v_{0i}$, of every grayscale symbol $x_i$ in the source, for $i > 1$, as well as the least significant bit $v_{01}$ of symbol $x_1$ to embed the secret message.
\medskip
Each symbol $x_i$ will be associated with one or more column vectors $\mathbf{h}_i$ in $H$, depending on the grayscale symbol:
\begin{enumerate}
\item Grayscale symbol $x_1$ is associated with column vector $\mathbf{h}_1$ by taking the least significant bit $v_{01}$ of $x_1$.
\item Grayscale symbol $x_i$, for $2\leq i \leq (\alpha+1)/2$, is associated with the two column vectors $\mathbf{h}_i$ and $\bar{\mathbf{h}}_i$, by taking, respectively, the two least significant bits, $v_{1i}, v_{0i}$, of $x_i$.
\item Grayscale symbol $x_j$, for $\alpha < j \leq N$, is associated with column vector $\mathbf{h}_{j+(\alpha-1)/2}$ by taking its two least significant bits $v_{1j},v_{0j}$ and interpreting them as an integer number $\phi^{-1}(v_{1j},v_{0j})$ in ${\mathbb{Z}}_4$.
\end{enumerate}
In this way, the given $N$-length packet $x$ of symbols is translated into a vector $\mathbf{w}$ of $\alpha$ binary coordinates and $\beta$ quaternary coordinates.
The embedding process we are proposing is based on the matrix encoding method~\cite{cra,west}. The secret message can be any vector $\mathbf{s}$ $\in$ ${\mathbb{Z}}_2^{\gamma}\times {\mathbb{Z}}_4^{\delta}$. Vector $\epsilon \cdot \mathbf{h}_i$ indicates the changes needed to embed $\mathbf{s}$ within $x$; that is $H{\mathbf{w}}^T + \epsilon \cdot \mathbf{h}_i=\mathbf{s}$, where $\epsilon$ is an integer whose value will be described bellow, $H{\mathbf{w}}^T$ is the syndrome vector of $\mathbf{w}$ and $\mathbf{h}_i$ is a column vector in $H$. The following situations can occur, depending on which column $\mathbf{h}_i$ needs to be modified:
\begin{enumerate}
\item If $\mathbf{h}_i=\mathbf{h}_1$, then the embedder is required to change the least significant bit of $x_1$ by adding or substracting one unit to/from $x_1$, depending on which operation will flip its least significant bit, $v_{01}$.
\item If $\mathbf{h}_i$ is among the first $\alpha$ column vectors in $H$ and $2 \leq i \leq \alpha$, then $\epsilon$ can only be $\epsilon=1$. In this case, since $\mathbf{h}_i$ was paired up with its complementary column vector $\bar{\mathbf{h}}_i$, then this situation is equivalent to make $(v_{1i},1+v_{0i})$ or $(1+v_{1i},v_{0i})$, where $v_{1i}$ and $v_{0i}$ are the least significant bits of the symbol $x_i$ which had been associated with those two column vectors. Hence, after the inverse of Gray map, by changing one or another least significant bit we are actually adding or subtracting one unit to/from $x_i$. Note that a problem may crop up at this point when we need to add $1$ to a symbol $x_i$ of value $2^B-1$ or, likewise, when $x_i$ has value $0$ and we need to subtract $1$ from it.
\item If $\mathbf{h}_i$ is one of the last $\beta$ columns in $H$ we can see that this situation corresponds to add $\epsilon \in \{0,1,2,3\}$ to $x_{i-(\alpha-1)/2}$. Note that because we are using a $\Z_2\Z_4$-additive perfect code, $\epsilon$ will never be $2$. Hence, the embedder should add ($\epsilon=1$) or subtract ($\epsilon=3$) one unit to/from symbol $x_{i-(\alpha-1)/2}$. Once again, a problem may arise with the extreme grayscale values.
\end{enumerate}
\bigskip
\begin{ex} \label{ex:noAnomalies} Let $x=(239,251,90,224,226,187,229,180)$ be an $N$-length source of grayscale symbols, where $x_i \in \{0,\ldots,255\}$ and $N=8$, and let $H$ be the matrix in (\ref{pcm}). The source $x$ is then translated into the vector $\mathbf{w}=(010|202310)$ in the way specified above. Let $\mathbf{s}={0 \choose 2}$ be the vector representing the secret message we want to embed in $x$. We then compute $H{\mathbf{w}}^T={2 \choose 3}$ and see, by the matrix encoding method, that $\epsilon=3$ and $\mathbf{h}_i=\mathbf{h}_9$. According to the method just described, we should apply the change $x_8-1$. In this way, $x_8$ becomes $x_8=179$, and then $\mathbf{w}=(010|202313)$, which has the expected syndrome ${0 \choose 2}$.
\end{ex}
\medskip
As already mentioned at the beginning of this paper, the problematic cases related to the extreme grayscale values are also present in the methods from \cite{fri} and \cite{WiMa}, but their authors assume that the probability of gray value saturation is not too large. We argue that, though rare, this gray saturation can still occur. However, in order to compare our proposal with these others we will not consider these problems either until the next section.
Therefore, we proceed to compute the values of the average distortion $D$ and the embedding rate $E$.
Our method is able to hide any secret vector $\mathbf{s}$ $\in$ ${\mathbb{Z}}_2^{\gamma}\times {\mathbb{Z}}_4^{\delta}$ into the given $N$ symbols. Hence, the embedding rate is $(\gamma+2\delta)$ bits per $N$ symbols, $E=\displaystyle \frac{\gamma+2\delta}{N}=\frac{m}{2^{m-1}}$.
Concerning the average distortion $D$, we are using a perfect code of binary length $2^{m}-1$, which corresponds to $N=2^{m-1}$ grayscale symbols. There are $N-1$ symbols $x_i$, for $2 \leq i \leq N$, with a probability $2/2^{m}$ of being subjected to a change; a symbol $x_1$ with a probability $1/2^{m}$ of being the one changed; and, finally, there is a probability of $1/2^{m}$ that neither of the symbols will need to be changed to embed $\mathbf{s}$. Hence, $D=\displaystyle \frac{2N-1}{N2^{m}}=\frac{2^m-1}{2^{2m-1}}$.
The described method has a $CI$-rate $(D_m,E_m) = \left ( \displaystyle \frac{2N-1}{2N^2}, \frac{1+\log(N)}{N} \right )$, where $N=2^{m-1}$ and $m$ is any integer $m\geq 2$.
We are able to generate a specific embedding scheme for any value of $m$ but not for any $CI$-rate.
With the aim of improving this situation, convex combinations of $CI$-rates of two codes related to their direct sum are extensively treated in~\cite{fri}. Actually, it is possible to choose the $D$ coordinate and cover more $CI$-rates by taking convex combinations. Therefore, if $D$ is a non-allowable parameter for the average distortion we can still take $D_1 < D < D_2$, where $D_1,D_2$ are two contiguous allowable parameters, and by means of the direct sum of the two codes with embedding rate $E_1$ and $E_2$, respectively, we can obtain a new $CI$-rate $(D,E)$, with $D=\lambda D_1+(1-\lambda) D_2$ and $E=\lambda E_1+(1-\lambda) E_2$. From~a graphic point of view, this is equivalent to draw a line between two contiguous points $(D_1,E_1)$ and $(D_2,E_2)$, as it is shown in \figurename~\ref{fig:graphicWoAnom}.
In the following theorem we claim that the $CI$-rate of our method improves the one given by direct sum of ternary Hamming codes from~\cite{WiMa}.
\begin{theo}
For $m\geq 4$, the $CI$-rate given by the method based on $\Z_2\Z_4$-additive perfect codes improves the $CI$-rate obtained by direct sum of ternary Hamming codes with the same average distortion.
\end{theo}
\begin{IEEEproof}
Optimal embedding (of course, in the allowable values of $D$) can be obtained by using ternary codes, as it is shown in~\cite{WiMa}. The $CI$-rate of these codes is
$(D_\mu,E_\mu) = \left ( \displaystyle \frac{2}{3^\mu}, \frac{2\mu}{3^\mu-1} \right )$
for any integer $\mu$. Our method, based on $\Z_2\Z_4$-additive perfect codes, has $CI$-rate $(D_m,E_m) = \left ( \displaystyle \frac{2N-1}{2N^2}, \frac{1+\log(N)}{N} \right )$, for any integer $m\geq 2$ and $N=2^{m-1}$.
Take, for any $m\geq 2$, two contiguous values for $\mu$ such that $D_{\mu+1} < D_m < D_{\mu}$ and write $D_m=\lambda D_{\mu+1}+(1-\lambda) D_\mu$, where $0\leq \lambda \leq 1$.
We want to prove that, for $m\geq 4$, we have $E_m \geq \lambda E_{\mu+1}+(1-\lambda) E_\mu$, which is straightforward. However, since it is neither short nor contributes to the well understanding of the method, we do not include all computations here. The graphic bellow compares the $CI$-rate of the method based on ternary Hamming codes with that one based on $\Z_2\Z_4$-additive perfect codes. As one may see in this graphic, for some values of the average distortion $D$, the scheme based on $\Z_2\Z_4$-additive perfect codes has greater embedding rate $E$ than the one based on ternary Hamming codes.
\end{IEEEproof}
\textbf{Remark:} The same argumentation can be used and the same conclusion can be reached taking $q$ instead of $3$ and comparing our method with the method described in \cite{fri}.
\begin{figure}[htp]
\centering
\includegraphics[scale=0.4]{graphicWithoutAnomalies_Mod.png}
\caption{$CI$-rate $(D,E)$, for $B=8$, of steganographic methods based on ternary Hamming codes and on $\Z_2\Z_4$-additive perfect codes.}\label{fig:graphicWoAnom}
\end{figure}
\section{Solving the extreme grayscale values problem}\label{sec:anomalies}
In Section~\ref{sec:stegoz2z4} we described a problem which may raise when, according to our method, the embedder is required to add one unit to a source symbol $x_i$ containing the maximum allowed value ($2^B-1$), or to substract one unit from a symbol $x_i$ containing the minimum allowed value, $0$. To face this problem, we will use the complementary column vector $\bar{\mathbf{h}}_i$ of columns $\mathbf{h}_i$ in matrix $H$, where $\bar{\mathbf{h}}_i = 3\mathbf{h}_i + \textbf{2}$ and $\mathbf{h}_i$ is among the last $\beta$ columns in $H$. Note that $\mathbf{h}_i$ and $\bar{\mathbf{h}}_i$ can coincide.
The first $\alpha$ column vectors in $H$ will be paired up as before, and the association between each $x_i$ and each column vector $\mathbf{h}_i$ in $H$ will be also the same as in Section~\ref{sec:stegoz2z4}. However, given an $N$-length source of grayscale symbols
$x=(x_1,\ldots,x_N)$, a secret message $\mathbf{s} \in {\mathbb{Z}}_2^{\gamma}\times {\mathbb{Z}}_4^{\delta}$ and the vector $\epsilon \cdot \mathbf{h}_i$, such that $H{\mathbf{w}}^T + \epsilon \cdot \mathbf{h}_i= \mathbf{s}$, indicating the changes needed to embed $\mathbf{s}$ within $x$, we can now make some variations on the kinds of changes to be done for the specific problematic cases:
\begin{itemize}
\item If $\mathbf{h}_i$ is among the first $\alpha$ columns in $H$, for $2 \leq i \leq \alpha$, and the embedder is required to add $1$ to a symbol $x_i=2^B-1$, then the embedder should instead substract $1$ from $x_i$ as well as perform the appropiate operation ($+1$ or $-1$) over $x_1$ to have $v_{01}$ flipped. Likewise, if the embedder is required to substract $1$ from a symbol $x_i=0$, then (s)he should instead add $1$ to $x_i$ and also change $x_1$ to flip $v_{01}$.
\item If $\mathbf{h}_i$ is one of the last $\beta$ columns in $H$, and the embedder has to add $1$ to a symbol $x_i=2^B-1$, (s)he should instead substract $1$ from the grayscale symbol associated to $\bar{\mathbf{h}}_i$ and also change $x_1$ to flip $v_{01}$.
If the method requires substracting $1$ from $x_i=0$, then we should instead add $1$ to the symbol associated to $\bar{\mathbf{h}}_i$ and, again, change $x_1$ to flip $v_{01}$.
\end{itemize}
\bigskip
\begin{ex}\label{ex:Anomalies} Let $x=(239,251,90,224,226,187,229,0)$ be an $N$-length source of grayscale symbols, where $x_i \in \{0,\ldots,255\}$ and $N=8$, and let $H$ be the matrix (\ref{pcm}). As in Example~\ref{ex:noAnomalies}, the packet $x$ is translated into vector $\mathbf{w}=(010|202310)$, and $\mathbf{s}={0 \choose 2}$. However, note that now we are not able to make $x_8-1$ because $x_8=0$. Instead of this, we will add one unit to $x_3$, which is the symbol associated with $\bar{\mathbf{h}}_9=\mathbf{h}_4$, and substract one unit from $x_1$ so as to have its least significant bit flipped. Therefore, we obtain $x=(238,251,91,224,226,187,229,0)$ and then $\mathbf{w}=(110|302310)$, which has the desired syndrome.
\end{ex}
\medskip
The method above described has the same embedding rate $E=\displaystyle \frac{m}{2^{m-1}}$ as the one from Section~\ref{sec:stegoz2z4} but a slightly worse average distortion. We will take into account the squared-error distortion defined in~\cite{WiMa} for our reasoning.
As before, among the total number of grayscale symbols $N=2^{m-1}$, there are $N-1$ symbols $x_i$, for $2 \leq i \leq N$, with a probability $2/2^{m}$ of being changed; a symbol $x_1$ with a probability $1/2^{m}$ of being the one changed; and, finally, there is a probability of $1/2^{m}$ that neither of the symbols will need to be changed.
As one may have noted in this scheme, performing a certain change to a symbol $x_i$, associated with a column $\mathbf{h}_i$ in $H$, has the same effect as performing the opposite change to the grayscale symbol associated with $\bar{\mathbf{h}}_i$ and also changing the least significant bit $v_{01}$ of $x_1$. This means that with probability $\frac{2^B-2}{2^B}$ we will change a symbol $x_i$, for $2 \leq i \leq N$, a magnitude of $1$; and with probability $\frac{2}{2^B}$ we will change two other symbols also a magnitude of $1$. Therefore, $R_a = \displaystyle (N-1) \frac{2}{2^m} \left ( \frac{2^B-2}{2^B}+2\frac{2}{2^B} \right ) + \frac{1}{2^m}$ and the average distortion is thus $D= \displaystyle \frac{ 2N - 1 + \frac{N-1}{2^{B-2}} }{ N2^m }$. Hence, the described method has $CI$-rate $\displaystyle (D_m,E_m) = \left ( \frac{2N-1 + \frac{N-1}{2^{B-2}}}{2N^2} , \frac{1+\log(N)}{N} \right ) $.
\medskip
As we have already mentioned, the problem of grayscale symbols with $0$ and $2^B-1$ values was previously detected in both \cite{fri} and \cite{WiMa}. With the aim of providing a possible solution to this problem, the authors suggested to perform a change of a magnitude greater than $1$. However, the effects of doing this were are out of the scope of $\pm 1$-steganography.
In the remainder of this section we proceed to compare the $CI$-rate of our method with the $CI$-rate that those methods would have if their proposed solution was implemented.
\medskip
The scheme presented by Willems et al.~\cite{WiMa} is based on ternary Hamming codes, which are known to have length $n=(3^\mu-1)/2$, where $\mu$ denotes the number of parity check equations. Let us assume that whenever the embedder is required to perform a change ($+1$ or $-1$) that would lead the corresponding symbol $x_i$ to a non-allowed value, then a change of magnitude $2$ ($-2$ or $+2$) is made instead. While the embedding rate $E$ of this scheme would still be $E=\displaystyle \frac{2\mu \log(3)}{3^\mu-1}$, the average distortion $D$ would no longer be $D=\frac{2}{3^\mu}$. The actual expected number of changes $R_a$ is computed by noting that a symbol will be changed with probability $\frac{3^\mu-1}{3^\mu}$, and will not with probability $\frac{1}{3^\mu}$. Among the cases in which a symbol would need to be changed, there is a probability of $\frac{2^B-2}{2^B}$ that a symbol will be changed a magnitude of $1$, and a probability of $\frac{2}{2^B}$ that it will be changed a magnitude of $2$. By the squared-error distortion, $R_a=\frac{3^\mu-1}{3^\mu} \left ( \frac{2^B-2}{2^B}\cdot1+\frac{2}{2^B}\cdot2^2 \right )$ and therefore $D=\displaystyle \frac{2}{3^\mu} \left ( 1+\frac{3}{2^{B-1}} \right )$.
\medskip
Fridrich and Lison\v{e}k propose in their paper to pool the grayscale symbols source $x$ into cells of size $d$, then rainbow colour these cells and apply a $q$-ary Hamming code, where $q=2d+1$ is a prime power.
They measure the distortion by counting the maximum number of embedding changes, thus just considering the covering radius of the $q$-ary Hamming codes. However, we will now consider the average number of embedding changes (see~\cite{friEff}). As Willems et at., the authors from \cite{fri} also suggest to perform a change of magnitude $q-1>1$ to solve the extreme grayscale values problem. If this is done, the embedding rate would still be the same, $E=\displaystyle \frac{2\mu \log(q)}{q^\mu-1}$, but the average distortion would now be $D=\frac{2}{q^{\mu}} \left ( \frac{2^B-2}{2^B} + \frac{2}{2^B} \cdot (q-1)^2 \right ) = \displaystyle \frac{2}{q^{\mu}} \left ( 1 + \frac{q(q-2)}{2^{B-1}} \right )$.
\medskip
One can see in \figurename~\ref{fig:graphicWAnom} how our steganographic method for $\Z_2\Z_4$-additive perfect codes deals with the extreme grayscale values problem, for some values of $D$, better than those using ternary Hamming codes ($q=3$) from \cite{fri} and \cite{WiMa}.
\begin{figure}[htb!]
\centering
\includegraphics[scale=0.4]{graphicWithAnomalies_Mod.png}
\caption{$CI$-rates $(D,E)$, for $B=8$, of steganographic methods based on ternary Hamming codes and on $\Z_2\Z_4$-additive perfect codes, when they are dealing with the extreme grayscale values problem described in Section~\ref{sec:anomalies}.}\label{fig:graphicWAnom}
\end{figure}
\section{Conclusions}\label{sec:conclusions}
In this paper, we have presented a new method for $\pm 1$-steganography, based on perfect $\Z_2\Z_4$-linear codes. These codes are non-linear but still there exists a parity check matrix representation that makes them efficient to work with.
As we have shown in sections~\ref{sec:stegoz2z4} and~\ref{sec:anomalies}, this new scheme outperforms the one obtained by direct sum of ternary Hamming codes (see~\cite{WiMa}) as well as the one obtained after rainbow colouring graphs by using $q$-ary Hamming codes for $q=3$.
If we consider the special cases in which the technique might require to substract one unit from a grayscale symbol containing the minimum allowed value, or to add one unit to a symbol containing the maximum allowed value, our method performs even better than those aforementioned schemes. This is so because unlike them, our method never applies any change of magnitude greater than $1$, but two changes of magnitude $1$ instead, which is better in terms of distortion. Therefore, our method makes the embedding less statistically detectable.
As for further research, since the approach based on product Hamming codes in~\cite{riri} improved the performance of basic LSB steganography and the basic $F5$ algorithm, we would also expect a considerable improvement of the $CI$-rate by using product $\Z_2\Z_4$-additive codes or subspaces of product $\Z_2\Z_4$-additive codes in $\pm 1$-steganography.
|
\section{Introduction}
Even though the concept of inflation was proposed
almost thirty years ago \cite{Guth:1981}, \cite{Linde:1982}, \cite%
{Albrecht:1982} only a handful of models implementing it can be considered
to have a compelling physical basis. Essentially the problem arises due to
the fact that the inflaton has to be a very light field with mass less than
the Hubble expansion parameter during inflation. Quantum or gravitational
contributions to the inflaton mass are typically too large unless there is
an underlying symmetry protecting the inflaton from such corrections.
Although supergravity can protect against large radiative contributions the gravitational corrections are typically of order the Hubble
scale and too large (the \textquotedblleft eta\textquotedblright problem); only a Goldstone (shift) symmetry can avoid such
contributions. In this case the inflaton is a pseudo Goldstone boson, its
small mass being due to the breaking of the continuous Goldstone symmetry by
an anomaly at the quantum level or by explicit breaking.
The original inflationary model based on a pseudo Goldstone inflaton was the \textit{Natural Inflation} model of
Freese \textit{et al.} \cite{Freese:1990rb,Freese:1993,Freese:2008if}. It was based
on an anomalous Abelian symmetry such that the quantum anomaly generated mass for the would-be Goldstone mode. A potential problem for the model is due to the
fact that, in order to generate sufficient inflation, the scale of symmetry
breaking must be bigger than the Planck scale and in this case there may be
large Quantum Gravity corrections of $O(f/M_{Planck}).$ \footnote{In models with
additional scalar fields it is possible to avoid such super Plankian scales%
\cite{Kim:2004rp,Dimopoulos:2005ac} while maintaining the \textit{Natural Inflation} form of the inflaton potential with an effective super Plankian scale $f.$}
Models of \textit{Natural Inflation} with explicit breaking of the Goldstone symmetry
have received relatively little attention. Small terms explicitly breaking
the continuous symmetry arise if the underlying symmetry is a discrete
symmetry that gets promoted to a continous one if only the low dimension
(renormalisable) terms are kept in the Lagrangian. In this case the higher
dimension terms explicitly breaking the continuous symmetry are suppressed
by a large inverse mass scale and are naturally small. In these models to avoid the
need for a super-Plankian scale, $f,$ the end of inflation must be triggered
by a second scalar field \cite{Linde:1994}. In this case
the number of e-folds of inflation depends on an additional parameter
allowing for viable models with a sub-Plankian value for $f$ \cite{Shaun:2008}. It proves to
be impossible to arrange for such hybrid natural inflation based on an Abelian
global symmetry but models of hybrid natural inflation have been developed
with an underlying non-Abelian discrete symmetry \cite{Stewart:2000pa,Graham:2009}. In this letter we explore the phenomenology of these models
in detail, concentrating on the possibility that the scale of inflation may be very low. To achieve low scale inflation the hybrid field (which is not a pseudo Goldstone boson) must be light and this in turn requires that the underlying theory should be supersymmetric to protect the mass of the hybrid field from large radiative corrections.
We start with a model independent discussion of the structure and phenomenological implications of hybrid natural inflation before turning to a more detailed discussion of a particular supersymmetric model.
\section{Slow-roll parameters and observables}
As discussed in \cite{Graham:2009}, the terms of the (hybrid) inflaton
potential relevant when density perturbations are being produced have a
simple universal form corresponding to the slow roll of a single inflaton
field $\phi $:%
\begin{equation}
V\simeq V_0\left( 1+a\cos \left( \frac{\phi }{f}\right) \right) ,
\label{potential}
\end{equation}%
where $a$ is a constant. \textit{Natural Inflation} corresponds to the case $a=1$
while \textit{Hybrid Natural Inflation} has $a<1.$ We first list the detailed
expressions for the observables, the spectral index $n_{s},$ the density
perturbation at wave number $k$, the tensor to scalar perturbations ratio, $%
r,$ and the \textquotedblleft running\textquotedblright\ index $n_{r}.$
These are described in terms of the scale $\phi _{\mathrm{H}}$ at which the
perturbations are produced, some $40- 60$ e-folds before the end of
inflation, together with the scale of inflation given in terms of $V_{0}$
and the parameters $a$ and $f.$ We will discuss later the model constraints
on these parameters. \textit{Hybrid Natural Inflation} is consistent with recent observational bounds coming from the five years data of the WMAP team \cite{Komatsu:2008}, \cite{Dunkley:2008} and those from BAO \cite{Percival:2007} and SN surveys \cite{Riess:2004}, \cite{Astier:2006}, \cite{Woods-Vasey:2006}. A Table displaying these results can be found in \cite{Graham:2009}.
The observables are given in terms of the usual slow-roll parameters \cite%
{Liddle:2000cg}.
\begin{eqnarray}
\epsilon &\equiv & \frac{M^{2}}{2}\left( \frac{V^{\prime }}{V}\right) ^{2}\simeq%
\frac{1}{2}M^{2}\frac{a^{2} }{f^{2}}\sin ^{2}\left( \frac{\phi }{f}\right)%
,\qquad \\
\eta &\equiv &M^{2}\frac{V^{\prime \prime }}{V}\simeq-M^{2}\frac{a }{f^{2}}\cos \left(
\frac{\phi}{f}\right), \\
\xi &\equiv &M^{4}\frac{V^{\prime }V^{\prime \prime \prime }}{V^{2}}\simeq-M^{4}%
\frac{a^{2} }{f^{4}}\sin ^{2}\left( \frac{\phi}{f}\right)=-2\left( \frac{M}{f}\right)^2\epsilon,
\end{eqnarray}%
where $M$ is the reduced Planck scale, $M=2.44\times 10^{18}$~GeV.
In terms of these the observables are given by
\begin{eqnarray}
r &=&16\epsilon , \\
n_{\mathrm{s}} &=&1+2\eta -6\epsilon , \label{spectral} \\
n_{\mathrm{r}} &=&16\epsilon \eta -24\epsilon ^{2}-2\xi, \label{run} \\
\delta _{\mathrm{H}}^{2}(k) &=&\frac{1}{150\pi ^{2}}\frac{V_{\mathrm{H}}}{%
\epsilon _{\mathrm{H}}M^{4}} .
\end{eqnarray}
\subsection{\textit{Natural Inflation.}}
For the case of \textit{Natural Inflation} $\phi $ is the only scalar field and the
end of inflation is determined by the potential of Eq.(\ref{potential}) with $%
a=1.$ Thus the inflaton value, $\phi _{e},$ at the end of inflation is
determined and, up to an uncertainty about the number of intermediate
e-folds of inflation, hence so too is the inflaton value, $\phi _{\mathrm{H}}$, at
the time the density perturbations relevant today leave the horizon. As a
result \textit{Natural Inflation} has only two free parameters, the scale of inflation $\Delta $ \ and $f$.
The spectral index is always less than one and using the WMAP 3-year data
gives the bound \cite{Freese:2008if} $f>3.5M$. For $f$ close to this bound
the density fluctuations constrain the scale of inflation to lie in the
range $\Delta =10^{15}GeV-10^{16}GeV.$ With this $r$ and $n_{r}$ are
determined
\begin{equation}
r=4\left( \delta _{\mathrm{ns}}-\frac{M^{2}}{f^{2}}\right) ,
\label{indicesnatr}
\end{equation}%
\begin{equation}
n_{r}=-\frac{1}{2}\left( \delta _{\mathrm{ns}}^{2}-\frac{M^{4}}{f^{4}}%
\right) , \label{indicesnatnr}
\end{equation}%
where $\delta _{\mathrm{ns}}\equiv 1-n_{s}.$ This implies $r\leq 0.2$ and $%
\left\vert n_{r}\right\vert <10^{-3}$, where we have used the value $n_{s}\geq 0.947$ \cite {Komatsu:2008}, \cite {Dunkley:2008} .
\subsection{\textit{Hybrid Natural Inflation.}}
For the case of \textit{Hybrid Natural Inflation} $\phi _{e}$ is determined by the
hybrid sector of the theory so $\phi _{e},$ or more conveniently $\phi
_{\mathrm{H}},$ is essentially a free parameter. Thus \textit{Hybrid Natural Inflation} has
four free parameters namely $\Delta $, $f$, $a$ and $\phi _{\mathrm{H}}$. Using the
equation for the spectral index Eq.(\ref{spectral}) at $\phi _{\mathrm{H}}$
we find
\begin{equation}
\delta _{\mathrm{ns}}=\frac{2a}{f^{2}}\frac{c_{\mathrm{H}}}{(1+ac_{\mathrm{H}%
})}+\frac{3a^{2}}{f^{2}}\frac{(1-c_{\mathrm{H}}^{2})}{(1+ac_{\mathrm{H}})^{2}%
}, \label{spectral2}
\end{equation}%
where $c_{\mathrm{H}}\equiv \cos \left(
\frac{\phi _{\mathrm{H}}}{f}\right) .$ Then
\begin{equation}
f^{2}\delta _{\mathrm{ns}}\equiv z=\frac{a(2c_{\mathrm{H}}+a(3-c_{\mathrm{H}%
}^{2}))}{(1+ac_{\mathrm{H}})^{2}} , \label{z}
\end{equation}%
showing that $f\varpropto \sqrt{a}$ for small $a.$
At $\phi _{\mathrm{H}}$ the solutions to Eq.(\ref{z}) are given by
\begin{equation}
c_{1\mathrm{H}}=\frac{1-z+\sqrt{1+3a^{2}-3\left( 1-a^{2}\right) z}}{a\left(
1+z\right) },\text{\ }a\geqslant \frac{1}{3} \label{solution1}
\end{equation}%
and
\begin{equation}
c_{2\mathrm{H}}=\frac{1-z-\sqrt{1+3a^{2}-3\left( 1-a^{2}\right) z}}{a\left(
1+z\right) },\;a<1. \label{solution2}
\end{equation}
The first solution in the limit $a=1$ corresponds to \textit{Natural Inflation}. However to avoid the possibility of large gravitational corrections to the
potential we will concentrate on the case $f<M.$ Then it follows from Eq.$%
\left( \ref{spectral2}\right) $ that, in order to obtain a small $\delta _{%
\mathrm{ns}},$ the parameter $a$ must be small ($a<0.026$ for $n_{s}\geq 0.947$) and so the
relevant solution is the second one.
We are now able to discuss the phenomenological implications of
\textit{Hybrid Natural Inflation}. Eq.$\left( \ref{z}\right) $ implies
\begin{equation}
\delta _{\mathrm{ns}}\simeq 2a\left( \frac{M}{f}\right)^{2} c_{\mathrm{H}},
\label{fluctuations}
\end{equation}%
where now and in what follows $c_{\mathrm{H}}$ corresponds to $c_{2\mathrm{H}}$ of Eq.(\ref{solution2}).
The hybrid sector triggers the end of inflation through the growth of a term
proportional to $\sin \left( \phi /f\right) $ driving the mass squared of
the hybrid field to be negative. To avoid introducing a fine tuning between
terms in this sector it is necessary that $\sin \left( \phi /f\right) $
should be varying rapidly for $\phi \approx \phi _{e}$ corresponding to $%
\phi /f\ll \pi /2,$ $c_{\mathrm{H}}\simeq 1.$ Thus, $c.f.$ Eq.(\ref{fluctuations}), fitting the spectral index
essentially fixes the ratio $a/f^{2}$. The remaining observables are given by
\begin{eqnarray}
r &=&\frac{8a^{2}(1-c_{\mathrm{H}}^{2})}{f^{2}(1+ac_{\mathrm{H}})^{2}} \simeq
4\,a\,\delta _{\mathrm{ns}}\,(1-c_{\mathrm{H}}^{2})\, \simeq 4\,a\,\delta _{\mathrm{ns}}\,\left(
\frac{\phi _{\mathrm{H}}}{f}\right)^2 < 2\,\delta _{\mathrm{ns}}^{2}\,\left(
\frac{f}{M}\right)^2 , \label{tensor} \\
n_{r} &=&\frac{2a^{2}(1-3a^{2}-2ac_{\mathrm{H}})(1-c_{\mathrm{H}}^{2})}{%
f^{4}(1+ac_{\mathrm{H}})^{4}}\simeq \frac{1}{2}\delta _{\mathrm{ns}%
}^{2}\,(1-c_{\mathrm{H}}^{2}) \, \simeq \frac{1}{2}{\delta _{\mathrm{ns}}^2 \left(
\frac{\phi _{\mathrm{H}}}{f}\right)^2 < \frac{1}{2}\delta _{\mathrm{ns}}^{2}\, ,} \label{running} \\
A_{\mathrm{H}}^{2} &= &75\pi ^{2}\delta _{\mathrm{H}}^{2}=8\frac{V_{\mathrm{H}}}{%
M^{4}}\frac{1}{r}\equiv8\left(\frac{\Delta}{M}\right) ^{4}\frac{1}{r}\, , \label{delta4}
\end{eqnarray}
where the observed magnitude of the density perturbations corresponds to $%
A_{\mathrm{H}}\,\equiv\, \sqrt{75}\,\pi\, \delta _{\mathrm{H}}$ and $\delta _{\mathrm{H}}\simeq \left(1.91\times 10^{-5}\right).$ From this we
see that the tensor to scalar ratio is small bounded by $r<5.6\times10^{-3}$ but is
typically much smaller since $c_{\mathrm{H}}\simeq 1$ and $a$ may be much smaller
than $0.026.$ The running index is bounded as $ n_{r}<1.4\times 10^{-3}$ and the scale of inflation is also bounded, $\Delta^{4}=\frac{1}{8}M^4A_{\mathrm{H}}^{2}r$ so $\Delta <9\times10^{15}\left(
\frac{f}{M}\right)^{1/2} GeV.$ However much lower scales of inflation through the choice of $a\left( 1-c_{\mathrm{H}}^{2}\right) $ can be obtained. For
small $\phi _{\mathrm{H}}/f$ we have
\begin{equation}
\Delta =M\sqrt{\frac{A_{\mathrm{H}}\delta _{\mathrm{ns}}}{2}\left( \frac{%
\phi _{\mathrm{H}}}{M}\right) }\simeq 9\times10^{15}\sqrt{ \frac{\phi _{\mathrm{H}}}{M} } \, GeV . \label{scale}
\end{equation}
From this it is clear that to achieve low scale inflation it is necessary that $\phi_{\mathrm{H}}$ should be small. As mentioned above $\phi_{\mathrm{H}}$ is determined by the field value $\phi_e$ at the end of inflation and this is in turn determined by the hybrid sector of the theory. In non-supersymmetric hybrid models $\Delta \ge m_\chi$ in order to have zero cosmological constant after inflation and so low scale inflation requires low hybrid mass ($m_\chi$ is the mass of the hybrid field $\chi$ defined below). This in turn points to a supersymmetric model of inflation since the hybrid field, which is not a pseudo-Goldstone boson, needs supersymmetric protection against large radiative corrections to its mass (the hierarchy problem again). Thus we turn to a specific supersymmetric example of a hybrid sector in order to illustrate the expectation for the range of $\phi_e$ and hence of the scale of inflation.
\section{A supersymmetric example}
The field content of the supersymmetric model \cite{Graham:2009} consists of chiral supermultiplets of the $N=1$ supersymmetry which
transform non-trivially under a $D_{4}$ non Abelian discrete symmetry. The
supermultiplets consist of a $D_{4\text{ }}$doublet $\varphi =\left(
\begin{array}{c}
\varphi _{1} \\
\varphi _{2}%
\end{array}%
\right) $ and three singlet representations $\chi _{1,2}$ and $A$
transforming as $1^{-+}$, $1^{--}$ and $1^{++}$ respectively where the superscripts refer to the transformation properties under the $Z_2$ semidirect product factors of $D_4$ ($D_4=Z_2\ltimes Z_2'$ ). The
interactions are determined by the superpotential given by%
\begin{equation}
P=A\left( \Delta ^{2}-\lambda _{3}\frac{1}{M^{2}}\varphi _{1}\varphi
_{2}\chi _{1}\chi _{2}\right) ,
\end{equation}%
where $\Delta$ is a constant with units of mass.
These are the leading order terms consistent with an additional $R-$symmetry
under which the fields $A,\varphi ,\chi $ have charges $2,-2$ and $2$
respectively.
The field $\varphi $ has the form
\begin{equation}
\varphi =e^{i\mathbf{\phi \cdot \sigma }}\left(
\begin{array}{c}
0 \\
\rho +v%
\end{array}%
\right) =\frac{\rho +v}{\phi }\left(
\begin{array}{c}
\left( \phi _{2}+i\phi _{1}\right) \sin \left( \frac{\phi }{v}\right) \\
\phi \cos \left( \frac{\phi }{v}\right) -i\phi _{3}\sin \left( \frac{\phi }{v%
}\right)%
\end{array}%
\right) ,
\label{V}
\end{equation}%
where $\sigma _{i}$ are the Pauli spin matrices, $\phi _{i}$ are the pseudo
Goldstone fields. The field $\rho $ acquires a mass of $%
O(m_{\varphi })$ and plays no role in the inflationary era.
Inflation occurs for small $\frac{\phi _{i}}{v}.$
In this region, for positive $\lambda _{2},$ the field $\phi _{3}$ has a
positive mass squared while $\phi _{1,2}$ have negative mass squared. Thus $%
\phi _{3}$ does not develop a $vev.$ The full potential for the fields
acquiring $vev$s then has the form%
\begin{eqnarray}
V\left( \phi ,\chi \right) &\approx&\Delta ^{^{\prime }4}+64\lambda _{2}\frac{m^{2}%
}{M^{2}_M}f^{4}\cos \left( \frac{\phi }{f}\right)
+\sum\limits_{i=1}^{2}m_{\chi _{i}}^{2}\left\vert \chi _{i}\right\vert ^{2}-
\notag \\
&&-8\lambda _{3}e^{i\alpha }\Delta ^{2}\frac{f^{2}}{M^{2}}\sin \left( \frac{%
\phi }{2f}\right) \chi _{1}\chi _{2}+h.c.+O\left( \frac{m^{2}}{M^{2}}\chi
^{4}\right) , \label{susy2}
\end{eqnarray}%
where $\phi ^{2}=\phi _{1}^{2}+\phi _{2}^{2},$ $\Delta ^{^{\prime }4}=\Delta
^{4}+16m_{\varphi }^{2}f^{2}+192\lambda _{2} \frac{m^{2}}{M^{2}_M}f^{4}$
and $\alpha =\tan ^{-1}\left( \frac{\phi _{1}}{%
\phi _{2}}\right).$
The second term is a $D-$term which only arise when
supersymmetry is broken and hence is proportional to the SUSY breaking
scale, $m^{2}$. It can come from radiative corrections with a messenger
field, $M$, of mass $M_{M}$ in the loop.
A particularly simple case to analyse and one that allows for the lowest scale of inflation is the case that the hybrid fields have only soft supersymmetry breaking mass, $m_{\chi_i}=O(\Delta^2/M)$. The condition for the end of inflation is
\begin{equation}
8\lambda _{3}\Delta ^{2}\frac{f^{2}}{M^{2}}\sin \left( \frac{\phi _{e}}{2f}%
\right) \approx m_{\chi i}^{2}, \label{end}
\end{equation}
where $\lambda _{3}$ is a coefficient expected to be of $O(1)$ and $m_{\chi _i}
$ is a soft supersymmetry breaking mass of $O(\frac{\Delta ^{2}}{M})$. Thus in this case, taking $\lambda_3=O(1)$, $\phi_e$ is determined by two of the remaining inflationary parameters
\begin{equation}
\phi _{e}=O\left( \frac{\Delta ^{2}}{4f}\right) . \label{phi}
\end{equation}
Since $\phi _{\mathrm{H}}\approx \phi _{e}$ one sees that the
dependence on $\Delta $ cancels in Eq.(\ref{scale}) and the density fluctuations
are determined by $f$. To get the observed density fluctuations
\begin{equation}
4\frac{f}{M}\simeq 10^{-5}. \label{f}
\end{equation}
Using this Eq.$\left( \ref{fluctuations}\right) $ gives
\begin{equation}
a\simeq \frac{1}{2}\delta _{\mathrm{ns}}\left(\frac{f}{M}\right)^2\simeq 10^{-13} . \label{a}
\end{equation}%
For the supersymmetric model we have from Eq.(\ref{susy2})
\begin{equation}
a=64\lambda _{2}\frac{m^{2}}{M^{2}_M}\frac{f^{4}}{\Delta ^{^{\prime }4}}=64\lambda_2\left(\frac{f}{M}\right)^2 \left(\frac{f}{M_M}\right)^2
,\end{equation}
so Eq.(\ref{a}) can be satisfied by the choice of the messenger mass scale, $f/M_M\approx 10^{-2}/\sqrt{\lambda_2}$.
From Eqs.(\ref{tensor}), (\ref{running}) and Eq.(\ref{f}) we get
\begin{eqnarray}
r & < & 2\,\delta _{\mathrm{ns}}^{2}\,\left(
\frac{f}{M}\right)^2 \,\simeq \,6\times 10^{-14}, \label{tensorapp} \\
n_{r} & < & \, \frac{1}{2}\,\delta _{\mathrm{ns}}^{2} \,\simeq\, 1.4\times 10^{-3} .\label{runningapp}
\end{eqnarray}
Since $\Delta^{4}=\frac{1}{8}M^4A_{\mathrm{H}}^{2}r$ we obtain the bound for $\Delta < 1.6\times 10^{13} GeV .$
To summarise, for the case the hybrid fields have only soft supersymmetry breaking masses, fitting the observed values of the spectral index and magnitude of the density perturbations give two constraints on the three hybrid inflation
parameters. The remaining parameter, that can be taken to be the scale of
inflation, $\Delta ,$ is undetermined. Its upper bound is $10^{13}GeV$.
It has a lower bound because gravitational corrections give an irreducible mass of $O(\bar{\Delta} ^{2}/M)\ $to the $\chi $ field where $\bar{\Delta}$ is the zero temperature supersymmetry breaking scale after inflation has ended. Keeping only this contribution to $m_{\chi i}$ and solving Eq.(\ref{end}) gives
$
\phi_e=O\left(\bar{\Delta}^4/(4f\Delta^2)\right)
$
and using this in Eq.(\ref{scale}) one finds
$\Delta^4\approx 10^{-5}\left(\frac{M}{4f}\right)\bar\Delta^4$. To avoid large gravitational corrections we require $f/M<1$ which implies $\Delta>4\times 10^{-2}\bar\Delta$. In
gravity mediated schemes the need to split the superpartners from the
Standard Model states requires $\bar{\Delta }=O(10^{10}GeV)$ but in
gauge mediated schemes $\bar{\Delta}$ can be as low
as $10^{4}GeV$ implying $\Delta >400GeV$, close to the electroweak breaking scale.
\section{Initial conditions for inflation}
We have argued that hybrid natural inflation with no fine tuning of parameters requires a small value for $\frac{\phi_{\mathrm{H}}}{2f}$. Moreover one may see from Eq.(\ref{phi}) that this ratio becomes very small indeed for the scale of inflation near the lower bound just discussed. This immediately raises the question of initial conditions and how a very small initial value for $\phi$ can be achieved. To answer this question consider the thermal effects present before inflation. In general $\phi$ will have coupling to other fields with a coupling that is $D_4$ symmetric but not $SU(2)$ symmetric. For example if there is a second doublet field $\psi=\left(
\begin{array}{c}
\psi _{1} \\
\psi _{2}%
\end{array}%
\right) $
there is a $D_4$ invariant superpotential coupling $(\phi_1\psi_1-\phi_2\psi_2)\chi_1^{-+}$ that is not $SU(2)$ invariant. From Eq.(\ref{V}) one may see that this includes a term of the form $\psi_1\chi_1^{-+}\sin(\phi/f)$. Such terms can maintain the fields in thermal equilibrium and generate a term of the form $\sin(\phi/f)^2T^2$ in the effective scalar potential at temperature $T$. This in turn drives the initial value of $\phi$ to be small.
To quantify this we suppose that at temperature $T_1$ the phase transition associated with the field $\varphi$ occurs. As this occurs the $vev$ of the angular variable $\phi_{T_1}/f$ is undetermined so we expect $\phi_{T_1}/f=O(1)$. The start of inflation occurs at a temperature $T_2$ of $O(\Delta)$ and subsequently the fields fall out of thermal equilibrium. In the intervening period the $vev$ of the field is reduced $\phi_{T_2}=\phi_{T_1}e^{-m_\phi/H}\approx\phi_{T_1}e^{-M/\Delta}$ where the Hubble parameter $H=T^2/M$ and the final estimate follows using the thermal mass. One sees that such thermal effects readily set the necessary initial conditions for low scale natural hybrid inflation. One may worry that thermal fluctuations $\delta\phi=O(T)$ negates this conclusion, destabilising the hybrid field through the fourth term on the right hand side of Eq.(\ref{susy2}). However this is not the case as it is the mean field values that are relevant to the phase transitions and the thermal fluctuations do not contribute to the mean value of the $\sin(\phi/f)$ term, being odd in $\phi$.
\section{Summary and conclusions}
In summary, we have explored the phenomenological aspects of
\textit{Hybrid Natural Inflation} in which the inflaton is a pseudo Goldstone boson and hence does not suffer from the \textquotedblleft eta\textquotedblright problem. The end of inflation is driven by a hybrid sector and as a result the scale of symmetry breaking associated with the pseudo Goldstone boson inflaton can be smaller than the Planck scale. In contrast with \textit{Natural Inflation} there is no conflict of such a sub-Plankian value with the requirement of generating enough inflation or with bounds imposed by the spectral index. For the supersymmetric model discussed here the inflaton potential relevant at the time density perturbations are produced is governed by three parameters. Two of the parameters are determined by fitting the observed spectral index and the magnitude of density perturbations. The scale of inflation is set by the remaining parameter and can be anywhere in the range $400GeV<\Delta<10^{16}GeV$ and thermal effects can set the initial conditions necessary for low scale inflation provided there is a period of thermal equilibrium after the initial phase transition associated with the Goldstone boson. Thus \textit{Hybrid Natural Inflation} provides another example of inflationary models capable of generating very low scales of inflation. However in contrast with previous examples \cite{German:2001tz} there is no difficulty in fitting the observed spectral index for even the lowest scale of inflation.
\section{Acknowledgements}
The research was partially supported by the EU RTN grant UNILHC 23792. The work by GG is part of the Instituto Avanzado de Cosmolog\'{\i}a (IAC) collaboration.
|
\section{Introduction} \label{s:Introduction}
Suppose that $M$ is a complete connected noncompact
Riemannian manifold with Ricci curvature bounded from below
and positive injectivity radius.
Denote by~$-\cL$ the Laplace--Beltrami operator on $M$:
$\cL$ is a symmetric operator on $C_c^\infty(M)$ (the space of
compactly supported smooth complex-valued functions on $M$). Its closure
is a self adjoint operator on $\ld{M}$ which, with a slight
abuse of notation, we still denote by $\cL$.
We assume throughout that the bottom $b$ of the
spectrum of $\cL$ is \emph{strictly positive}.
Important examples of manifolds with these properties are nonamenable
connected unimodular Lie groups equipped with a left
invariant Riemannian distance, and symmetric spaces
of the noncompact type with the Killing metric.
It is known \cite[Section~8]{CMM1}
that for manifolds with Ricci curvature bounded
from below the assumption $b>0$ is equivalent to
an isoperimetric property,
which implies that $M$ has exponential volume growth,
\emph{ergo} {the Riemannian measure} is nondoubling.
In \cite{MMV2} we introduced a sequence $\Xu{M}, X^2(M), \ldots$
of new spaces of Hardy type on $M$ with the property that
$$
\hu{M} \supset \Xu{M} \supset X^2(M), \ldots,
$$
and the sequence $\Yu{M}, Y^2(M), \ldots$
of their dual spaces, and showed that these spaces may
be used to obtain endpoint estimates for
interesting spectral multipliers of $\cL$, including the
purely imaginary powers of $\cL$, and the first order Riesz transform.
Here $\hu{M}$ is the atomic Hardy space introduced in \cite{CMM1}.
Each of the inclusions above is proper and each of
the spaces $\Xh{M}$ is an isometric copy of $\hu{M}$.
We refer the reader to Section~\ref{s: Background material} for
the definitions of the spaces $\hu{M}$, $\Xh{M}$ and $\Yh{M}$.
Since $\Xh{M}$ is continuously included in $\hu{M}$,
each function in $\Xh{M}$ admits an atomic
decomposition in terms of $H^1$-atoms (these are defined
as classical Euclidean atoms {\cite{CW,St2}}, but their
support is contained in balls of radius at most {$1$})).
Recall that an atom $a$ in $\hu{M}$ must have integral $0$.
This cancellation condition may
also be expressed by saying that $a$ is orthogonal to the subspace
of $\ld{M}$ of functions that are constant on
the support of $a$.
E.M.~Stein posed the question whether functions in $\Xh{M}$
may be characterised as those functions in $\hu{M}$
that admit a decomposition in terms of atoms satisfying
further cancellation conditions.
The purpose of this paper is to
prove such an atomic characterisation of $\Xh{M}$ on manifolds
as above satisfying, in the case where {$k\geq 1$}, the additional
requirement that the first {$k$} covariant derivatives of the
Ricci tensor of $M$ be uniformly bounded.
Specifically,
we say that $A$ is an $X^k$-atom if $A$ is an $H^1$-atom supported in
a ball $B$ of radius at most $1$ and is orthogonal
in $\ld{B}$ to the space $\HBh$
of all functions $V$ in $\ld{M}$ such that $\cL^k V$ is constant
on a neighbourhood of $\OV{B}$.
Note that, contrary to the classical case,
the cancellation condition required for $\Xh{M}$-atoms
is expressed as orthogonality to a
\emph{infinite dimensional} subspace of $\ld{M}$.
{As far as we know, this is the first time that
an ``infinite dimensional'' cancellation condition
appears in the literature in connection with Hardy spaces.}
{An interesting and challenging problem is to prove
$\lp{M}$ bounds for the Riesz transforms for $p$ in $(1,\infty)$
and endpoint results for $p=1$. After the pioneering
works of Stein \cite{St1} and R.S.~Strichartz \cite{Str},
several contributions
have appeared recently on the subject.
We refer the reader to \cite{CD,ACDH} and
the references therein for $\lp{M}$ bounds.
Endpoint results in the case where $\mu$
is doubling and $M$ satisfies
some extra assumptions, such as
appropriate on-diagonal estimate for the heat kernel
or scaled Poincar\'e inequality~
have been obtained in \cite{CD,Ru,MRu,AMR}.
To the best of our knowledge, very little is known
about $\lp{M}$ bounds for higher order Riesz transforms.
N.~Lohou\'e \cite{Lo} proved that if $M$ is a Cartan--Hadamard
manifold such that the first $2k$ covariant derivatives of the
Riemann tensor of $M$ are uniformly bounded, and
the Laplace--Beltrami operator has spectral gap, then
the Riesz transforms of even order $\nabla^{2k}\cL^{-k}$
are bounded on $\lp{M}$ for every $p$ in $(1,\infty)$.
}
The atomic characterization of the spaces
$\Xh{M}$ enables us to prove, in a more general setting,
an endpoint result for $\nabla^{2k}\cL^{-k}$ when $p=1$,
namely that these operators are bounded from $\Xh{M}$ to $\lu{M}$
(see Theorem~\ref{t: RT}). We then obtain the
$\lp{M}$ boundedness for $p$ in $(1,2)$ by interpolation with
a classical $\ld{M}$ result of T.~Aubin~\cite{Au}.
We emphasise the fact that our proof is very short and simple.
Now we briefly outline the content of this paper.
In Section~\ref{s: Background material}, after stating the basic
geometric assumptions on the manifold $M$ and
their analytic consequences, we recall the definition
of the spaces $\Xh{M}$ and their properties. In Section~\ref{s:
Special atoms} we define the atoms in $\Xh{M}$,
we prove some of their properties and we define the atomic
space $\Xhat{M}$. In Section~\ref{s: The main result},
we prove that $\Xh{M}=\Xhat{M}$, with equivalent norms. The argument uses
a technical lemma, whose proof is rather long and
is deferred to {Section~\ref{s: proof of Lemma}}. In Section~\ref{s: Riesz}
we prove the boundedness {results for} the
Riesz transforms of even order.
We will use the ``variable constant convention'', and denote by $C,$
possibly with sub- or superscripts, a constant that may vary from place to
place and may depend on any factor quantified (implicitly or explicitly)
before its occurrence, but not on factors quantified afterwards.
If $\cT$ is a bounded linear operator from the Banach
space $A$ to the Banach space $B$, we
shall denote by $\bigopnorm{\cT}{A;B}$ its norm.
If $A=B$ we shall simply write $\bigopnorm{\cT}{A}$
instead of $\bigopnorm{\cT}{A;A}$.
\section{Basic definitions and background material}
\label{s: Background material}
Suppose that $M$ is a connected $n$-dimensional Riemannian manifold
of infinite volume with Riemannian measure $\mu$.
Denote by $\Ric$ the Ricci tensor, by $-\cL$ the Laplace--Beltrami operator
on $M$,
by $b$ the bottom of the $\ld{M}$ spectrum of $\cL$,
and set $\be =
\limsup_{r\to\infty} \bigl[\log\mu\bigl(B(o,r)\bigr)\bigr]/(2r)$.
By a result of {R.}~Brooks $b\leq \be^2$ \cite{Br}.
\begin{definition} \label{def: bounded geometry}
We say that $M$ has $C^\ell$ \emph{bounded geometry}
if the injectivity radius is positive and the following hold:
\begin{itemize}
\item{}
if $\ell =0$, then the Ricci tensor is bounded from below;
\item{}
if $\ell$ is positive, then the covariant
derivatives $\nabla^j \Ric$ of the Ricci tensor are uniformly
bounded on $M$ for all $j$ in $\{0,\ldots, \ell\}$.
\end{itemize}
\end{definition}
\begin{Basic assumptions} \label{Ba: on M}
We make the following assumptions on $M$:
\begin{enumerate}
\item[\itemno1] $b>0$;
\item[\itemno2]
$M$ has $C^\ell$ bounded geometry for some nonnegative integer $\ell$.
\end{enumerate}
\end{Basic assumptions}
We denote by~$\kappa$ the smallest positive number such that
$\Ric \geq -\kappa^2$.
\begin{remark} \label{rem: unif ball size cond}
It is well known that for manifolds with properties \rmi-\rmii\ above
there are positive constants $\al$, $\be$ and $C$ such that
\begin{equation} \label{f: volume growth}
\mu\bigl(B(p,r)\bigr)
\leq C \, r^{\al} \, \e^{2\be \, r}
\quant r \in [1,\infty) \quant p \in M,
\end{equation}
where $B(p,r)$ denotes the
geodesic ball with centre $p$ and radius~$r$.
\par
Moreover, they
satisfy the \emph{uniform ball size condition}, i.e., for every $r>0$
\begin{equation}\label{f: ubsc}
\inf\, \bigl\{ \mu\bigl(B(p, r)\bigr): p \in M \bigr\} > 0
\qquad\hbox{and}\qquad
\sup\, \bigl\{ \mu\bigl(B(p, r)\bigr): p \in M \bigr\} < \infty.
\end{equation}
See, for instance, \cite{CMP}, where complete references are given.
\end{remark}
\begin{remark}\label{ultra}
By \cite[Section~7.5]{Gr} there exists
a nonnegative number $\de$ such that the following
ultracontractive estimate holds
{
$$
\bigopnorm{\e^{-t\cL }}{1;2}
\leq C\, \e^{-bt} \, t^{-n/4} \, (1+t)^{n/4-\de/2} \quant t \in \BR^+ .
$$
Clearly this implies
$$
\bigopnorm{\e^{-t\cL }}{1;\infty}
\leq C\, \e^{-bt} \, t^{-\de} \quant t \in [1,\infty).
$$
}
\end{remark}
\medskip
We denote by $\cB$ the family of all balls on $M$.
For each $B$ in $\cB$ we denote by $c_B$ and $r_B$
the centre and the radius of $B$ respectively.
Furthermore, we denote by $c \, B$ the
ball with centre $c_B$ and radius $c \, r_B$.
For each \emph{scale parameter} $s$ in $\BR^+$,
we denote by $\cB_s$ the family of all
balls $B$ in $\cB$ such that $r_B \leq s$.
We recall the definitions of the atomic Hardy space $H^1(M)$ and its
dual space $BMO(M)$ given in \cite{CMM1}.
\begin{definition}
An $H^1$-\emph{atom} $a$
is a function in $\lu{M}$ supported in a ball $B$
with the following properties:
\begin{enumerate}
\item[\itemno1]
$\int_B a \wrt \mu = 0$;
\item[\itemno2]
$\norm{a}{2} \leq \mu (B)^{-1/2}$.
\end{enumerate}
\end{definition}
\begin{definition}
Suppose that $s$ is in $\BR^+$.
The \emph{Hardy space} $H_s^{1}({M})$ is the
space of all functions~$g$ in $\lu{M}$
that admit a decomposition of the form
\begin{equation} \label{f: decomposition}
g = \sum_{k=1}^\infty \la_k \, a_k,
\end{equation}
where $a_k$ is a $H^1$-atom \emph{supported in a ball $B$ of $\cB_s$},
and $\sum_{k=1}^\infty \mod{\la_k} < \infty$.
The norm $\norm{g}{H_s^{1}}$
of $g$ is the infimum of $\sum_{k=1}^\infty \mod{\la_k}$
over all decompositions (\ref{f: decomposition})
of $g$.
\end{definition}
The vector space $H_s^{1}(M)$
is independent of~$s$ in $\BR^+$.
Furthermore, given $s_1$ and $s_2$ in $\BR^+$,
the norms $\norm{\cdot}{H_{s_1}^{1}}$ and
$\norm{\cdot}{H_{s_2}^{1}}$ are equivalent
{\cite{CMM1}}.
\begin{notation}
\emph{We shall denote the space $H_s^{1}(M)$ simply by $\hu{M}$,
and we endow $\hu{M}$ with the norm $H_{1}^{1}(M)$.}
\end{notation}
\begin{definition}
The space $BMO(M)$ is the
space of all locally integrable functions~$f$ such that $N(f) < \infty$,
where
$$
N(f)
= \sup_{B\in \cB_{1}} \frac{1}{\mu(B)}
\int_B \mod{f-f_B} \wrt\mu,
$$
and $f_B$ denotes the average of $f$ over $B$.
We endow $BMO(M)$ with the ``norm''
$$
\norm{f}{BMO}
= N(f).
$$
\end{definition}
\begin{remark}
It is straightforward to check that $f$ is in $BMO(M)$
if and only if its sharp maximal function $f^{\sharp}$,
defined by
$$
f^{\sharp}(x)
= \sup_{B \in \cB_{1}(x)} \frac{1}{\mu(B)}
\int_B \mod{f-f_B} \wrt\mu
\quant x \in M,
$$
is in $\ly{M}$. Here $\cB_{1}(x)$ denotes the family
of all balls in $\cB_{1}$ that contain the point $x$.
\end{remark}
The Banach dual of $\hu{M}$
is isomorphic to $BMO(M)$ \cite[Thm~5.1]{CMM1}.
\par
\medskip
Now we recall the definition of the new Hardy spaces $\Xh{M}$,
introduced in \cite{MMV2}.
For every $\si$ in $\BR^+$ denote by $\cU_\si$ the operator
$\cL\, (\si I + \cL)^{-1}$.
Observe that
$$
\cU_\si
= \cI-\si \, (\si I + \cL)^{-1}.
$$
It is known that $\cU_\si$ is injective on $\lu{M}+\ld{M}$.
\begin{definition} \label{def: Hardy space}
For each positive integer $k$
we denote by $\Xh{M}$ the Banach space of all
$\lu{M}$ functions~$f$ such that $\cU_{\be^2}^{-k}f$ is in $\hu{M}$,
endowed with the norm
$$
\norm{f}{\Xk}
= \norm{\cU_{\be^2}^{-k} f}{H^1}.
$$
\end{definition}
\noindent
Clearly $\cU_{\be^2}^{-k}$ is
an isometric isomorphism between $\Xh{M}$ and $\hu{M}$.
\begin{definition} \label{def: BMO space}
For each positive integer $k$
we denote by $\Yh{M}$ the Banach dual of $\Xh{M}$.
\end{definition}
\begin{remark}
Since $\cU_{\be^2}^{-k}$ is
an isometric isomorphism between $\Xh{M}$ and $\hu{M}$,
its adjoint map~$\bigl(\cU_{\be^2}^{-k}\bigr)^*$ is an isometric isomorphism
between the dual of $\hu{M}$, i.e., $BMO(M)$, and $\Yh{M}$.
Hence
$$
\bignorm{\bigl(\cU_{\be^2}^{-k}\bigr)^*f}{\Yk}
= \norm{f}{BMO}.
$$
\end{remark}
\begin{remark} \label{rem: fund prop Xh}
We recall the following properties of the spaces $\Xh{M}$,
proved in \cite{MMV2}:
\begin{enumerate}
\item[\itemno1]
if $\si$ is in $(\be^2-b, \infty)$, then $\cU_\si\!\!^k \hu{M}$
agrees with $\Xh{M}$;
\item[\itemno2] $\cU_{\beta^2}: X^{k-1}(M)\to \Xh{M}$ is an isomorphism for every positive integer $k$;
\item[\itemno3]
$
\hu{M} \supset X^1(M) \supset X^2(M) \supset \cdots
$
with proper continuous inclusions;
{\item[\itemno4]
if $1/p = 1-\te/2$, then the complex interpolation space
$(\Xh{M},\ld{M})_{[\te]}$ is $\lp{M}$ (this is the analogue for
the spaces $\Xh{M}$ of the celebrated result of C.~Fefferman
and Stein \cite{FeS}).}
\end{enumerate}
\end{remark}
\section{Special atoms}
\label{s: Special atoms}
Atoms in $\Xh{M}$ will be $\ld{M}$ functions supported in a ball $B$
that satisfy a size condition analogous
to that for $H^1$-atoms and an infinite dimensional cancellation
condition, which will be expressed as orthogonality
to the space of ``$k$-quasi-harmonic'' functions
on $\OV{B}$ defined below.
\begin{definition}
Suppose that $k$ is a positive integer, and that
$B$ is a ball in $M$. We say that a function $V$ in $\ld{M}$
is \emph{$k$-quasi-harmonic} on $\OV{B}$ if
$\cL^k V$ is constant (in the sense of distributions)
in a neighbourhood of $\OV{B}$. We shall denote by $\HBh$ the space of
$k$-quasi-harmonic functions
on $\OV{B}$.
\end{definition}
\begin{remark} \label{rem: elliptic reg}
Observe that the following are equivalent:
\begin{enumerate}
\item[\itemno1]
$V$ is in $\HBh$;
\item[\itemno2]
$V$ is in $\ld{M}$ and is smooth in a neighbourhood of $\OV{B}$,
and $\cL^k V$ is constant therein.
\end{enumerate}
Indeed, if $V$ is in $\HBh$, then $V$ is in $\ld{M}$ by the definition
of the space $\HBh$, and $\cL^k V$ is a constant in the sense
of distributions in a neighbourhood of $\OV{B}$.
Hence $V$ is smooth on that neighbourhood by
elliptic regularity.
The converse is obvious.
\end{remark}
\noindent
Observe the following inclusions, which are direct consequences
of the definition of~$\HBh$:
$$
Q_B^1 \subset Q_B^2 \subset \cdots;
\qquad
(Q_B^1)^\perp
\supset (Q_B^2)^\perp
\supset \cdots
$$
\par\medskip
For each ball $B$ in $M$ we denote by $\ldO{B}$
the subspace of $\ld{M}$ consisting
of all $\ld{M}$ functions $f$ with support contained in the
ball $\overline{B}$, and satisfying $\int_B f \wrt \mu = 0$.
\begin{proposition} \label{p: canc II}
Suppose that $k$ is a positive integer, and that $B$ is a ball in $M$.
The following hold:
\begin{enumerate}
\item[\itemno1]
$
\HBhperp
= \set{F\in L^2(M):\cL^{-k} F\in\ldO{B}};
$
\item[\itemno2]
$\cL^{-k} \bigl( \HBhperp\bigr)$ is contained in $\ldO{B} \cap
\Dom(\cL^{k})$. Furthermore, functions in $\HBhperp$
have support contained in $\OV{B}$;
\item[\itemno3]
$\Jbemenoh \bigl( \HBhperp\bigr)$ is contained in $\ldO{B}$.
\end{enumerate}
\end{proposition}
\begin{proof}
We prove \rmi. First we show that $\HBhperp$
is contained in \break $\set{F\in L^2(M): \cL^{-k} F\in\ldO{B}}$.
Suppose that $F$ is in $\HBhperp$. To show that the
support of $\cL^{-k}F$ is contained in $\OV{B}$
it suffices to prove that $(\cL^{-k}F, \One_{B'}) = 0$
for every ball $B'$ contained in $(\OV{B})^c$.
Since $\cL$ is self adjoint,
$$
(\cL^{-k} F, \One_{B'})
= (F,\cL^{-k} \One_{B'}).
$$
Notice that $\cL^{-k} \One_{B'}$ is in $\HBh$, hence the last
inner product vanishes, as required.
Next we prove that the integral of $\cL^{-k}F$ is $0$.
Since the support of $\cL^{-k}F$ is contained in~$\OV{B}$
and $\cL$ is self adjoint,
$$
\int_M \cL^{-k}F \wrt \mu
= (\cL^{-k}F, \One_{2B})
= (F, \cL^{-k}\One_{2B}).
$$
Now, the last inner product vanishes, because $F$ is in $\HBhperp$ by assumption
and $\cL^{-k}\One_{2B}$ is in $\HBh$, as required.
Next we prove that $\set{F\in L^2(M):\cL^{-k} F\in\ldO{B}}$
is contained in $\HBhperp$.
Suppose that $\cL^{-k} F$ is in $\ldO{B}$.
Observe that $F$ is in $\Dom(\cL^k)$ and that
$F = \cL^k \cL^{-k} F$. Suppose now that $V$ is in $\HBh$.
Then $V$ is smooth in a neighbourhood of $\OV{B}$ by
Remark~\ref{rem: elliptic reg}, and
$$
\begin{aligned}
(F,V)
= (\cL^k \cL^{-k} F, V)
= (\cL^{-k} F, \cL^k V)
= 0.
\end{aligned}
$$
The last equality follows from the facts that $\cL^k V$ is constant
in a neighbourhood of $\OV{B}$,
and that $\cL^{-k}F$ is in $\ldO{B}$, so that its integral on $B$ vanishes.
Next we prove \rmii.
Clearly if $F$ is in $\ld{M}$, then $\cL^{-k}F$ is in $\Dom(\cL^k)$
by abstract set theory. Moreover $\cL^{-k}F$ is in $\ldO{B}$ by \rmi,
and the first statement of \rmii\ follows.
To prove the second statement of \rmii, observe that the support of
$\cL^{-k}F$ is contained in $\OV{B}$,
hence so is the support of $\cL^k\cL^{-k}F$, i.e., of $F$.
Finally, we prove \rmiii. Observe that $\Jbemenoh = (\cI+\be^2\, \cL)^k \,
\cL^{-k}$. Since $\cL^{-k} \bigl( \HBhperp\bigr)$
is contained in $\ldO{B} \cap \Dom(\cL^{k})$ by \rmii, it suffices to show
that $\cL^j\bigl(\ldO{B} \cap \Dom(\cL^{k})\bigr)$ is contained in
$\ldO{B}$ for all $j$ in $\{0,1,\ldots, k\}$. Suppose that $F$
is in $\ldO{B} \cap \Dom(\cL^{k})$. Denote by $\phi$ a function
in $C_c^\infty(M)$ such that $\phi=1$ on $\OV{B}$. Since $\cL$ is self adjoint
and the support of $F$ is contained in $\OV{B}$,
$$
\int_M \cL^j F \wrt \mu
= (\cL^j F, \phi)
= (F, \cL^j \phi)
= 0,
$$
as required to conclude the proof of \rmiii\ and of the proposition.
\end{proof}
\begin{definition}
Suppose that $k$ is a positive integer.
An $X^k$-\emph{atom} associated to
the ball $B$ is a function $A$ in $\ld{M}$, supported in $B$, such that
\begin{enumerate}
\item[\itemno1]
$A $ is in $\HBhperp$;
\item[\itemno2]
$\ds\norm{A}{2}\leq \mu(B)^{-1/2}$.
\end{enumerate}
Note that condition \rmi\ implies that
$\int_M A\wrt\mu=0$, because $\One_{2B}$ is in $\HBh$.
\end{definition}
\begin{remark} \label{rem: 12h atomi}
Note that if $A$ is a $X^k$-atom supported in $B$, then
$\cL^{-k}A/\opnorm{\cL^{-k}}{2}$ is
a $H^1$-atom with support contained in $\OV{B}$.
Indeed $A$ is in $\HBhperp$, so that $\cL^{-k}A$ is in $\ldO{B}$ by
Proposition~\ref{p: canc II}~\rmiii. Moreover
$$
\begin{aligned}
\norm{\cL^{-k}A}{2}
& \leq \opnorm{\cL^{-k}}{2} \, \norm{A}{2} \\
& \leq \opnorm{\cL^{-k}}{2} \, \mu(B)^{-1/2},
\end{aligned}
$$
so that $\cL^{-k}A/\opnorm{\cL^{-k}}{2}$ is a $H^1$-atom supported in $B$,
as required.
Note also that an $X^k$-atom $A$ is in $\Xh{M}$ and
\begin{equation} \label{f: norma atomica}
\norm{A}{\Xk}
\leq \opnorm{\Jbemenoh}{2}.
\end{equation}
Indeed, the function $\Jbemenoh A$ is in $\ldO{B}$
by Proposition~\ref{p: canc II}~\rmiii\ and
$$
\begin{aligned}
\norm{\Jbemenoh A}{2}
& \leq \opnorm{\Jbemenoh}{2} \, \norm{A}{2} \\
& \leq \opnorm{\Jbemenoh}{2} \, \mu(B)^{-1/2}.
\end{aligned}
$$
Therefore $\Jbemenoh A/\opnorm{\Jbemenoh}{2}$ is an $H^1$-atom, and the required
estimate follows from the definition of $\Xh{M}$.
\end{remark}
\begin{definition}
Suppose that $k$ is a positive integer. The space $\Xhat{M}$
is the space of all functions $F$ in $\hu{M}$ that admit a
decomposition of the form $F= \sum_j \la_j\, A_j$, where $\{\la_j\}$ is
a sequence in $\ell^1$ and $\{A_j\}$ is a sequence
of $X^k$-atoms supported in balls $B_j$ in $\cB_1$. Atoms supported
in balls in $\cB_1$ will be called \emph{admissible}. We endow
$\Xhat{M}$ with the norm
$$
\norm{F}{\Xkat}
= \inf\, \bigl\{\sum_{j} \mod{\la_j}: F = \sum_{j} \la_j \, A_j,\quad
\hbox{$A_j$ admissible $X^k$-atoms}\bigr\}.
$$
\end{definition}
\section{
The atomic decomposition of $\Xh{M}$.}
\label{s: The main result}
\
In this section we prove that $\Xh{M}=\Xhat{M}$ with equivalent norms.
We need two lemmata.
\begin{lemma}\label{l: Usk}
If $\sigma>\beta^2-b$ the operator $\cU_\sigma^k$ is bounded on $H^1(M)$
for every {positive} integer~$k$.
\end{lemma}
\begin{proof} Denote by $\cD$ the operator $\sqrt{\cL-b}$ and by $m_{\sigma,k}$ the function
defined by
$$
m_{\sigma,k}(\zeta)=\left(\frac{\zeta^2+b}{\zeta^2+b+\sigma}\right)^k.
$$
Then $\cU_\sigma^k=
m_{\sigma,k}(\cD)$.
The function $m_{\sigma,k}$ is bounded, even and holomorphic
in the strip $\bS_{\beta}=\set{\zeta\in\BC: \mod{\Im \zeta}<\beta}$ and there
exists a constant $C$ such that
\begin{equation}
\mod{D^j m_{\sigma,k}(\zeta)}
\leq {C} \, {(1+\mod{\zeta})^{-j}}
\quant \zeta\in \bS_{\beta} \quant j \in \BN.
\end{equation}
The conclusion follows from \cite[Thm 3.4]{MMV2}.
\end{proof}
\noindent
The main step in the proof of the atomic decomposition of $\Xh{M}$
is Lemma~\ref{l: dec resolvent} below,
which will be proved in Section~\ref{s: proof of Lemma}.
\begin{lemma} \label{l: dec resolvent}
Suppose that $k$ is a positive integer and that
$M$ has $C^k$ bounded geometry. If
$A$ is an admissible $X^{k-1}$-atom then $\Jbekappa A$ is
in $\Xhat{M}$, and
there exists a constant $C$, independent of $A$, such that
$$
\norm{\Jbekappa A}{\Xkat}
\leq C ,
$$
{where $\kappa$ is the constant which appears
in the lower bound of the Ricci tensor.}
\end{lemma}
\noindent
The main result of this section is the following.
\begin{theorem} \label{t: atomic dec}
Suppose that $k$ is a positive integer and that
$M$ has $C^k$ bounded geometry (see Definition~\ref{def: bounded
geometry}).
Then $\Xh{M}$ and $\Xhat{M}$
agree as vector spaces and there exists a constant $C$ such that
\begin{equation} \label{f: atomic dec}
C \, \norm{F}{\Xkat}
\leq \norm{F}{\Xk}
\leq \bigopnorm{\Jbe^{-k}}{2} \, \norm{F}{\Xkat}
\quant F \in \Xh{M}.
\end{equation}
\end{theorem}
\begin{proof}
First we prove that $\Xhat{M}$ is contained in $\Xh{M}$, and that
the right-hand inequality in (\ref{f: atomic dec}) holds.
Suppose that $F= \sum_{j} \la_j \, A_j$, where $A_j$ is an
admissible $X^k$-atom. By Remark~\ref{rem: 12h atomi}
(see (\ref{f: norma atomica})), the function
$\Jbemenoh A_j/\opnorm{\Jbemenoh}{2}$
is an $H^1$-atom.
The series $\sum_j \la_j \, \Jbemenoh A_j$ is then convergent in
$\hu{M}$. Denote by $f$ its sum. Since $\Jbeh$ is bounded
on $\hu{M}$ by Lemma \ref{l: Usk},
$$
\Jbeh f
= \sum_j \la_j \, \Jbeh\bigl(\Jbemenoh A_j\bigr)
= F.
$$
Thus, $F$ is in $\Xh{M}$, and
$$
\norm{F}{\Xk}
= \norm{f}{H^1}
\leq \, \bigopnorm{\Jbe^{-k}}{2}
\, \sum_j \mod{\la_j}.
$$
The right-hand inequality in (\ref{f: atomic dec}) follows
from this by taking the infimum over all the decompositions
of $F$ of the form $F = \sum_j \la_j \, A_j$.
Next we prove that $\Xh{M}$ is contained in $\Xhat{M}$ and that
the left-hand inequality in (\ref{f: atomic dec}) holds.
For notational convenience, in the rest of this proof
we denote $\hu{M}$ also by $X^0(M)$, and write $\cR$ instead of $\Rbekappa$
and $\cU$ instead of $\Jbekappa$.
We argue inductively.
The result is trivial in the case where $k=0$, because $\cU^0= \cI$.
Suppose that the result holds for $k-1$ and that
$F$ is in $\Xh{M}$. Then $f=\cU^{-1} F$
is a function in $\Xhmu{M}$, and
$\norm{f}{X^{k-1}} = \norm{F}{\Xk}$, by Remark \ref{rem: fund prop Xh}.
By the induction hypothesis for each $\vep$ in $\BR^+$
there exist a sequence $\{A_j\}$ of
admissible $X^{k-1}$-atoms and a summable
sequence~$\{c_j\}$ of complex numbers such that
\begin{equation}
f = \sum_{j} c_j \, A_j
\qquad\hbox{and}\qquad
\norm{f}{X^{k-1}}
\geq \sum_j \mod{c_j} - \vep.
\end{equation}
Observe that we may write
\begin{equation} \label{f: series for F}
F
=\cU f
= \sum_j c_j \, \ \cU A_j ,
\end{equation}
because
the series $\sum_j c_j \, A_j$ converges to $f$ in $\hu{M}$,
and $\cU$ is bounded on~$\hu{M}$ by Lemma \ref{l: Usk}.
Fr{}om (\ref{f: series for F}) and Lemma~\ref{l: dec resolvent},
we see that
$$
\begin{aligned}
\norm{F}{\Xkat}
& \leq \sum_j \mod{c_j} \, \norm{\cU A_j}{\Xkat} \\
& \leq C \,
\sum_j \mod{c_j} \\
& \leq C \,
\bigl( \norm{f}{X^{k-1}} + \vep\bigr) \\
& = C \,
\bigl( \norm{F}{\Xk} + \vep\bigr).
\end{aligned}
$$
Therefore $F$ is in $\Xhat{M}$, and $\norm{F}{\Xkat}
\leq C \, \norm{F}{\Xk}$, as required.
This concludes the proof of the theorem.
\end{proof}
\begin{remark}
Suppose that $k$ is a positive integer
and that $s$ is a scale parameter in $\BR^+$.
The space of all functions $F$ in $\hu{M}$ that admit a
decomposition of the form $F= \sum_j \la_j\, A_j$, where $\{\la_j\}$ is
a sequence in $\ell^1$ and $\{A_j\}$ is a sequence
of $X^k$-atoms supported in balls $B_j$ in $\cB_s$ agrees with
$\Xhat{M}$ (hence with $\Xh{M}$).
The norm on $\Xhat{M}$ defined by
$$
\inf\, \bigl\{\sum_{j} \mod{\la_j}: F = \sum_{j} \la_j \, A_j,\quad
\hbox{$A_j$ are $X^k$-atoms supported in balls
of $\cB_s$}\bigr\}
$$
is an equivalent norm on $\Xhat{M}$.
To prove this, it suffices to observe that
minor modifications in the proof of Theorem~\ref{t: atomic dec}
and Lemma~\ref{l: dec resolvent} show that $F$ is in $\Xh{M}$
if and only if it admits a decomposition in terms
of $X^k$-atoms supported in balls in $\cB_s$.
\end{remark}
\begin{remark}
Suppose that $p$ is in $(1,\infty)$
and denote by $p'$ the index conjugate to $p$.
Assume that $k$ is a positive integer and that $B$ is in $\cB$.
Define ${Q}_{B,p'}^k$ to be the space of all functions $V$
in $L^{p'}(M)$ such that $\cL^k V$ is constant (in the sense
of distributions) in a neighbourhood of $\OV{B}$.
Then denote by $({Q}_{B,p'}^k)^\perp$ the annihilator of
${Q}_{B,p'}^k$ in $\lp{M}$. Then a $X^k$-atom in $\lp{M}$ is
an element $A$ of $({Q}_{B,p'}^k)^\perp$, satisfying the size condition
$$
\norm{A}{p}
\leq \mu(B)^{-1/p'}.
$$
It is straightforward to modify
the theory of this section to show
that $\Xh{M}$ admits an atomic characterisation
in terms of $X^k$-atoms in $\lp{M}$. The fact that $\cU$
is an isomorphism of $\lp{M}$ for all $p$ in $(1,\infty)$
plays an important r\^ole here.
\end{remark}
As a consequence of the atomic decomposition
of the space $\Xh{M}$, we may describe explicitly the
action of elements of $Y^k(M)$, the dual of $\Xh{M}$, on
finite linear combinations of $X^k$-atoms.
\begin{definition}
Suppose that $k$ is a positive integer. We denote by
$\Xhfin{M}$ the vector space of all finite linear
combinations of $X^k$-atoms and by $\hufin{M}$
the vector space of all finite linear
combinations of $H^1$-atoms.
\end{definition}
Suppose that $\ell$ is a continuous linear functional on $\Xh{M}$.
Since $\Jbeh$ is an isomorphism between $\hu{M}$ and $\Xh{M}$
and $\Xh{M}$ and $\Xhat{M}$ are isomorphic by Theorem~\ref{t: atomic dec},
$\ell\circ \Jbeh$ is a continuous linear
functional on $\hu{M}$. By \cite[Thm~5.1]{CMM1}, there exists a function
$f$ in $BMO(M)$ such that
$$
(\ell\circ\Jbeh) (g)
= \int_M g \, f \wrt \mu
\quant g \in \hufin{M}.
$$
Clearly
$$
\norm{\ell}{Y^k}
= \norm{\ell\circ\Jbeh}{(H^1)^*}
= \norm{f}{BMO}.
$$
It may be worth describing how the functional $\ell$
acts on $X^k$-atoms, or, more generally, on functions
in $\Xhfin{M}$.
Suppose that $A$ is a $X^k$-atom with support contained in an
arbitrary ball $B$.
Since $\Jbe^{-k} = \bigl(\cI+\be^2\cL^{-1}\bigr)^k$,
there exist constants $c_j$ such that
$$
\Jbe^{-k}A
= \sum_{j=0}^k c_j \, \cL^{-j}A.
$$
Then $\Jbe^{-k}A$ is a finite linear combination of
$H^1$-atoms by Remark~\ref{rem: 12h atomi}.
Therefore
$$
\begin{aligned}
\ell(A)
& = (\ell\circ\Jbeh) (\Jbe^{-k} A) \\
& = \int_M (\Jbe^{-k}A) \, f \wrt \mu.
\end{aligned}
$$
Observe that $\Jbe^{-k}A$ is supported in $B$, so that the last
integral is just the inner product in $\ld{M}$
between $\Jbe^{-k}A$ and $\One_B\, f$. Since $\Jbe^{-k}$
is self adjoint, we may write
$$
\ell(A)
= \int_M A \, \, \Jbe^{-k}(\One_B f) \wrt \mu.
$$
A similar argument shows that
if $F$ is in $\Xhfin{M}$ and its support is contained
in the ball $B$, then
$$
\ell(F)
= \int_M F \, \, \Jbe^{-k}(\One_B f) \wrt \mu.
$$
It may be worth observing that a consequence of
this representation formula for $\ell$, and of
the fact that $\norm{\ell}{Y^k} = \norm{f}{BMO}$,
is that
$$
N'(f) \leq \norm{\ell}{Y^k} \leq \opnorm{\Jbeh}{2} \, N'(f),
$$
where
$$
N'(f)
= \sup_{B\in \cB_1} \Bigl(\frac{1}{\mu(B)} \int_B
\bigmod{\Jbe^{-k} (\One_B\, f)
- f_B \,\Jbe^{-k} \One_B }^2 \wrt\mu \Bigr)^{1/2}
$$
and $f_B$ denotes the average of $f$ over $B$.
The proof of this fact is straightforward and is omitted.
\section{Riesz transforms of even order}\label{s: Riesz}
Denote by $\nabla$ the covariant derivative on $M$. The Riesz
transform of order $\ell$ is the operator $\nabla^{\ell}\cL^{-\ell/2}$
mapping smooth functions
with compact support on $M$ to sections of the bundle $T_\ell(M)$ of
covariant tensors of order $\ell$. In this section we exploit the atomic decomposition
of the spaces $X^k(M)$ to prove that the Riesz
transforms of even order $\nabla^{2k}\cL^{-k}$ extend to bounded operators
from $X^k(M)$ to the space $L^1\big(T_{2k}(M)\big)$ of $L^1$ sections of $T_{2k}(M)$.
To prove this result we need to strengthen the bounded
geometry assumption on $M$, by replacing the derivatives of the Ricci
tensor with those
of the Riemann tensor in Definition \ref{def: bounded geometry}.
\begin{definition}\label{sbg}
We say that $M$ has $C^\ell$ \emph{strongly bounded} geometry
if the injectivity radius is positive and the following hold:
\begin{itemize}
\item{}
if $\ell =0$, then the Ricci tensor is bounded from below
$\Ric \geq -\kappa^2$ for some positive $\kappa$;
\item{}
if $\ell$ is positive, then the covariant
derivatives $\nabla^j R$ of the Riemann tensor are uniformly
bounded on $M$ for all $j$ in $\{0,\ldots, \ell\}$.
\end{itemize}
\end{definition}
\noindent
We recall that the boundedness of the first order Riesz transform on $L^2(M)$
follows from the identity $\cL=\nabla^*\nabla$ and the self-adjointness
of $\cL$ on $L^2(M)$ \cite{Str}.
From a result of Aubin \cite[Prop. 3]{Au}, it follows
also that if $b>0$ and $M$ has $C^{\ell-2}$ strongly bounded
geometry then the Riesz
transform
of order $\ell\ge2$ extends to a bounded operator from $L^2(M)$ to
the space $L^2\big(T_\ell(M)\big)$ of square integrable sections of $T_\ell
(M)$.
In \cite{MMV2} the authors,
{under the additional assumption that $b=\be^2$}, proved that
the Riesz transform of order $1$
maps {$\Xh{M}$} to $L^1\big(T_1(M)\big)$ {for $k$ large enough}.
{In general}, the Riesz transforms of
{order one} do not map $H^1(M)$ to $L^1\big(T_\ell(M)\big)$.
A counterexample on noncompact symmetric spaces will appear in a
forthcoming paper of the authors \cite{MMV3}.
{Notice that the modified Riesz transform of order $1$, i.e.
the operator $\nabla (\cL+\vep I)^{-1/2}$, {for $\vep >0$}, maps $\hu{M}$ into $\lu{T_1(M)}$
even when $M$ satisfies less stringent assumptions on $M$ \cite{Ru}}.
\begin{theorem}\label{t: RT}
If $b>0$ and $M$ has $C^{2k-2}$ strongly bounded geometry then
the Riesz transform of order $2k$
extends to a bounded operator from $X^k(M)$ to {$\lu{T_{2k}(M)}$ and from
$\lp{M}$ to $\lp{T_{2k}(M)}$ for all $p$ in $(1,2)$}.
\end{theorem}
\begin{proof}
{To prove that $\nabla^{2k}\cL^{-k}$ is bounded
from $X^k(M)$ to $\lu{T_{2k}(M)}$}
it suffices to show that $\nabla^{2k}\cL^{-k}$ maps $X^k$-atoms
into $L^1\big(T_{2k}(M)\big)$ uniformly. Suppose that $A$ is a
$X^k$-atom associated to the ball $B$. Then, by Remark \ref{rem: 12h atomi},
the function $\cL^{-k} A$ is supported in $\overline{B}$.
Hence, by H\"older's inequality and the boundedness of
$\nabla^{2k}\cL^{-k}$ on $L^2(M)$,
\begin{align*}
\norm{\nabla^{2k}\cL^{-k} A}{1}&\le \norm{\nabla^{2k}\cL^{-k} A}{2}
\, \mu(B)^{1/2}\\
&\le C\, \norm{ A}{2} \
\mu(B)^{1/2}\le \,C,
\end{align*}
{as required. The boundedness of $\nabla^{2k}\cL^{-k}$ from
$\lp{M}$ to $\lp{T_{2k}(M)}$ for all $p$ in $(1,2)$
follows by interpolation
(see \cite[Thm~2.15]{MMV2})
from {the $X^k(M)--\lu{T_{2k}(M)}$ boundedness} and the aforementioned
result of Aubin.}
\end{proof}
\section{Proof of Lemma \ref{l: dec resolvent} }
\label{s: proof of Lemma}
In this section we shall prove Lemma \ref{l: dec resolvent}.
First we need a variant of the ``economical decomposition of atoms" proved in \cite[Lemma 5.7]{MMV2}.
\begin{lemma} \label{l: economical decomposition I}
Suppose that $k$ is a positive integer and that
$M$ has $C^k$ bounded geometry (see Definition~\ref{def: bounded geometry}).
If $a$ is an $H^1$-atom in $\Dom(\cL^k)$,
then $\cL^k a$ is in $\Xhat{M}$.
Furthermore, if the support of $a$ is contained in the ball $B$,
then there exists a constant $C$ such that
$$
\norm{\cL^k a}{\Xkat}
\leq C\, (1+ \,r_B)\, \mu(B)^{1/2}\, \norm{\cL^k a}{2}.
$$
\end{lemma}
\begin{proof}
Suppose first that the support of $a$ is contained in
a ball $B$ such that $r_B\le 1$.
Since $\cL^ka$ is in $\HBhperp$ by Proposition \ref{p: canc II},
$\mu (B)^{-1/2}{\cL^k a}/\norm{\cL^k a}{2}$ is a $X^k$-atom supported
in a ball in $\cB_1$ and the lemma is proved. \par
Next, suppose that $r_B>1$.
Denote by $\fS$ a $1/3$-discretisation
of $M$, i.e. a set of points in $M$ that is
maximal with respect to the property
$$
\min\{d(z,w): z,w \in \fS, z \neq w \} >1/3,
\quad\hbox{and}\quad d(\fS, x) \leq 1/3 \quant x \in M.
$$
The family $\set{B(z,1):z\in \fS}$ is a covering of $M$ which is
uniformly locally finite, by the uniform ball size and the {local
doubling} property of the Riemannian measure
(see, for instance, \cite[Theorem~3.10]{Ch}).
By the same token, the set $B\cap\fS$ is
finite and has at most $N$ points $z_1,\ldots,z_N$, with $N \le
C\,\mu(B)$, where $C$ is a constant which does not depend on $B$.
Denote by $B_j$ the ball with centre $z_j$ and radius $1$,
and by $\set{\psi_j: j=1,\ldots,N}$ a partition of unity on $B$
subordinated to the covering $\set{B_j:j=1,\ldots,N}$.
Fix~$j$ in $\{1,\ldots,N\}$ and denote by $z_j^0,\ldots, z_j^{N_j}$
points on a minimizing geodesic joining $z_j$ and $c_B$,
with the property that
$z_j^0 = z_j$, $z_j^{N_j} = c_B$, and $d(z_j^h,z_j^{h+1})$ is
approximately equal to $1/3$. Note that $N_j\le 4r_B$.
Denote by $B^h_j$ the ball $B(z^h_j,1/12)$,
for $j=1,\ldots,N$ and $h=0,\ldots, N_j$. Then the balls
$B^h_j$ are disjoint, $B^h_j\subset B(z_j^h,1)\cap B(z_j^{h+1},1)$
and $B_j^{N_j}=B(c_B,1/12)$. \par
Denote by $\phi^h_j$ a nonnegative function in $C^\infty_c(B^h_j)$ that
has integral $1$. By the uniform ball size property we may choose the
functions $\phi^h_j$ so that there exists a constant $A$
such that $\norm{\phi^h_j}{2}\le A$ for all $h$ and $j$.
The existence of a uniform bound on the derivatives of the
Ricci tensor implies that
we can choose the functions $\psi_j$ and $\phi^h_j$ so that their covariant
derivatives of order up to $2k$ are uniformly bounded for all $j$ and $h$
(see \cite[p. 14]{He}).
Now, denote by $a_j^0$ the function $a\, \psi_j$. Clearly
$$
a
= \sum_{j=1}^N \psi_j \, a
= \sum_{j=1}^N \, a_j^0.
$$
Next, define
$$
a_j^1
= a_j^0-\phi_j^0 \, \int_M a_j^0\wrt\mu
\quad\hbox{and}\quad
a_j^h=(\phi_j^{h-2}-\phi_j^{h-1})\int_M a_j^0\wrt\mu,
\quad 2\le h\le N_j+1.
$$
Then, for every $h$ in $\set{1,\ldots,N_j}$,
the support of $a_j^h$ is contained in $B(z_j^{h-1},1)$, the integral
of $a_j^h$ vanishes and
$$
\begin{aligned}
\norm{a_j^h}{2}
& \le 2A \int_M \mod{a_j^0}\wrt\mu \\
& \le C\, \norm{a_j^0}{2} \, \mu(B_j)^{1/2}\\
& \le C\, \norm{a_j^0}{2} \, \mu(B_j^h)^{-1/2}.
\end{aligned}
$$
In the last two inequalities we have used
the uniform ball size property (\ref{f: ubsc}).
Hence there exists a constant $C$, independent of $j$ and $h$,
such that
\begin{equation} \label{f: normaHuno ajk}
\norm{a_j^h}{H^1}
\leq C \, \norm{a_j^0}{2}.
\end{equation}
Moreover
$$
a_j^0=\sum_{h=1}^{N_j+1}a_j^h+\phi_j^{N_j}\int_M a_j^0\wrt\mu.
$$
Thus
$$
a
= \sum_{j=1}^N\sum_{h=1}^{N_j+1} a^h_j,
$$
because $\sum_j \int_M a^0_j\wrt\mu=\int_M a\wrt\mu=0$ and all the functions
$\phi_j^{N_j}$, $j=1,\ldots,N_j$ coincide, since $B^{N_j}_j = B(c_B,1/12)$.
Moreover, $a^1_j$ is in $\Dom(\cL^k)$, and
$$
\begin{aligned}
\norm{\cL^k a^1_j}{2}
& \le \norm{\cL^k a^0_j}{2}+\norm{\cL^k \phi^0_j}{2}
\int_M \mod{a^0_j} \wrt\mu \\
& \le C\,\norm{\cL^k a^0_j}{2},
\end{aligned}
$$
where, in the last inequality, we have used the estimate
$\norm{a^0_j}{2}
\le C\, \norm{\cL^k a^0_j}{2}$, which holds because $\cL$ has spectral gap.
Similarly, if $h=2,\ldots,N_j+1$, then
$a^h_j$ is in $\Dom(\cL^k)$, and
$$
\norm{\cL^k a^h_j}{2}\le C\, \norm{\cL^k a^0_j}{2}.
$$
Hence $\cL^k a^h_j/\norm{\cL^k a^0_j}{2}$ is a multiple of
a $X^k$-atom supported in a ball of radius $1$, with a constant $C$
which does not depend on $j$ and $h$ by the uniform ball size property.
Thus
\begin{equation}\label{ajk}
\norm{\cL^ka^h_j}{\Xkat}
\le C\, \norm{\cL^ka^0_j}{2}
\quant j,h.
\end{equation}
Adding up the inequalities in (\ref{ajk}), we obtain
\begin{align*}
\norm{\cL^k a}{\Xkat}
&\le \sum_{j=1}^N \sum_{h=1}^{N_j+1} \norm{\cL^k a^h_j}{\Xkat} \\
&\le C\,\sum_{j=1}^N\sum_{h=1}^{N_j+1} \norm{\cL^k a^0_j}{2}.
\end{align*}
Remembering that $N_j\le C\,r_B$ and $N\le C\,\mu(B)$, and using Schwarz's
inequality, we see that the right-hand side is dominated by
\begin{align*}
C\, r_B \sum_{j=1}^N \norm{\cL^k a^0_j}{2}
& \le C\, r_B\,N^{1/2} \, \Bigl(\sum_{j=1}^N
\norm{\cL^k a^0_j}{2}^2\Bigr)^{1/2}\\
& \le C\, r_B\,\mu(B)^{1/2} \, \norm{\cL^k a}{2}.
\end{align*}
In the last inequality we have used the fact that $\set{\psi_j}$
is a partition of unity on $B$, subordinated to the
uniformly locally finite covering $\set{B_j}$.\par
This completes the proof of the lemma.
\end{proof}
\begin{remark}\label{c: large atoms}
There exists a constant $C$ such that
$$
\norm{f}{\Xkat}
\le C\, (1+r_B)\,\mu(B)^{1/2} \, \norm{f}{2}
\quant f\in \HBhperp.
$$
Indeed,
if $f$ is in $\HBhperp$, then the function $\cL^{-k}f$ is a
multiple of a $H^1$-atom, by Proposition~\ref{p: canc II}.
The conclusion follows, by Lemma~\ref{l: economical decomposition I}.
\end{remark}
\noindent
The second ingredient in our proof of
Lemma~\ref{l: dec resolvent} are two technical results in
one-dimensional Fourier analysis (see Lemma~\ref{l: P} and
Lemma~\ref{l: stima Runo} below). To state them we need some more notation.
For every $f$ in $\lu{\BR}$
define its Fourier transform $\wh f$~by
$$
\wh f(t) = \ir f(s) \, \e^{-ist} \wrt s
\quant t \in \BR.
$$
If $f$ is a function on $\BR$, and $\la$ is in $\BR^+$,
we denote by $f^\la$ and $f_\la$ the~$\la$-dilates of $f$, defined by
\begin{equation} \label{f: dilate}
f^\la(x)
= f(\la x)
\qquad\hbox{and}\qquad
f_\la(x)
= \la^{-1} \, f(x/\la) \quant x \in \BR.
\end{equation}
\noindent
For each $\nu\geq -1/2$, denote by $\cJ_\nu: \BR\setminus\{0\} \to \BC$
the modified Bessel function of order $\nu$, defined by
$$
\cJ_\nu(t) = \frac{J_\nu(t)}{t^\nu},
$$
where $J_\nu$ denotes the standard Bessel function of the first
kind and order $\nu$ (see, for instance,
\cite[formula~(5.10.2), p.~114]{L} for the definition). Recall that
$$
\cJ_{-1/2} (t)
= \sqrt{\frac{2}{\pi}}\ \cos t
\qquad\hbox{and that}\qquad
\cJ_{1/2} (t)
= \sqrt{\frac{2}{\pi}}\ \frac{\sin t}{t}.
$$
For each positive integer $\ell$,
we denote by $\cO^\ell$ the differential operator
$t^\ell\, D^\ell$ on the real~line. For the proof
of the following lemma, see \cite[Lemma~4.1]{MMV2}.
\begin{lemma}\label{l: P}
For every positive integer $k$ there exists a polynomial
$P_{k+1}$ of degree $k+1$ without constant term, such that
\begin{equation} \label{f: propertiesRiesz I}
\ir {f} (t) \, \cos (vt) \wrt t
= \ir P_{k+1}(\cO)f(t)\, \cJ_{k+1/2} (t v) \wrt t,
\end{equation}
for all functions $f$ such that
$\cO^\ell f\in \lu{\BR}\cap C_0(\BR)$ for all
$\ell$ in $\{0,1,\ldots, k+1\}$.
\end{lemma}
Denote by $\om$ an even function in
$C_c^\infty(\BR)$ which is supported in $[-3/4,3/4]$, is equal to~1
in $[-1/4,1/4]$, and satisfies
$$
\sum_{j\in \BZ} \om(t-j) = 1
\quant t \in \BR.
$$
Denote by $\phi$ the function $\om^{1/4}- \om$,
where $\om^{1/4}$ denotes the $1/4$-dilate of $\om$.
Then $\phi$ is smooth, even and vanishes in the complement of the set
$\{t \in \BR: 1/4\leq \mod{t} \leq 4\}$.
For a fixed $R$ in $(0,1]$ and for each positive integer $i$,
denote by $E_{i}$ the set
$\{t \in \BR: 4^{i-1}R \leq \mod{t} \leq 4^{i+1}R\}$.
Clearly $\phi^{1/(4^{i}R)}$ is supported in $E_{i}$,
and $\sum_{i=1}^\infty \phi^{1/(4^{i}R)} =~1$ in $\BR\setminus (-R,R)$.
Denote by $d$ the integer $[\!\![\log_4(3/R)]\!\!]+1$.
To avoid cumbersome notation, we write $\rho_i$ instead of $1/(4^{i}R)$.
Then
\begin{equation} \label{f: dec om phi}
\om^{\rho_0} + \sum_{i=1}^{d} \phi^{\rho_i} =~1
\qquad\hbox{on}\quad [-3,3].
\end{equation}
Suppose that $c$ is in $\BR^+$, and
denote by $r$ the function defined by
\begin{equation}\label{funzr}
r(\la)
= \frac{1}{c^2+\la^2}
\quant \la \in \BC \setminus \{\pm ic\}.
\end{equation}
Note the decomposition
\begin{equation} \label{f: dec om hat r}
\wh{\om}* r(\lambda)= \sum_{i=0}^{d} S_i (\la),
\end{equation}
where the functions $S_i: \BR\to\BC$ are defined by
\begin{equation} \label{f: A0}
S_0(\la)
= \frac{1}{2\pi} \ir \om^{\rho_0}(t) \,\, P_{N}(\cO)(\om\, \wh r)(t)
\,\, \cJ_{N-1/2}(\la\, t) \wrt t
\quant \la \in \BR,
\end{equation}
and, for $i$ in $\{1,\ldots,d\}$,
\begin{equation} \label{f: Ai}
S_i(\la)
= \frac{1}{2\pi} \, \ir \phi^{\rho_i}(t)\,
P_{N}(\cO) (\om\, \wh r)(t) \, \cJ_{N-1/2}(\la t) \wrt t
\quant \la \in \BR,
\end{equation}
{where $N$ is a positive integer.}
\begin{remark} \label{rem: sigmak}
Note that there exist constants $c_\ell$ such that
$$
t^{-1}P_{N}(\cO)
= \sum_{\ell=0}^{N-1} c_\ell \, t^\ell \, D^{\ell+1}.
$$
\end{remark}
\begin{lemma} \label{l: stima Runo}
Suppose that $N$ is a positive integer.
The following hold:
\begin{enumerate}
\item[\itemno1]
the norm
$\norm{t^{-1}P_{N}(\cO)
\wh{r}}{\infty}$ is finite;
\item[\itemno2]
if $N\geq 3$, then there exists a constant $C$,
independent of $R$ in $(0,1]$, such that
$$
\sup_{\la\ge 0}\, (\la^2+1)\, \mod{S_0(\la)}
\leq C .
$$
\end{enumerate}
\end{lemma}
\begin{proof}
By Remark ~\ref{rem: sigmak},
to prove \rmi\ it suffices to show that
\begin{equation}\label{tDr}
\sup_{t \in \BR} \mod{t^\ell \, D^{\ell+1} \wh{r}(t)}
< \infty
\quant \ell \in \{0,\ldots, N-1\}.
\end{equation}
This is a standard estimate in Fourier analysis.
Recall that $\wh{r}(t) = (1/c)\, \e^{-c\mod{t}}$.
It is straightforward to check that $D\wh r = -c\, \wh r\,\, \sgn$,
and that for each $k\geq 1$
$$
\begin{aligned}
D^{2k} \wh{r}
=&\ c^{2k} \, \wh r - 2 \, \sum_{j=0}^{k-1} c^{2(k-1-j)} \, D^{2j} \de_0
\\
D^{2k+1} \wh{r}
=&\, - c^{2k+1} \, \wh r \cdot \sgn - 2 \,
\sum_{j=0}^{k-1} c^{2(k-1-j)}\, D^{2j+1} \de_0.
\end{aligned}
$$
Hence
$$
t^{2k-1}\, D^{2k} \wh{r}(t)
= c^{2k}\, t^{2k-1}\, \wh r(t)
\quad\hbox{and}
\quad
t^{2k}\, D^{2k+1} \wh{r} (t)
= - c^{2k+1}\, t^{2k} \, \sgn(t)\, \cdot \wh r(t),
$$
so that
$$
\mod{t^\ell\, D^{\ell+1} \wh{r}(t)}
= c^{\ell} \, \mod{t}^\ell\, \e^{-c\mod{t}}
\quant t \in \BR,
$$
and the required estimate follows.
To prove \rmii, observe that, on the one hand, by (\ref{f: A0}) and (\ref{tDr})
$$
\begin{aligned}
\mod{S_0(\la)}
& \leq \norm{\om}{\infty} \, \norm{
t^{-1}P_{N}(\cO)
(\om\, \wh r)}{\infty}
\ir \mod{t}\, \mod{\cJ_{N-1/2} (t \la)} \wrt t \\
& \leq \, C\,\norm{\om}{\infty} \,
\,\, \la^{-2}
\quant \la \in [0,\infty).
\end{aligned}
$$
On the other hand, the function $\cJ_{N-1/2}$ is bounded, so that
$$
\begin{aligned}
\mod{S_0(\la)}
& \leq C\, \norm{t^{-1}P_{N}(\cO)
(\om\, \wh r)}{\infty}
\ir \om^{\rho_0}(t) \,\mod{t} \wrt t \\
& \leq C
\quant \la \in [0,\infty).
\end{aligned}
$$
We have used the fact that $\rho_0 = 1/R$ and $R\leq 1$ in the
last inequality.
The proof of the lemma is complete.
\end{proof}
The third, and last, ingredient in the proof of Lemma \ref{l: dec resolvent} is the following proposition, which shows that certain functions of the operator $\cL$ map $H^1$-atoms
into functions that have integral $0$.
For technical reasons, it is convenient to work with functions of the wave propagator
$$
\cD_1=\sqrt{\cL-b+\kappa^2}
$$
instead of functions of $\cL$.
We recall that $-\kappa^2$ is the greatest lower
bound of the Ricci curvature (see Basic assumptions~\ref{Ba: on M}).
The reason for considering the operator $\cD_1$ instead of
$\cD=\sqrt{\cL-b}$ is that, in order to prove estimates of
the gradient of the kernels associated to functions of $\cL$,
we need to exploit the identity $\wrt\cL=\mathbb{L} \wrt$,
where $\mathbb{L}$ is the Hodge Laplacian $\mathbb{L}$ on $1$-forms
and $\wrt$ denotes exterior differentiation (see \cite[Prop. 5.5]{MMV1}).
Whereas, in general, the operator $\mathbb{L}-b$ is not a
positive operator on $1$-forms, the operator
$\mathbb{L}-b+\kappa^2$ is nonnegative on manifolds whose
Ricci curvature satisfies the lower bound $\Ric\ge -\kappa^2$ .
\begin{proposition} \label{p: Mean}
Suppose that $\nu$ is in $[-1/2,\infty)$, that $w$ is
a complex measure on $\BR$
and that $a$ is an $H^1$-atom. Define the operator $\cW_\nu(\cD_1)$
on $\ld{M}$ spectrally by
$$
\cW_{\nu}(\cD_1) f
= \ir
\, \cJ_{\nu} (t\cD_1) f
\wrt w(t)
\quant f \in \ld{M}.
$$
The following hold:
\begin{enumerate}
\item[\itemno1]
$\int_M \cW_\nu(\cD_1) a \wrt \mu = 0$;
\item[\itemno2]
$\int_M S_i(\cD_1) a \wrt \mu = 0$ for $i=0,\ldots,d$
(the functions $S_i$ are defined in (\ref{f: A0}) and (\ref{f: Ai}).
\end{enumerate}
\end{proposition}
\begin{proof}
A simple argument, based on the finite speed of propagation property
of the operator $\cL-b+\kappa^2$, shows that
\begin{equation}\label{}
\int_M\cJ_\nu(t\cD_1)\,a\wrt\mu =0.
\end{equation}
(see \cite[Prop. 5.5]{MMV2}). Since $\cW_\nu$
and the function $S_i$ are integrals of
$\cJ_\nu(t\cdot)$ with respect to complex measures,
we obtain the desired conclusion by interchanging the order of integration.
\end{proof}
\begin{remark}\label{r: fps}
Note that for every $\nu$ in $[-1/2,\infty)$
the function $\la\mapsto\cJ_\nu(t\la)$ is even and of
entire of exponential type $t$, so that
kernel $k_{\cJ_\nu (t\cD_1)}$ of the operator
$\cJ_\nu(t\cD_1)$
is supported in
the set $\{(x,y) \in M\times M:
d(x,y) \leq t\}$ by the finite propagation speed.
\end{remark}
\noindent
The main step in the proof of our main result
is Lemma~\ref{l: dec resolvent}, which we restate for the
reader's convenience. {The idea, used in the proof, of subordinating
spectral functions of $\cL$ to the wave propagator
has been used several times
since its appearance in \cite{CGT,Ta}}.
\begin{lemma*}[{\bf 4.2}]
Suppose that $k$ is a positive integer and that
$M$ has $C^k$ bounded geometry.
Let
$A$ be an admissible $X^{k-1}$-atom. Then $\Jbekappa A$ is
in $\Xhat{M}$, and
there exists a constant $C$, independent of $A$, such that
$$
\norm{\Jbekappa A}{\Xkat}
\leq C.
$$
\end{lemma*}
\begin{proof}
Suppose that the atom $A$ is supported in the ball
$B(p,R)$. Then $R\leq 1$, because $A$ is admissible.
Denote by $N$ an integer $>n/2+3$.
For notational convenience, in this proof
we shall write $\cJ$ instead of $\cJ_{N-1/2}$,
$\cR$ instead of $\Rbekappa$, $\cU$ instead of $\Jbekappa$
and $c$ instead of $\sqrt{4\be^2+b }$.
Observe that $\cR = r(\cD_1)$ (the function $r$ was defined in (\ref{funzr})).
\par
\emph{Step I: splitting of the operator}.
Define the operators~$\cS$ and $\cT$ spectrally~by
\begin{equation} \label{f: decomposition R}
\cS
= (\wh\om\ast r) (\cD_1)
\qquad\hbox{and}\qquad
\cT
= (r-\wh\om\ast r) (\cD_1).
\end{equation}
Then
$\cU A=\cL
\cR A
=\cL \cS A+\cL \cT A.
$
We shall prove that both $\cL\cS A$ and $\cL\cT A$
are in $\Xhat{M}$ and that there exists a constant $C$, independent
of $A$, such that
\begin{equation} \label{f: required atomic}
\norm{\cL\cS A}{\Xkat}
\leq C
\qquad\hbox{and}\qquad
\norm{\cL\cT A}{\Xkat}
\leq C.
\end{equation}
The proof of estimates (\ref{f: required atomic})
will be given in Steps~II and III.
\emph{Step II: proof of the first inequality
in (\ref{f: required atomic})}.
Note that
$\om\,\wh r$ has support in $[-3/4,3/4]$.
Define the functions $S_i$ as in (\ref{f: A0}) and
(\ref{f: Ai}).
Observe that, by
(\ref{f: dec om hat r}),
\begin{equation}\label{f:S=sumSi}
\cS
= \sum_{i=0}^{d} S_i(\cD_1),
\end{equation}
where $d=[\!\![\log_{4}(3/R)+1]\!\!]$.
Denote by $B_i$ the ball with centre~$p$ and radius $(4^{i+1}+1)R$.
Since the support of the kernel of the operator $S_i(\cD_1)$
is contained in $\{(x,y): d(x,y) \leq 4^{i+1}R\}$ by the finite
propagation speed, the function $S_i(\cD_1)A$ is
supported in $B_i$.
Now we check that $\cL S_i(\cD_1)A$ is in $\bigl(Q_{B_i}^k\bigr)^{\perp}$.
By Proposition~\ref{p: canc II} it suffices to show
that $\cL^{-k}\cL S_i(\cD_1)A$ is in $L^2_0(B_i)$.
Now, $\cL^{-k}\cL S_i(\cD_1)A=S_i(\cD_1)\cL^{1-k}A$ and $\cL^{1-k}A$
is a constant multiple of
a $H^1$-atom with support contained in $B(p,R)$
by Remark~\ref{rem: 12h atomi}. Thus the support of $S_i(\cD_1)\cL^{1-k} A$
is contained in $B_i$ and its integral
over $M$ vanishes by Proposition~\ref{p: Mean}~\rmii.
Next, we claim that there exists a constant $C$, independent
of $A$, such that for $i$ in $\{0,\ldots,d\}$
\begin{equation} \label{f: at size est I}
\norm{S_i(\cD_1)A}{2}
\leq C \,
\, \mu(B_i)^{-1/2} \,4^{-i}
\end{equation}
and
\begin{equation} \label{f: at size est II}
\norm{\cL S_i(\cD_1)A}{2}
\leq C\,
\, \mu(B_i)^{-1/2}\,4^{-i}.
\end{equation}
Deferring momentarily the proof of the claim, we show that the
first inequality in (\ref{f: required atomic}) follows from it.
Indeed, by (\ref{f:S=sumSi}) and the triangle inequality,
$$
\norm{\cL\cS A}{\Xkat}
\leq C\, \sum_{i=0}^{d} \norm{\cL S_i(\cD_1) A}{\Xkat}.
$$
Now Remark~\ref{c: large atoms} and (\ref{f: at size est II})
imply that
$$
\begin{aligned}
\norm{\cL S_i(\cD_1) A}{\Xkat}
& \leq C \, \mu(B_i)^{1/2}\, \norm{\cL S_i(\cD_1) A}{2} \\
& \leq C \, 4^{-i}.
\end{aligned}
$$
Hence
$$
\norm{\cL\cS A}{\Xkat}
\leq C,
$$
as required to prove the first inequality in (\ref{f: required atomic}).
To conclude the proof of Step~II it remains to prove
(\ref{f: at size est I}) and (\ref{f: at size est II}).
The function $S_0$ is bounded by Lemma~\ref{l: stima Runo}
hence $S_0(\cD_1)$ is bounded on $\ld{M}$ by the spectral theorem, and
$$
\opnorm{S_0(\cD_1)}{2}\
\leq \norm{S_0}{\infty}.
$$
Since $S_{0}(\cD_1)A$ is supported in $B_0=B(p,5R)$, we have
$$
\norm{S_{0}(\cD_1)A}{2}
\leq
\opnorm{S_{0}(\cD_1)}{2}\, \norm{A}{2}
\leq C\, R^{-n/2}.
$$
Furthermore, the integral of $S_0(\cD_1)A$ vanishes
by Proposition~\ref{p: Mean}~\rmii,
so that $S_0(\cD_1)\,A$ is a constant multiple of an $H^1$-atom.
Denote by $k_{S_i(\cD_1)}$ the integral kernel of the
operator $S_i(\cD_1)$. Observe that
$$
S_i(\cD_1)\, A(x)
= \int_{B(p,R)} A(y) \, \bigl[k_{S_i(\cD_1)}(x,y)
- k_{S_i(\cD_1)}(x,p)\bigr] \wrt \mu(y).
$$
By Minkowski's integral inequality
and the fact that the support of $S_i(\cD_1)\, A$
is contained in $B_i$, we have that
$$
\begin{aligned}
\norm{S_i(\cD_1)\, A}{2}
& = \norm{S_i(\cD_1)\, A}{\ld{B_i}} \\
& \leq \int_{B(p,R)} \mod{A(y)} \, I_{i}(y) \wrt \mu(y),
\end{aligned}
$$
where
$$
I_{i}(y)
= \norm{k_{S_i(\cD_1)}(\cdot,y)
- k_{S_i(\cD_1)}(\cdot,p)}{L^2(B_i)}
\quant y \in B(p,R).
$$
To estimate $I_{i}(y)$, we observe that
$$
I_{i}(y)
\leq d(y,p) \, \sup_{z\in M} \, \bignorm{\dest_2
k_{S_i(\cD_1)}(\cdot,z)}{\ld{B_{i}}},
$$
and, by
(\ref{f: Ai}) and (\ref{f: decomposition R}),
$$
\dest_2
k_{S_i(\cD_1)}(\cdot,z)=\frac{1}{2\pi} \, \ir \phi^{\rho_i}(t)\,
P_{N}(\cO)(\om\wh{r})
(t) \ \dest_2
k_{\cJ(t\cD_1)}(\cdot,z) \wrt t.
$$
Recall that $\phi^{\rho_i}$ is supported in $E_i
= \{ t\in \BR: 4^{i-1} R \leq \mod{t} \leq 4^{i+1}R \}$,
that the support of $\om\wh{r}
$
is contained in $[-1,1]$ and that $d(p,y)<R$.
Then, by \cite[Prop.~2.2~\rmiii]{MMV1} (with $\cJ$
in place of $F$), there exists a constant $C$, independent of $i$
and $R$, such that
$$
\begin{aligned}
I_{i}(y)
& \leq C \, d(y,p) \,
\ir \phi^{\rho_i}(t)\, \mod{P_{N}\cO)
(\om\wh{r})(t) }
\, \,
\sup_{z\in M} \bignorm{\dest_2
k_{\cJ (t\cD_1)}(\cdot,z)}{\ld{B_{i}}} \wrt t \\
& \leq C \, \norm{t P_{N}(\cO)
(\om\wh{r}) }
{\infty} \,R \,
\int_{E_{i}} \mod{t}^{-n/2-2} \wrt t \\
& \leq C \,
\, R \, (4^iR)^{-n/2-1} \,.
\end{aligned}
$$
Thus,
$$
\begin{aligned}
\norm{S_i(\cD_1)\, A}{2}
& \leq C\,
\, 4^{-i}\,
(4^i R)^{-n/2} \, \norm{A}{1} \\
& \leq C\,
\,
4^{-i}\, \mu(B_i)^{-1/2}.
\end{aligned}
$$
This concludes the proof of (\ref{f: at size est I}).
Now we prove (\ref{f: at size est II}). Recall that
$
\cL = \cD^2 + b\, \cI = \cD_1^2 + (b-\kappa^2) \, \cI.
$
Therefore
\begin{equation} \label{f: stima LRunoa}
\norm{\cL S_i(\cD_1) A}{2}
\leq \norm{\cD_1^2 S_i(\cD_1) A}{2} + \mod{b-\kappa^2}
\, \norm{S_i(\cD_1) A}{2}.
\end{equation}
We first estimate $\norm{\cD_1^2 S_i(\cD_1) A}{2}$
when $i$ is in $\{1,\ldots,d\}$.
Observe that
$$
\cD_1^2\, S_i(\cD_1)
=\frac{2^{N-1}}{\sqrt{2\pi}} \ir \frac{\phi^{\rho_i}(t)}{t^2}\,
P_N(\cO) (\om \, \wh r)(t)\, F (t \cD_1) \wrt t,
$$
where $F(\lambda)=\lambda^2 \cJ(\lambda)$.
Since the function $\la\mapsto F(t\la)$ is an even entire function of
exponential type $\mod{t}$ and the support of
$\phi^{\rho_i}$ is contained in the set {$E_i$},
the support of $F(t\cD_1)\, A$ is contained in $B_i$,
by finite propagation speed. Thus
$$
F(t\cD_1)\, A(x)
= \int_{B(p,R)} A(y) \, \bigl[k_{F(t\cD_1)}(x,y)
- k_{F(t\cD_1)}(x,p)\bigr] \wrt \mu(y),
$$
and, by Minkowski's integral inequality,
$$
\begin{aligned}
\norm{F(t\cD_1)\, A}{2}
& = \norm{F(t\cD_1)\, A}{\ld{B_i}} \\
& \leq \int_{B(p,R)} \mod{A(y)} \, I_{i}(y) \wrt \mu(y) \\
& \leq \norm{A}{1} \,\sup_{y\in B(p,R)} I_{i}(y),
\end{aligned}
$$
where
$$
I_{i}(y)
=\norm{k_{F(t\cD_1)}(\cdot,y)
- k_{F(t\cD_1)}(\cdot,p)}{\ld{B_i}}\quant y \in B(p,R).
$$
Observe that
$$
I_{i}(y)
\leq d(y,p) \,\, \sup_{z\in M} \, \bignorm{\dest_2
k_{F(t\cD_1)}(\cdot,z)}{\ld{B_i}}.
$$
Since
$
\sup_{\la \in \BR^+} (1+\la)^{N-2}\, \mod{F(\la)} < \infty
$
by the asymptotics of Bessel functions of the first kind and
$N-2 > n/2+1$ by assumption,
we may use \cite[Prop.~2.2~\rmiii]{MMV1},
and conclude that
$$
\sup_{z\in M} \, \norm{ \dest_2 k_{F(t\cD_1)}(\cdot,z)}{\ld{B_i}}
\leq C \, \mod{t}^{-n/2-1}
\quant t \in [-1,1].
$$
Therefore, since the support of $\phi^{\rho_i}$
is contained in the set {$E_i$},
$$
\begin{aligned}
\norm{\cD_1^2 S_i(\cD_1) A}{2}
& \leq C \, R \, \ir \frac{\phi^{\rho_i}(t)}{t^2}\,
\mod{P_N(\cO) (\om \, \wh r)(t)}\,
\sup_{z\in M} \bignorm{\dest_2
k_{F(t\cD_1)}(\cdot,z)}{\ld{B_i}} \wrt t \\
& \leq C \,\norm{
t^{-1}P_N(\cO)
(\om \, \wh r)}{\infty} \,
\frac{R}{(4^{i}R)^{n/2+3}} \ir \phi^{\rho_i}(t) \, \mod{t} \wrt t \\
& \leq C \,
4^{-i} \, \mu(B_i)^{-1/2}
\quant i \in \{1,\ldots,d\}.
\end{aligned}
$$
Now, the inequality (\ref{f: at size est II}) for $i\in \{1,...,d\}$ follows directly
fr{}om this, (\ref{f: stima LRunoa}) and (\ref{f: at size est I}).
Next we consider $\cL S_0(\cD_1)A$.
Observe that $\cL S_0(\cD_1)A$ is supported in $B(p,5R)$, and that
$$
\begin{aligned}
\norm{\cL \,S_0(\cD_1)A}{2}
& \leq \bigopnorm{\cL \,S_0(\cD_1)}{2} \, \norm{A}{2} \\
& \leq \mu\bigl(B(p,R)\bigr)^{-1/2} \,
\bigopnorm{\cL \,S_0(\cD_1)}{2} \\
& \leq C\, \mu\bigl(B(p,5R)\bigr)^{-1/2} \,
\bigopnorm{\cL \,S_0(\cD_1)}{2}.
\end{aligned}
$$
To prove that $\cL\, S_0(\cD_1)$
is bounded on $\ld{M}$, with norm independent
of $R$ in $(0,1]$ observe that, by the spectral theorem and
Lemma~\ref{l: stima Runo}
$$
\begin{aligned}
\bigopnorm{\cL\, S_0(\cD_1)}{2}
& \leq \sup_{\la \geq 0} \, (\la^2+b) \, \mod{S_0(\la)} \\
& \leq C,
\end{aligned}
$$
where $C$ is independent of $R$.
This concludes the proof of
~(\ref{f: at size est II}), and of Step~II.
\emph{Step III: proof of the second inequality in~(\ref{f: required atomic})}.
For each $j$ in $\{1,2,3,\ldots\}$, define~$\om_j$ by the formula
\begin{equation}\label{omj}
\om_j(t) = \om (t-j) + \om(t+j) \quant t \in \BR.
\end{equation}
Observe that $\sum_{j=1}^\infty \om_j=1-\om$
and that the support of $\om_j$ is contained in the set of all
$t$ in $\BR$ such that $j-3/4\le\mod{t}\le j+3/4$.
In the rest of this proof, we write $\Om_{j,N}$ instead of
$P_N(\cO)(\om_j\, \wh r)$.
Observe that the support of $\Om_{j,N}$
is contained in $\set{t\in\BR:j-3/4\le\mod{t}\le j+3/4}$.
Moreover, since $\wh{r}(t)={c}^{-1}\, \e^{-c\mod{t}}$ and
$c>2\beta$ there exist constants $C,\varepsilon>0$ such that
\begin{equation}\label{propOM}
\norm{\Om_{j,N}}{\infty}
\le C\, \e^{-2\beta j}
\quant j\in \{1,2,\ldots\}.
\end{equation}
Define the function $T_j: \BR\to \BC$ by
\begin{equation} \label{f:ridef}
T_j(\la)
= \ir \Om_{j,N}(t)\, \cJ (t \la)\wrt t \quant \la \in \BR.
\end{equation}
We may use the observation that $(m-\wh\om\ast m)\wh{\phantom a}
= \sum_{j=1}^\infty \om_j\, \wh m$
and formula~(\ref{f: propertiesRiesz I}), and write
$$
\begin{aligned}
(m-\wh\om\ast m)(\la)
& = \frac{1}{2\pi} \ir \bigl(1-\om(t)\bigr) \, \wh{r}(t)
\, \cos (t\la) \wrt t \\
& = \sum_{j=1}^\infty T_j(\la).
\end{aligned}
$$
Then, by the spectral theorem,
$$
\cT A
=\sum_{j=1}^\infty T_j(\cD_1) A.
$$
Now we estimate $\norm{T_j(\cD_1) A}{2}$.
By the asymptotics of $J_{N-1/2}$ \cite[formula (5.11.6), p.~122]{L}
$$
\sup_{s>0} \mod{(1+s)^{N} \, \cJ (s)} < \infty.
$$
By Remark \ref{ultra} we may apply
\cite[Proposition~2.2~\rmi]{MMV1}, since $N-1/2>(n+1)/2$,
and conclude that
$$
\begin{aligned}
\norm{\cJ (t\cD_1) A}{2}
& \leq \norm{A}{1} \, \bigopnorm{\cJ (t\cD_1)}{1;2} \\
& \leq \sup_{y\in M} \bignorm{k_{\cJ (t\cD_1)}(\cdot,y)}{2} \\
& \leq C\, \mod{t}^{-n/2}\, \bigl(1+\mod{t}\bigr)^{n/2-\de}
\quant t \in \BR\setminus \{0\}.
\end{aligned}
$$
for some $\de>0$.
The function
$\cJ (t\cD_1)A$ is supported in $B(p,t+R)$
by Remark \ref{r: fps},
and has integral $0$ by Proposition~\ref{p: Mean}~\rmi.
Moreover
\begin{align*}
\norm{T_j(\cD_1)A}{2}
& \leq C\, \ir \mod{\Om_{j,N}(t)}
\, \norm{\cJ (t \cD_1)A}{2} \wrt t \nonumber \\
& \leq C\, \int_{j-3/4}^{j+3/4} \mod{{\Om_{j,N}(t)}} \,
\mod{t}^{-n/2}\, \bigl(1+\mod{t}\bigr)^{n/2-\de} \wrt t \\
& \leq C \,
\e^{-2\beta \, j} \quant j \in \{1,2,\ldots\}. \\
\end{align*}
By (\ref{f: volume growth}) there exist $\vep>0$ such that
$\e^{-2\be \, j}\leq C\, \mu\bigl(B(p,j+1)\bigr)^{-1/2} \, \e^{-\vep \, j}$.
Hence
\begin{equation} \label{f: est aj2}
\norm{T_j(\cD_1) A}{2}
\leq C \,
\mu\bigl(B(p,j+1)\bigr)^{-1/2} \, \e^{-\vep \, j}
\quant j \in \{1,2,\ldots\}.
\end{equation}
Observe that, at least formally,
$$
\cL \cT A
= \sum_{j=1}^\infty \, \cL \, T_j(\cD_1) A.
$$
To prove that the series converges in $\Xhat{M}$ we estimate $\norm{\cL T_j(\cD_1) A}{2}$. Note that
\begin{equation} \label{f: stima Laj}
\norm{\cL T_j(\cD_1) A}{2}
\leq \norm{\cD_1^2 T_j(\cD_1) A}{2} + \mod{b-\kappa^2}
\, \norm{T_j(\cD_1) A}{2}.
\end{equation}
We have already estimated $\norm{T_j(\cD_1) A}{2}$ in (\ref{f: est aj2}),
so we concentrate on $\norm{\cD_1^2 T_j(\cD_1) A}{2}$.
By (\ref{f:ridef}) and the spectral theorem
$$
\cD_1^2 T_j(\cD_1)
= \ir \Omega_{j,N}(t)\,F(t\cD_1) \frac{\wrt t}{t^2},
$$
where $F(\la)=\la^2 \cJ(\la)$.
By using (\ref{propOM}), \cite[Proposition~2.2~\rmii]{MMV1}
and the fact that the support of $\Om_{j,N}$
is contained in $\set{t: j-3/4\le \mod{t}\le j+3/4}$, we obtain that
there exist constants $C$ and $\vep>0$ such~that
$$
\begin{aligned}
\norm{\cD_1^2T_j(\cD_1) A}{2}
& \leq C\, \ir \mod{\Om_{j,N}(t)}
\, \norm{F (t \cD_1)A}{2} \frac{\wrt t}{t^2} \\
& \leq C\, \norm{A_{}}{1}\, \ir \mod{\Om_{j,N}(t)} \,
\bigopnorm{F (t \cD_1)}{1;2} \frac{\wrt t}{t^2} \\
& \leq C \,
\, \e^{-2\be \, j} \\
& \leq C \,
\,
\mu\bigl(B(p,j+1)\bigr)^{-1/2} \, \e^{-\vep\, j}
\quant j \in {\{1,2,\ldots\}}.
\end{aligned}
$$
This estimate, (\ref{f: est aj2}) and (\ref{f: stima Laj}) then imply
that
\begin{equation} \label{f: Tja}
\norm{\cL \, T_j(\cD_1) A}{2}
\leq C\,
\, \mu\bigl(B(p, j+1)\bigr)^{-1/2}\,\e^{-\vep j}.
\end{equation}
Now, by (\ref{f:ridef}) we may write
\begin{equation}
\begin{aligned}
\cL T_j(\cD_1)A
& = \cL \ir \Om_{j,N}(t)\, \cJ (t \cD_1)A \wrt t \\
& = \cL^{k} \ir \Om_{j,N}(t)\, \cJ (t \cD_1)\cL^{1-k}A \wrt t \\
&= \cL^{k} a_j,
\end{aligned}
\end{equation}
where $a_j = \ir \Om_{j,N}(t)\, \cJ (t \cD_1)\cL^{1-k}A \wrt t$.
\par
The function $a_j$ is supported in $B(p,j+1)$, since $\cL^{1-k}A$
is in $L^2_0\big(B(p,R)\big)$ and the kernel of the operator
$\int_{-\infty}^\infty \Om_{j,N}(t)\, \cJ (t \cD_1) \wrt t$
is supported in $\set{(x,y): d(x,y)\le j}$. Moreover, $\int_M a_j \wrt\mu=0$ by Proposition~\ref{p: Mean}~\rmii, and
\begin{equation} \label{f: est int TjA}
\norm{a_j}{2}
\leq C\, \opnorm{\cL^{-k}}{2}
\, \mu\bigl(B(p, j+1)\bigr)^{-1/2}\,\e^{-\vep j},
\end{equation}
by (\ref{f: Tja}). Hence $a_j$ is a multiple of an $H^1$-atom supported in $B(p,j+1)$.Then we may apply Lemma~\ref{l: economical decomposition I} to the function
$a_j$, and conclude
that $\cL T_j(\cD_1)A=\cL^k a_j$ is in $\Xhat{M}$, and that,
by (\ref{f: Tja}),
$$
\begin{aligned}
\norm{\cL T_j(\cD_1)A}{\Xkat}
&\leq C\, j\,\big(\mu(B(p,j+1)\big)^{1/2} \, \norm{\cL T_j(\cD_1)A}{2}\\
&\leq C\,
\, j\, \e^{-\vep j}.
\end{aligned}
$$
By summing over $j$, we see that
$$
\norm{\cL\cT A}{\Xkat}
\leq C\,
\sum_{j=1}^\infty \, j\, \e^{-\vep j},
$$
thereby concluding the proof of Step~III and of the lemma.
\end{proof}
|
\section{Introduction}
As one of the closest and most massive Galactic globular clusters
(GCs), \objectname[47 Tuc]{47 Tucanae} (47 Tuc) is a test-bed for
Galaxy formation models \cite[][and references therein]{Salaris07},
distance measurement techniques \cite[e.g.][]{Bono08} and metallicity
calibrations \cite[e.g.][]{McWilliam08,Lane09b}. \new{This close
examination, however, has left several unresolved conundrums. While
not unique in this respect \cite[see][for a review of elemental
abundance variations in GCs]{Gratton04}, 47 Tuc has a bimodal
distribution of carbon and nitrogen line strengths
\cite[e.g.][]{Harbeck03}. Furthermore, 47 Tuc has a complex stellar
population, exhibiting, for example, multiple sub-giant branches
\cite[e.g.][]{Anderson09}. Although multiple Red Giant and Horizontal
Branches can also be found in other GCs, and may be due to chemical
anomalies \cite[e.g.][]{Ferraro09,Lee09}, 47 Tuc is particularly
unusual in many respects. It has an extreme rotational velocity
\cite[a property it shares only with M22 and $\omega$ Centauri,
e.g.][]{Merritt97,Anderson03,Lane09b} and an apparently unique rise in
its velocity dispersion profile at large radii \cite[][]{Lane09b}.}
Furthermore, the mass-to-light ratio (M/L$_{\rm V}$) of 47 Tuc is very
low for its mass \cite[][]{Lane10}, that is, it does not obey the
mass-M/L$_{\rm V}$ relation described by \cite{Rejkuba07}. Note that
this mass-M/L$_{\rm V}$ relation is not due to the presence of dark
matter but because of dynamical effects
\cite[][]{Kruijssen08}. \new{Explanations for these unusual properties
may be intimately linked to its evolutionary history. In this Letter
we describe and analyse various explanations for the rise in velocity
dispersion in the outer regions of 47 Tuc initially reported by
\cite{Lane09b}.}
\cite{Lane09b} provided a complete description of the data acquisition
and reduction, radial velocity uncertainty estimates, the membership
selection process, and statistical analysis of cluster membership for
all data presented in this Letter.
\section{Plummer Model Fits}\label{Plummerfits}
The \cite{Plummer11} model \cite[see also][]{Dejonghe87} predicts that
the isotropic, projected velocity dispersion $\sigma$ falls off with
radius $r$ as:
\begin{equation}
\sigma(r; \left\{\sigma_0, r_0\right\}) = \frac{\sigma_0}{\left(1 + \left(\frac{r}{r_0}\right)\right)^{1/4}},
\end{equation}
where $\sigma_0$ is the central velocity dispersion and $r_0$ is the
scale radius of the cluster. We now describe how we fitted Plummer
profiles to the 47 Tuc radial velocity data to infer the values of the
parameters $\sigma_0$ and $r_0$, and also to evaluate the overall
appropriateness of the Plummer hypothesis ($H_1$) by comparing it to a
more complex double Plummer model ($H_2$).
The mechanism for carrying out this comparison is Bayesian model
selection \citep{sivia}. Suppose we have two (or more) competing
hypotheses, $H_1$ and $H_2$, with each possibly containing different
parameters $\theta_1$ and $\theta_2$. We wish to judge the
plausibility of these two hypotheses in the light of some data $D$,
and some prior information. Bayes' rule provides the means to update
our plausibilities of these two models, to take into account the data
$D$:
\begin{eqnarray}\label{bayes}
\nonumber\frac{P(H_2|D)}{P(H_1|D)} =
\frac{P(H_2)}{P(H_1)}\frac{P(D|H_2)}{P(D|H_1)} \\ =
\frac{P(H_2)}{P(H_1)}\times\frac{\int_{\theta_1}
p(\theta_1|H_1)p(D|\theta_1, H_1)\,d\theta_1}{\int_{\theta_2}
p(\theta_2|H_2)p(D|\theta_2, H_2)\,d\theta_2}.
\end{eqnarray}
Thus, the ratio of the posterior probabilities for the two models
depends on the ratio of the prior probabilities and the ratio of the
{\it evidence} values. The latter measure how well the models predict
the observed data, not just at the best-fit values of the parameters,
but averaged over all plausible values of the parameters. It should be
noted that we rely solely on the velocity information for our Plummer
model fits. Taking the stellar density as a function of radius into
account would be useful in further constraining the models. However,
the Plummer model fits by Lane et al. (in prep.) based exclusively on
velocity information are also good fits to the surface brightness
profiles of the four GCs analysed in that study.
\subsection{Single Plummer Model}
The data are a vector of radial velocity measurements $\mathbf{v} =
\{v_1, v_2, ..., v_N\}$ of $N$ stars, along with the corresponding
distances from the cluster centre $\mathbf{r} = \{r_1, r_2, ...,
r_N\}$ and observational uncertainties on the velocities
$\mathbf{\sigma_{\textnormal{obs}}} = \{\sigma_{\textnormal{obs}, 1},
..., \sigma_{\textnormal{obs},N}\}$. We will consider $\mathbf{v}$ to
be the data, whereas $\mathbf{r}$ and
$\mathbf{\sigma_{\textnormal{obs}}}$ are considered part of the prior
information. In this case the probability distribution for the data
given the parameters is the product of independent Gaussians, whose
standard deviations vary with radius:
\begin{align}
\nonumber p(\mathbf{v} | \mu, \sigma_0, &r_0)=\\
\prod_{i=1}^N &\left(\frac{1}{\sqrt{2\pi\sigma_i^2}}
\exp\left(-\frac{1}{2}\left(\frac{v_i -
\mu}{\sigma_i}\right)^2\right)\right),
\end{align}
where $\mu$ is the systemic velocity of the cluster. The standard
deviation $\sigma_i$ for each data point is given by a combination of
the standard deviation predicted by the Plummer model, and the
observational uncertainty:
\begin{equation}
\sigma_i = \sqrt{\sigma(r_i; \left\{\sigma_0, r_0\right\})^2 +
\sigma_{\textnormal{obs}, i}^2}
\end{equation}
To carry out Bayesian Inference, prior distributions for the
parameters must also be defined. We assigned a uniform prior for $\mu$
(between $-30$ and 30 km s$^{-1}$). For $\sigma_0$, we assigned
Jeffreys' scale-invariant prior $p(\sigma_0) \propto 1/\sigma_0$ for
$\sigma_0$ in the range $0.1-100\textnormal{ km s}^{-1}$. Finally,
$r_0$ was assigned the Jeffreys' prior $p(r_0) \propto 1/r_0$ for
$r_0$ in the range $0.2-220$~pc. These three prior distributions were
all chosen to be independent and to cover the approximate range of
values that we expect the parameters to take.
\subsection{Double Plummer Model}\label{doublePlummer}
The double Plummer model is a simple extension of the Plummer
model. The stars are hypothesised to come from two distinct
populations, each having its own Plummer profile parameters (but with
a common systemic velocity $\mu$). Thus, at any radius $r$ from the
cluster centre, we model the velocity distribution as a mixture
(weighted sum) of two Gaussians. From previous work \cite[][]{Lane09b}
we also expect the inner regions of the cluster to be well fitted by a
single Plummer profile, so the weight of the second population of
stars should become more significant at larger radii.
Thus, instead of having a single $\sigma_0$ parameter and a single
$r_0$ parameter, there are now two of each. The probability
distribution for the data given the parameters is then the weighted
sum of two Gaussians:
\begin{align}
\nonumber p(\mathbf{v} | \mu&, \{\sigma_0\}, \{r_0\}, w(r)) =\\
&\prod_{i=1}^N \left(\frac{w(r_i)}{\sqrt{2\pi\sigma_{i, 1}^2}} \exp\left(-\frac{1}{2}\left(\frac{v_i - \mu}{\sigma_{i, 1}}\right)^2\right)\right. \nonumber\\
&+ \left.\frac{1-w(r_i)}{\sqrt{2\pi\sigma_{i, 2}^2}} \exp\left(-\frac{1}{2}\left(\frac{v_i - \mu}{\sigma_{i, 2}}\right)^2\right)\right).
\end{align}
Here, $w(r)$ is a weight function that determines the relative
strength of one Plummer profile with respect to the other, as a
function of radius. We expect one component to dominate at smaller
radii, and to eventually fade away as the second component becomes
dominant. Hence, we parameterise the function $w(r)$ as:
\begin{eqnarray}
w(r) = \frac{\exp\left(u(r)\right)}{1 + \exp\left(u(r)\right)}
\end{eqnarray}
where
\begin{eqnarray}
u(r) = u_{\alpha} + \frac{r - r_{\textnormal{min}}}{r_{\textnormal{max}} - r_{\textnormal{min}}}\left(u_{\beta} - u_{\alpha}\right).
\end{eqnarray}
That is, the log of the relative weight between one Plummer component
and the other increases linearly over the range of radii spanned by
the data, starting at $u_{\alpha}$ and ending at a value
$u_{\beta}$. Parameterising $w$ via $u$ makes it easier to enforce the
condition that $w(r)$ must be in the range $0-1$ for all $r$. The
prior distributions for $u_{\alpha}$ and $u_{\beta}$ were chosen to be
Gaussian with mean zero and standard deviation 3. This implies that
$w(r)$ will probably lie between 0.05 and 0.95, with a small but not
negligible chance of extending lower than 0.001 or above 0.999.
The standard deviations of the two Gaussians at each data point are
given by a combination of that predicted by the Plummer model and the
observational uncertainty:
\begin{eqnarray}
\sigma_{i, 1} &=& \sqrt{\sigma(r_i; \{\sigma_0, r_0\}_1)^2 + \sigma_{\textnormal{obs}, i}^2} \\
\sigma_{i, 2} &=& \sqrt{\sigma(r_i; \{\sigma_0, r_0\}_2)^2 + \sigma_{\textnormal{obs}, i}^2}
\end{eqnarray}
The priors for all the parameters $\mu$, $\{\sigma_0, r_0\}$ were
chosen to be the same as in the single Plummer case.
\section{Results}
An obvious rise in the velocity dispersion of 47 Tuc was described by
\citet[][their Figure 11]{Lane09b} at approximately half the tidal
radius ($\sim28$\,pc). The tidal radius is $\sim56$\,pc
\cite[][]{Harris96}. To confirm the reality of this rise, several
tests were performed, including resizing the bins and shifting the bin
centres, as described by \cite{Lane09b}. No difference in the overall
shape of the dispersion profile was found during any tests.
We used a variant \new{\cite[][]{Brewer09}} of Nested Sampling
\citep{skilling} to sample the posterior distributions for the
parameters, and to calculate the evidence values for the single and
double Plummer models. The results are listed in
Table~\ref{results_table}. The double Plummer model is favoured by a
factor $\sim 3 \times 10^7$, and consists of a dominant Plummer
profile that fits the inner parts of the radial velocity data
(Figure~\ref{results_table}), and a second, wider profile that models
the stars at large radius.
As a test of the veracity of our model we altered the model so that
$w$($r$) was linear (with endpoints in the range $0-1$ and with a
uniform prior) rather than $u$($r$) being linear. This had the effect
of reducing the log evidence to $\approx-7748$, so in this case the
double Plummer model is favoured by a factor $\sim 10^5$. This
best-fit model has a more subtle increase in width at large radii,
when compared with Figure \ref{results_figure1}. Presumably this is
because $w$($r$) being linear prevents more rapid fade-outs.
Correspondingly, the Plummer profile represented by the thin curve in
Figure \ref{results_figure2} was shifted slightly lower. Note that the
model in Figure~\ref{results_figure1} is narrower than the spread of
the data, because the spread also arises partly from observational
errors. The most extreme points at large radii are likely to be those
for which the intrinsic velocity dispersion is large {\it and} the
observational errors have pushed the points further away from the
mean. In Figure~\ref{results_figure2}, the Plummer profiles of the two
population components are shown. The inner component is a good fit to
the binned velocity dispersions by \cite{Lane09b}. \new{We now discuss
possible explanations for this two-component population, and calculate
an upper limit on when the second component was introduced.}
\begin{table}
\begin{center}
\caption{Inferred parameter values for the single Plummer and the
double Plummer fits to the 47 Tuc data. The values quoted are the
posterior mean $\pm$ the posterior standard deviation, when the
marginal posterior distributions were sufficiently symmetric for this
to be a reasonable summary. For the few parameters with asymmetric
posterior distributions, we instead give the symmetric 68\% credible
interval. The evidence values imply that if the two models were
equally likely before taking into account the data, the data makes the
double Plummer model more likely by a factor of $e^{17.3} \approx 3
\times 10^7$. For the double Plummer model, the first value listed is
for the component that dominates at $r=0$.}\label{results_table}
\begin{tabular}{lc}
Parameter & Value\\
\hline
{\bf Single Plummer Profile}\\
\hline
$\mu$ & -16.87 $\pm$ 0.17 km s$^{-1}$\\
$\sigma_0$ & 9.37 $\pm$ 0.32 km s$^{-1}$\\
$r_0$ & 9.27 $\pm$ 0.98 pc \\
log(evidence) & -7759.5 \\
\hline
{\bf Double Plummer Profile}\\
\hline
$\mu$ & -16.94 $\pm$ 0.12 km s$^{-1}$\\
$\sigma_0$ & 9.93 $\pm$ 0.43 km s$^{-1}$\\
& [5.70, 13.51] $\pm$ km s$^{-1}$\\
$r_0$ & 6.76 $\pm$ 0.94 pc\\
& [53.6, 4560] pc\\
$u_{\alpha}$ & 6.30 $\pm$ 1.30 \\
$u_{\beta}$& -2.30 $\pm$ 1.12\\
log(evidence) & -7742.2
\end{tabular}
\end{center}
\end{table}
\begin{figure*}
\begin{centering}
\includegraphics[scale=0.8]{figures/velocities.eps}
\caption{The radial velocities of stars in 47 Tuc together with the
best fit double Plummer model for the velocity distribution as a function
of radius. The inner part of the cluster is well modelled by the
dominant Plummer profile, while at larger radii, the second plummer
profile dominates. The radius at which the two profiles have equal
weight is 55 pc.\label{results_figure1}}
\end{centering}
\end{figure*}
\begin{figure*}
\begin{centering}
\includegraphics[scale=0.8]{figures/plummers.eps}
\caption{Binned velocity dispersion as a function of radius
\citep[from][]{Lane09b}, with the radial velocity profiles of the two
stellar populations from the best fit double Plummer model. The
Plummer profile that dominates at small radii is shown as the thick
black curve, while the thin curve shows the Plummer profile for the
stellar population that dominates at large
radii.\label{results_figure2}}
\end{centering}
\end{figure*}
\subsection{Evaporation}\label{evaporation}
\cite{Drukier07} carried out $N$-body simulations of GCs through
core-collapse and into post-core-collapse. They showed that the
evaporation of low-mass stars due to two-body interactions during
these phases alters the velocity dispersion profiles in predictable
ways. At approximately half the tidal radius ($r_t/2$) the velocity
dispersion reaches a minimum of $\sim40$\% of the central dispersion,
then rises to $\sim60$\% of the central level at $r_t$. These
criteria are certainly met within 47 Tuc \cite[again, see Figure 11
of][]{Lane09b}. \new{Furthermore, \cite{Lane09b} conclude that the
rise in velocity dispersion could be explained by evaporation due to
the core collapsing, and the Fokker-Planck models by \cite{Behler03}
show that 47 Tuc is nearing core-collapse.}
This appears to be reasonable evidence that 47 Tuc is
evaporating. However, based on the conclusions drawn by
\cite{Drukier07} and \cite{Lane10}, it is unclear how much Galactic
tidal fields affect the outer regions of Galactic globular clusters,
and what effect this has on the external velocity dispersion
profile. \new{While our best-fit double Plummer model matches the
overall form of the trend shown by \cite{Drukier07} reasonably well
(see Figure \ref{results_figure1}), this scenario does not explain its
multiple stellar populations \cite[although these might be explained
by chemical anomalies, e.g.][and references therein]{Piotto07}, nor
its low M/L$_{\rm V}$ in comparison with its mass \cite[see Figure 6
of][]{Lane10}, or its extreme rotation. In addition, the extra-tidal
stars are spread uniformly across all regions of the colour magnitude
diagram \cite[see Figure 3 of][]{Lane09b}, which is inconsistent with
the the two-component population being a consequence of evaporating
low-mass stars. We cannot, however, completely discount evaporation
without detailed chemical abundance information.}
\subsection{Merger}\label{merger}
\new{Another scenario, which appears to explain most of the unusual
properties of 47 Tuc, is that it has undergone a merger in its past
\cite[note that this is not the first evidence for such a merger
within Galactic GCs, see][]{Ferraro02}. Several observed quantities
can be explained by this hypothesis: (1) the bimodality of the carbon
and nitrogen line strengths, (2) the mixed stellar populations, (3)
the large rotational velocity, (4) the low M/L$_{\rm V}$ compared with
total mass and (5) the increase in velocity dispersion in the
outskirts of the cluster.}
\new{In addition to our evidence for two kinematically distinct
stellar populations,} \cite{Anderson09} showed that 47 Tuc also has
two distinct sub-giant branches, one of which is much broader than the
other, as well as a broad main sequence. The authors determined that
this broadening may be due to a combination of metallicities and ages,
and it is known from previous studies \cite[e.g.][]{Harbeck03} that 47
Tuc has a \new{bimodality in its carbon and nitrogen line
strengths. While this bimodality is not unique, a merger could explain
its origin. Therefore, another possible scenario, which explains many
of the properties of 47 Tuc, is a past merger event.}
\new{While it might seem unlikely that this is the remnant of a
merger, extant kinematic signatures of subgalactic scale hierarchical
merging do exist \cite[e.g. within $\omega$ Centauri;][]{Ferraro02},
and there have been hints that the distinct photometric populations in
47 Tuc might be remnants of a past merger \cite[e.g.][]{Anderson09}. A
possible explanation for this merger hypothesis is given in Section
\ref{primordial}.}
\new{\subsection{Initial Formation}\label{primordial}
The two components may have formed at the same epoch, and the distinct
kinematic populations are, therefore, extant remnants from the
formation of the cluster itself. If GCs form from a single cloud
\cite[see][for a discussion of GCs as simple stellar
populations]{Kalirai09}, it is possible for the proto-cluster cloud to
initially contain two overdensities undergoing star formation
independently at almost the same time. In this case, the two
proto-clusters, which would inevitably be in mutual orbit due to the
initially bound nature of the proto-cluster cloud, eventually
coalesced through the loss of angular momentum due to dynamical
friction.
Note that this scenario is similar to the capture of a satellite
described by \cite{Ferraro02} to explain the merger hypothesis for
$\omega$ Centauri, and explains the two kinematic and photometric
populations, and the high rotation rate assuming the proto-cluster
cloud initially had a large angular momentum. It might also explain
the low ${\rm M/L_{V}}$ of this cluster.
Interestingly, \cite{Vesperini09} show that clusters with initially
segregated masses evolve more slowly than non-segregated clusters,
having looser cores and reaching core-collapse much later. Because the
core of 47 Tuc is highly concentrated and near core-collapse
\cite[e.g.][]{Behler03} it must be very old if it was initially
mass-segregated. 47 Tuc is thought to be 11--14\,Gyr old
\cite[e.g.][]{Gratton03,Kaluzny07}, hence initial mass segregation is
plausible. However, even if 47 Tuc is $\sim11$\,Gyr old the original
populations would be kinematically indistinguishable at the present
epoch (Section \ref{time}) indicating that some other process was the
cause of the two extant populations described in this
Letter. Furthermore, Milky Way GC formation ended about 10.8\,Gyr ago
\cite[e.g.][]{Gratton03}, long before the upper limit for the initial
mixing of the two populations derived in Section \ref{time}.
\subsection{Multiple Star Formation Epochs}\label{SF}
Two star formation epochs in GCs result in the radial separation of
the two populations, with the second generation initially concentrated
in the core \cite[e.g.][and references
therein]{D'Ercole08}. Furthermore, the kinematics of the second
generation are virtually independent of that of the first generation
and the second generation contain chemical anomalies which are
consistent with having arisen in the envelopes of the first generation
\cite[e.g.][]{Decressin07,D'Ercole08}.
This scenario might explain the two kinematic populations and chemical
anomalies of 47 Tuc, however, it is unclear how this would cause the
extreme rotation, or the anomalous ${\rm M/L_{V}}$.}
\subsection{Time-scale for the Initial Mixing of the Second Population}\label{time}
\new{The scenarios described in Sections \ref{merger},
\ref{primordial}, and \ref{SF} require a second population beginning
to mix with an initial population at a particular epoch.}
\cite{Decressin08} performed detailed $N$-body simulations of globular
clusters containing two distinct populations of stars to determine
their dynamical mixing time via two-body relaxation. They concluded
that $\sim2$ relaxation times are required to completely homogenise
the populations, a timescale that is virtually independent of the
number of stars in the cluster. They also concluded that the
information on the stellar orbital angular momenta of the two
populations is lost on a similar timescale. Since we observe two
distinct, extant kinematic populations\new{, an upper limit can be
placed on when the two populations began to mix.}
\cite{Decressin08} showed that the relaxation time ($t_{rh}$) of a GC
decreases by $\sim0.29t_{rh}$ for each consecutive relaxation time,
i.e. $t_{rh}(i)\sim0.71t_{rh}(i+1)$. Because the two populations are
kinematically distinct at the present epoch, and the current
relaxation time of 47 Tuc is $t_{rh}\approx3.02$\,Gyr
\cite[][]{Harris96}, \new{an upper limit on when the two populations
began to mix} is $7.3\pm1.5$\,Gyr ago. \new{Note that the scenarios
discussed in Sections \ref{merger} and \ref{primordial} can be
temporally assessed in this way only if the merger did not disrupt the
core of the larger component. This is true for minor mergers in blue
compact dwarf galaxies \cite[][]{Sung02}, therefore, if we assume a
merger between objects with a mass ratio of 9:1 \cite[][show the ratio
of the number densities on the two sub-giant branches of 47 Tuc is
$\sim$9:1]{Anderson09}, this is also likely to hold for 47 Tuc.}
\section{Conclusions}
With a Bayesian analysis of the velocity distribution of 47 Tuc, we
conclude that \new{the scenario which best explains the observed
properties of 47 Tuc is that} we are seeing the first kinematic
evidence of a merger in 47 Tuc, which occurred
$\lesssim7.3\pm1.5$\,Gyr ago. \new{Extant kinematic populations from
the merger of formation remnants is a plausible explanation as to the
reason for this merger, assuming the two components evolved separately
and merged $\lesssim7.3\pm1.5$\,Gyr ago.} This scenario \new{could
explain the two-component population,} its extreme rotational
velocity, mixed stellar populations and low M/L$_{\rm V}$ compared
with its mass.
\new{All the other explanations for this two-component population are
less plausible than the merger hypothesis. Evaporation of low mass
stars is unlikely due to the various stellar types that are found
beyond the tidal radius, and this also cannot explain its low
M/L$_{\rm V}$. The possibility of multiple star formation epochs does
not explain the large rotational velocity, nor the low M/L$_{\rm V}$.}
\new{Detailed chemical abundances and} high resolution $N$-body
simulations of merging globular clusters are now required to further
\new{analyse the} merger scenario. \new{Several observed quantities
need to be} addressed, namely how much angular momentum can be
imparted through a 9:1 merger, what consequence it has on the velocity
dispersion in the outer regions of the cluster over dynamical
timescales, and what effect it would have on the global M/L$_{\rm V}$.
Of course, alternative explanations exist for the observed rise in
dispersion. For example, if GCs form in a similar fashion to Ultra
Compact Dwarf galaxies, there may be a large quantity of DM in the
outskirts of the cluster as discussed by \cite{Baumgardt08}. However,
no evidence exists supporting GCs forming in this manner and GCs do
not appear to have significant dark matter components
\cite[e.g.][]{Lane09a,Lane09b,Lane10}.
\acknowledgments
This project has been supported by the University of Sydney, the
Anglo-Australian Observatory, the Australian Research Council, the
Hungarian OTKA grant K76816 and the Lend\"ulet Young Researchers
Program of the Hungarian Academy of Sciences. \new{GyMSz acknowledges
the Bolyai Fellowship of the HAS.} RRL thanks Martine L. Wilson for
everything.
|
\section{Introduction}
The aim of our work is to construct and research the fundamental solution of the formal
KZ (Knizhnik-Zamolodchikov) equation via iterated integrals.
First we establish the decomposition theorem for the normalized fundamental solution of
the formal KZ equation on the moduli space ${\mathcal M}_{0,5}$
(or, the formal KZ equation of two variables).
Next we show that,
by using iterated integrals, it can be viewed as a generating function of hyperlogarithms
of the type ${\mathcal M}_{0,5}$. The decomposition theorem says that the normalized fundamental
solution decomposes to a product of two factors which are the normalized fundamental
solutions of the formal (generalized) KZ equations of one variable. Comparing the different ways
of decomposition gives the generalized harmonic product relations of the hyperlogarithms.
These relations properly contain the harmonic product of multiple polylogarithms.
The most simple case of the harmonic product is the following: Let us define
\begin{align*}
\operatorname{Li}_{k_1,\ldots,k_r}(z)&=\sum_{n_1>\cdots>n_r>0}\frac{z^{n_1}}{n_1^{k_1}\cdots n_r^{k_r}},\\
\operatorname{Li}_{k_1,\ldots,k_{i+j}}(i,j;z_1,z_2)&=\sum_{n_1>\cdots>n_{i+j}>0}
\frac{z_1^{n_1} z_2^{n_{i+1}}}{n_1^{k_1}\cdots n_{i+j}^{k_{i+j}}}.
\end{align*}
Then we obtain
\begin{align}
\operatorname{Li}_{k}(z_1)\operatorname{Li}_{l}(z_2)&=\sum_{m>0}\frac{z_1^m}{m^k} \sum_{n>0}\frac{z_2^n}{n^l}
=\left(\sum_{m>n>0}+\sum_{m=n>0}+\sum_{n>m>0}\right)\frac{z_1^m z_2^n}{m^k n^l} \nonumber \\
&=\operatorname{Li}_{k,l}(1,1;z_1,z_2)+\operatorname{Li}_{k+l}(z_1z_2)+\operatorname{Li}_{l,k}(1,1;z_2,z_1). \tag{HPMPL} \label{hpmpl}
\end{align}
Taking the limit, we have the harmonic product of multiple zeta values
\begin{align}\tag{HPMZV} \label{hpmzv}
\zeta(k) \zeta(l) = \zeta(k,l)+\zeta(k+l)+\zeta(l,k).
\end{align}
(The harmonic product of multiple zeta values is considered from the viewpoint of
arithmetic geometry in \cite{BF},\ \cite{DT},\ \cite{F}.)
Moreover we consider the transformation theory of the fundamental solution
of the formal KZ equation of two variables
and derive the five term relation for the dilogarithm due to Hill \cite{Le},
\begin{align}
\operatorname{Li}_2(z_1z_2)= \operatorname{Li}_2 & \left(\frac{-z_1(1-z_2)}{1-z_1}\right) +
\operatorname{Li}_2\left(\frac{-z_2(1-z_1)}{1-z_2}\right) \nonumber \\
& + \operatorname{Li}_2(z_1)+\operatorname{Li}_2(z_2)+\frac{1}{2}\log^2\left(\frac{1-z_1}{1-z_2}\right).
\tag{5TERM} \label{5term}
\end{align}
For detailed accounts of the results in this note, see \cite{OU1} and \cite{OU2}.
The transformation theory of the formal KZ equation of one variable
(or the formal KZ equation on ${\mathcal M}_{0,4}$) is studied in \cite{OkU}.
\paragraph{Acknowledgment}
The authors express their gratitude to Professor Hideaki Morita for giving them
a chance of a lecture.
The second author is partially supported by JPSP Grant-in-Aid No. 19540056.
\section{The formal KZ equation on ${\mathcal M}_{0,n}$}
\subsection{Definition of the formal KZ equation}
First we introduce the formal KZ equation: It is defined on the configuration space
of $n$ points of ${\mathbf P}^1$ ($=$ the complement of the hyperplane arrangement associated
with Dynkin diagram of $A_{n-1}$-type), which is by definition
\begin{align*}
({\mathbf P}^1)^n_*=\{(x_1,\ldots,x_n) \in
\underbrace{{\mathbf P}^1\times\cdots\times{\mathbf P}^1}_{n} \;|\; x_i\neq x_j\;\; (i\neq j)\}.
\end{align*}
The \textbf{infinitesimal pure braid Lie algebra}
\begin{align*}
{\mathfrak X}={\mathfrak X}(\{X_{ij}\}_{1\le i,j \le n}) := {\mathbf C}\{X_{ij}\;|\;{1\le i,j \le n}\}
\Big/ \eqref{ipbr}
\end{align*}
is a graded Lie algebra for the lower central series of the fundamental group
of $({\mathbf P}^1)^n_*$ \ \cite{I}. It is generated by the formal elements
$\{X_{ij}\}_{1\le i,j \le n}$ with the defining relations \eqref{ipbr}
(the infinitesimal pure braid relations)
\begin{align}
\begin{cases}
X_{ij}=X_{ji}, \quad &X_{ii}=0,\\
\sum_j X_{ij}=0 \quad (\forall i ),
\quad &[X_{ij},X_{kl}]=0 \quad (\{i,j\} \cap \{k,l\} = \emptyset).
\tag{IPBR}\label{ipbr}
\end{cases}
\end{align}
By ${\mathcal U}({\mathfrak X})$, we denote the universal enveloping algebra of ${\mathfrak X}$. It has the unit ${\mathbf I}$
and has the grading with respect to the homogeneous degree of an element:
\begin{align*}
{\mathcal U}({\mathfrak X})= \bigoplus_{s=0}^{\infty} \, {\mathcal U}_s({\mathfrak X}).
\end{align*}
The formal KZ equation is by definition
\begin{align}\tag{KZ}\label{KZeq}
dG=\varOmega G, \qquad
\varOmega = \sum_{i<j}\xi_{ij}X_{ij}, \qquad \xi_{ij}=d\log(x_i-x_j),
\end{align}
which is a ${\mathfrak X}$-valued total differential equation (or, a connection) on $({\mathbf P}^1)^n_*$.
(Such a formal equation was considered in \cite{Ha},\ \cite{De},\ \cite{Dr},\ \cite{W}.)
The 1-forms $\xi_{ij}$'s satisfy only the \textbf{Arnold relations} \cite{A}
as non-trivial relations of degree 2:
\begin{align}\tag{AR}\label{AR}
\xi_{ij} \wedge \xi_{ik} + \xi_{ik} \wedge \xi_{jk} +
\xi_{jk} \wedge \xi_{ij} = 0.
\end{align}
From \eqref{ipbr} and \eqref{AR}, one can see that \eqref{KZeq} is integrable and
has $\mathrm{PGL}(2,{\mathbf C})$-invariance.
Hence \eqref{KZeq} can be viewed as an equation on the moduli space
\begin{align*}
{\mathcal M}_{0,n}=\operatorname{PGL}(2,{\mathbf C}) \backslash ({\mathbf P}^1)^n_*.
\end{align*}
Hereafter we will call \eqref{KZeq} the \textbf{formal KZ equation on the moduli space} ${\mathcal M}_{0,n}$.
\subsection{The formal KZ equation on ${\mathcal M}_{0,4}$ and ${\mathcal M}_{0,5}$}
For analysis of \eqref{KZeq}, it is convenient to use the \textbf{cubic coordinates} on
${\mathcal M}_{0,n}$ \cite{B}. Introducing the simplicial coordinates $\{y_i\}$ by
\begin{align*}
y_i=\frac{x_i-x_{n-2}}{x_i-x_n}\frac{x_{n-1}-x_{n}}{x_{n-1}-x_{n-2}}
\qquad (i=1,\ldots,n-3),
\end{align*}
(fixing three points $y_n=\infty,\; y_{n-1}=1,\; y_{n-2}=0$)
the cubic coordinates $\{z_i\}$ are defined by blowing up at the origin,
\begin{align*}
y_i \,=\, z_1 \cdots z_i \qquad (i=1,\dots,n-3).
\end{align*}
We give representations of \eqref{KZeq} for $n=4, 5$. In the cubic coordinates of ${\mathcal M}_{0,4}$,
we put $z=z_1$ and $Z_1=X_{12}, Z_{11}=-X_{13}$. Then \eqref{KZeq} is represented as
\begin{align}\tag{1KZ}\label{1KZeq}
dG=\varOmega G, \qquad
\varOmega=\zeta_1Z_1+\zeta_{11}Z_{11}, \qquad
\zeta_1=\frac{dz}{z},\; \zeta_{11}=\frac{dz}{1-z}.
\end{align}
which is referred to as the \textbf{formal KZ equation of one variable}.
The singular divisors of this equation are $D({\mathcal M}_{0,4}^{cubic}):=\{z=0,1,\infty\}$.
The Lie algebra ${\mathfrak X}$ is a free Lie algebra generated by $Z_1, Z_{11}$, and
\eqref{AR} reduces to the trivial one $\zeta_1 \wedge \zeta_{11}=0$.
In the case of ${\mathcal M}_{0,5}$, we put
\begin{align*}
Z_1=X_{12}+X_{13}+X_{23},\; Z_{11}=-X_{14},\; Z_2=X_{23},\; Z_{22}=-X_{12},\ Z_{12}=-X_{24}.
\end{align*}
In the cubic coordinates of ${\mathcal M}_{0,5}$, (KZ) reads as
\begin{gather}\tag{2KZ}\label{2KZeq}
dG = \varOmega G, \qquad
\varOmega=\zeta_1 Z_1 + \zeta_{11} Z_{11} + \zeta_2 Z_2 + \zeta_{22} Z_{22} + \zeta_{12} Z_{12}, \\
\zeta_1=\frac{dz_1}{z_1},\;\; \zeta_{11}=\frac{dz_1}{1-z_1},\;\;
\zeta_2=\frac{dz_2}{z_2},\;\; \zeta_{22}=\frac{dz_2}{1-z_2},\;\;
\zeta_{12}=\frac{d(z_1z_2)}{1-z_1z_2}, \notag
\end{gather}
which is referred to as the \textbf{formal KZ equation of two variables}.
The singular divisors of this equation are
$D({\mathcal M}_{0,5}^{cubic}):=\{z_1=0,1,\infty\}\cup\{z_2=0,1,\infty\}\cup\{z_1z_2=1\}$.
The Lie algebra ${\mathfrak X}$ is generated by the five elements $Z_1, Z_{11}, Z_2, Z_{22}, Z_{12}$
with the defining relations
\begin{align}
\begin{cases}
[Z_1,Z_2]=[Z_{11},Z_2]=[Z_1,Z_{22}]=0, \\
[Z_{11},Z_{22}]=[-Z_{11},Z_{12}]=[Z_{22},Z_{12}]=[-Z_1+Z_2,Z_{12}].
\end{cases} \tag{IPBR'}\label{ipbr'}
\end{align}
Non trivial relations among \eqref{AR} are
\begin{align}
\begin{cases}
(\zeta_1+\zeta_2) \wedge \zeta_{12} = 0,\\
\zeta_{11} \wedge \zeta_{12} + \zeta_{22} \wedge (\zeta_{11} - \zeta_{12})
- \zeta_2 \wedge \zeta_{12}=0.
\end{cases} \tag{AR'}\label{AR'}
\end{align}
The following is a figure of the divisors $D({\mathcal M}^{cubic}_{0,5})$.
Note that they are normal crossing at $(z_1,z_2)=(0,0),(1,0),(0,1)$.
\begin{center}
\setlength{\unitlength}{1cm}
\begin{picture}(6,5.5)(0,0)
\put(0,0){\scalebox{0.5}{\includegraphics{m05.eps}}}
\put(-0.5,0){$(0,0)$}
\put(3.1,0){$(1,0)$}
\put(-0.5,3.1){$(0,1)$}
\put(3.1,3.1){$(1,1)$}
\put(4.4,0.3){$z_1$}
\put(0.3,4.4){$z_2$}
\end{picture}
\end{center}
\section{The fundamental solution of the formal KZ equation on ${\mathcal M}_{0,4}$}
\subsection{A free shuffle algebra and iterated integral on ${\mathcal M}_{0,4}$}
For a free shuffle algebra $S=S(a_1,\ldots,a_r)$ generated by the alphabet $a_1,\ldots,a_r$,
we denote by ${\mathbf 1}$ the unit, by $\circ$ the product of concatenation and by $\sh$
the shuffle product:
\begin{gather*}
S=({\mathbf C}\langle a_1,\ldots,a_r\rangle, \sh),\\
w \sh {\mathbf 1} = {\mathbf 1} \sh w = {\mathbf 1},\\
(a_i \circ w) \sh (a_j \circ w')=a_i \circ (w \sh (a_j \circ w'))
+ a_j \circ ((a_i \circ w) \sh w').
\end{gather*}
It is a graded algebra with respect to the homogeneous degree of an element.
Let $\zeta_1, \zeta_{11}$ be the 1-forms in \eqref{1KZeq}, and $S(\zeta_1,\zeta_{11})$
a free shuffle algebra generated by them. For any word
$\varphi=\omega_1 \circ \cdots \circ \omega_r \quad (\omega_i=\zeta_1, \mbox{or}, \zeta_{11})$
in $S(\zeta_1,\zeta_{11})$, we set the iterated integral by
\begin{align*}
\int_{z_0}^z \,\, \varphi
=\int_{z_0}^z \,\, \omega_1(z') \, \int_{z_0}^{z'}\,\,\omega_2 \circ \cdots \circ \omega_r,
\end{align*}
which gives a many-valued analytic function on ${\mathbf P}^1-D({\mathcal M}^{cubic}_{0,4})$.
For $\varphi,\psi \in S(\zeta_1,\zeta_{11})$, we have
\begin{align*}
\int (\varphi \sh \psi)=\left(\int \varphi \right)\left(\int \psi \right).
\end{align*}
A free shuffle algebra has the structure of a Hopf algebra, and
$S(\zeta_1,\zeta_{11})$ is a dual Hopf algebra of the universal enveloping
algebra ${\mathcal U}({\mathfrak X})$.
\subsection{The fundamental solution of \eqref{1KZeq}}
Next we consider the fundamental solution of \eqref{1KZeq} normalized at the origin $z=0$.
We denote it by ${\mathcal L}(z)$. It is a solution satisfying the following condition:
\begin{align*}
{\mathcal L}(z)={\hat{\mathcal L}}(z)z^{Z_1}
\end{align*}
where ${\hat{\mathcal L}}(z)$ is represented as
\begin{align*}
{\hat{\mathcal L}}(z)= \sum_{s=0}^{\infty} {\hat{\mathcal L}}_s(z), \quad {\hat{\mathcal L}}_s(z) \in {\mathcal U}_s({\mathfrak X}),
\quad {\hat{\mathcal L}}_s(0)=0\;\; (s>0), \quad {\hat{\mathcal L}}_0(z)={\mathbf I}.
\end{align*}
It is easy to see that ${\hat{\mathcal L}}_s(z)$ satisfies the following recursive equation:
\begin{align*}
\frac{d{\hat{\mathcal L}}_{s+1}}{dz} = \frac{1}{z}[Z_1, {\hat{\mathcal L}}_{s}] + \frac{1}{1-z}Z_{11} {\hat{\mathcal L}}_{s}
\quad (s=0,1,2,\dots).
\end{align*}
Since the term $\frac{1}{z}[Z_1, {\hat{\mathcal L}}_{s}]$ is holomorphic at $z=0$, ${\hat{\mathcal L}}_{s+1}(z)$
is uniquely determined by
\begin{align*}
{\hat{\mathcal L}}_{s+1}(z)= \int_0^z \Big( \frac{1}{z}[Z_1, {\hat{\mathcal L}}_{s}] + \frac{1}{1-z}Z_{11}{\hat{\mathcal L}}_{s} \Big) \, dz.
\end{align*}
In terms of iterated integral, it is expressed as
\begin{align*}
{\hat{\mathcal L}}_s(z) &=\sum_{k_1+\cdots+k_r=s}
\Big\{\int_0^z \zeta_1^{k_1-1}\circ \zeta_{11} \circ\cdots\circ \zeta_1^{k_r-1}
\circ \zeta_{11}\Big\} \notag \\
& \hphantom{\sum_{k_1+\cdots+k_r=s} \Big\{} \times \operatorname{ad}(Z_1)^{k_1-1}\mu(Z_{11})
\cdots\operatorname{ad}(Z_1)^{k_r-1}\mu(Z_{11})({\mathbf I}).
\end{align*}
Here $\operatorname{ad}(Z_1) \in \operatorname{End}({\mathcal U}({\mathfrak X}))$ stands for the adjoint operator by $Z_1$,
and $\mu(Z_{11}) \in \operatorname{End}({\mathcal U}({\mathfrak X}))$ the multiplication of $Z_{11}$ from the left.
From these considerations, it follows that \textbf{the fundamental solution normalized
at $z=0$ exists and is unique}.
The iterated integral in the right hand side is a \textbf{multiple polylogarithm
of one variable}:
\begin{align}\tag{1MPL} \label{1mpl}
\operatorname{Li}_{k_1,\dots,k_r}(z) = \int_0^z \zeta_1^{k_1-1}\circ \zeta_{11}
\circ\cdots\circ \zeta_1^{k_r-1} \circ \zeta_{11}.
\end{align}
If $|z|<1$, it has a Taylor expansion
\begin{align*}
\operatorname{Li}_{k_1,\dots,k_r}(z)=\sum_{n_1>n_2>\cdots>n_r>0} \frac{z^{n_1}}{n_1^{k_1}\cdots n_r^{k_r}}.
\end{align*}
If $k_1 \geq 2$, we have
\begin{align*}
\lim_{z \to 1-0}\, \operatorname{Li}_{k_1,\dots,k_r}(z) = \zeta(k_1,\dots,k_r),
\end{align*}
where the right side above is a \textbf{multiple zeta value},
\begin{align}\tag{MZV}\label{mzv}
\zeta(k_1,\dots,k_r)=\sum_{n_1>\cdots>n_r>0} \frac{1}{n_1^{k_1}\cdots n_r^{k_r}}.
\end{align}
\subsection{The fundamental solution of the formal generalized KZ equation of one variable}
Let us consider a generalization of \eqref{1KZeq}.
For mutually distinct points $a_1, \ldots, a_m \in {\mathbf C}-\{0\}$ we set
\begin{equation}\tag{G1KZ} \label{g1KZeq}
dG= \varOmega G, \quad \varOmega=\frac{dz}{z}X_0+\sum_{i=1}^{m}\frac{a_i dz}{1-a_i z}X_i.
\end{equation}
Here the coefficients $X_0, X_1,\ldots, X_m$ are free formal elements.
For $r=1, a_1=1$, this is the formal KZ equation of one variable.
This is a differential equation of the Schlesinger type
with regular singular points $0,1/a_1,\ldots, 1/a_m, \infty$.
We call \eqref{g1KZeq} the \textbf{formal generalized KZ equation of one variable}.
Let ${\mathfrak X}={\mathbf C}\{X_0,X_1,\ldots,X_m \}$ be a free Lie algebra generated by
$X_0,X_1,\dots,X_m$, and ${\mathcal U}({\mathfrak X})$ the universal enveloping algebra.
The free shuffle algebra
$S(\xi_0, \xi_1,\ldots, \xi_m)$ where
\begin{align*}
\xi_0=\frac{dz}{z}, \quad\quad \xi_i=\frac{a_idz}{1-a_iz}, \quad (1 \leq i \leq m),
\end{align*}
is a dual Hopf algebra of ${\mathcal U}({\mathfrak X})$.
\textbf{The fundamental solution ${\mathcal L}(z)$ normalized at the origin $z=0$ of this equation
exists and is unique}. It satisfies the following conditions:
\begin{align*}
{\mathcal L}(z)={\hat{\mathcal L}}(z)z^{X_0}
\end{align*}
where ${\hat{\mathcal L}}(z)$ is represented as
\begin{align*}
{\hat{\mathcal L}}(z)= & \sum_{s=0}^{\infty} {\hat{\mathcal L}}_s(z), \quad {\hat{\mathcal L}}_s(z) \in {\mathcal U}_s({\mathfrak X}),
\quad {\hat{\mathcal L}}_s(0)=0\;\; (s>0), \quad {\hat{\mathcal L}}_0(z)={\mathbf I}, \\*
{\hat{\mathcal L}}_s(z) &= \sum_{\substack{k_1+\cdots+k_r=s\\i_1,\ldots,i_r \in \{1,\ldots,m\}}}
L({}^{k_1}a_{i_1}\cdots{}^{k_r}a_{i_r};z) \notag \\*
& \hphantom{\sum_{i_1,\ldots,i_r \in \{1,\ldots,m\}}} \times
\operatorname{ad}(X_0)^{k_1-1}\mu(X_{i_1})\cdots\operatorname{ad}(X_0)^{k_r-1}\mu(X_{i_r})({\mathbf I}).
\end{align*}
Here $L({}^{k_1}a_{i_1}\cdots{}^{k_r}a_{i_r};z)$ is a
\textbf{hyperlogarithm of the general type}:
\begin{align}\tag{HLOG} \label{hlog}
L({}^{k_1}a_{i_1}\cdots{}^{k_r}a_{i_r};z)
:= \int_0^z \xi_0^{k_1-1}\circ \xi_{i_1} \circ \xi_0^{k_2-1}\circ
\xi_{i_2} \circ\cdots\circ \xi_0^{k_r-1} \circ \xi_{i_r}.
\end{align}
For $r=1$ and $a_1=1$, this is \eqref{1mpl}.
If $|z|<\min\{\frac{1}{|a_{i_1}|},\ldots,\frac{1}{|a_{i_r}|}\}$, it has a Taylor expansion
\begin{align*}
L({}^{k_1}a_{i_1}\cdots{}^{k_r}a_{i_r};z)=
\sum_{n_1>n_2>\cdots>n_r>0}\frac{a_{i_1}^{n_1-n_2}a_{i_2}^{n_2-n_3}\cdots
a_{i_r}^{n_r}}{n_1^{k_1}\cdots n_r^{k_r}} z^{n_1}.
\end{align*}
\section{The fundamental solution of the formal KZ equation on ${\mathcal M}_{0,5}$}
\subsection{The reduced bar algebra and iterated integrals on ${\mathcal M}_{0,5}$}
Let $S=S(\zeta_1,\zeta_{11},\zeta_2,\zeta_{22},\zeta_{12})$ be a free shuffle algebra
generated by $\zeta_1,\zeta_{11},\zeta_2,\zeta_{22},\zeta_{12}$ which are 1-forms in \eqref{2KZeq}.
The iterated integral of an element in S, in general, depends on the integral path.
We want to construct a shuffle subalgebra of S such that the iterated integral of any
element in this subalgebra depends only on the homotopy class of the integral path.
We say that an element
\begin{align*}
S \ni \varphi \,=\, \sum_{I=(i_1,\ldots,i_s)} \ c_I \omega_{i_1}\circ\cdots\circ\omega_{i_s},
\end{align*}
where $\omega_{i} \in \{\zeta_1,\zeta_{11},\zeta_2,\zeta_{22},\zeta_{12}\}$, satisfies
\textbf{Chen's integrability condition} \cite{C1} if and only if
\begin{align}\tag{CIC}\label{cic}
\sum_I c_I \; \omega_{i_1}\otimes\cdots\otimes\omega_{i_l}\wedge
\omega_{i_{l+1}}\otimes\cdots\otimes\omega_{i_s}=0
\end{align}
holds for any $l$ ($1 \le l<s$) as a multiple differential form.
Let ${\mathcal B}$ be the subalgebra of elements satisfying \eqref{cic}. We call it
the \textbf{reduced bar algebra}, which coincides with the 0-th cohomology of
the reduced bar complex \cite{C2} associated with the Orlik-Solomon algebra \cite{OT}
generated by $\zeta_1,\zeta_{11},\zeta_2,\zeta_{22},\zeta_{12}$.
For any element $\varphi \in {\mathcal B}$, the iterated integral
\begin{align*}
\int_{(z^{(0)}_{1},z^{(0)}_{2})}^{(z_1,z_2)} \ \varphi
\end{align*}
gives a many-valued analytic function on ${\mathbf P}^1 \times {\mathbf P}^1 - D({\mathcal M}_{0,5}^{cubic})$.
Let us consider more on the structure of ${\mathcal B}$: It is a graded algebra;
${\mathcal B}=\bigoplus_{s=0}^\infty {\mathcal B}_s, \;\; {\mathcal B}_s={\mathcal B} \cap S_s$ where $S_s$ denotes
the degree s part of $S$: We have
\begin{align*}
& \hspace{10mm} {\mathcal B}_0={\mathbf C}{\mathbf 1},\quad
{\mathcal B}_1={\mathbf C}\zeta_1\oplus{\mathbf C}\zeta_{11}\oplus{\mathbf C}\zeta_2\oplus{\mathbf C}\zeta_{22}\oplus{\mathbf C}\zeta_{12}, \\
& \hspace{10mm} {\mathcal B}_2=\bigoplus_{\omega \in A}{\mathbf C} \omega\circ\omega \oplus
\bigoplus_{i=1,2}{\mathbf C} \zeta_i\circ\zeta_{ii} \oplus \bigoplus_{i=1,2}{\mathbf C} \zeta_{ii}\circ\zeta_i\\*
& \hspace{15mm} \oplus \bigoplus_{\substack{\omega_1=\zeta_1,\zeta_{11}\\ \omega_2=\zeta_2,\zeta_{22}}}
{\mathbf C}(\omega_1\circ\omega_2+\omega_2\circ\omega_1) \oplus \bigoplus_{\omega \in A-\{\zeta_{12}\}}{\mathbf C}
(\omega\circ\zeta_{12}+\zeta_{12}\circ\omega)\\*
& \hspace{15mm} \oplus {\mathbf C} (\zeta_1\circ\zeta_{12}+\zeta_2\circ\zeta_{12}) \oplus
{\mathbf C} (\zeta_{11}\circ\zeta_{12}+\zeta_{22}\circ\zeta_{11}-\zeta_{22}\circ\zeta_{12}
-\zeta_2\circ\zeta_{12})
\end{align*}
where $A:=\{\zeta_1,\zeta_{11},\zeta_2,\zeta_{22},\zeta_{12}\}$.
For $s>2$, ${\mathcal B}_s$ is characterized as follows \cite{B};
\begin{align*}
{\mathcal B}_s&=\bigcap_{j=1}^{s-1}{\mathcal B}_j\circ{\mathcal B}_{s-j}
=\bigcap_{j=0}^{s-2}\underbrace{{\mathcal B}_1\circ\cdots\circ{\mathcal B}_1}_{j\text{ times}}
\circ{\mathcal B}_2\circ\underbrace{{\mathcal B}_1\circ\cdots\circ{\mathcal B}_1}_{s-j-2\text{ times}}.
\end{align*}
Put
\begin{align*}
\zeta_{12}^{(1)}= \frac{z_2dz_1}{1-z_1z_2}, \qquad
\zeta_{12}^{(2)}= \frac{z_1dz_2}{1-z_1z_2}.
\end{align*}
One can define a linear map
\begin{align*}
\iota_{1 \otimes 2 }:{\mathcal B} \longrightarrow
S(\zeta_1,\zeta_{11},\zeta_{12}^{(1)}) \otimes S(\zeta_2,\zeta_{22})
\end{align*}
by the following procedure;
\begin{enumerate}
\item \ pick up the terms only having a form $\psi_1 \circ \psi_2 \in
S(\zeta_1,\zeta_{11},\zeta_{12}) \circ S(\zeta_2,\zeta_{22})$.
\item \ change each term $\psi_1 \circ \psi_2$ to $\psi_1 \otimes \psi_2 \in
S(\zeta_1,\zeta_{11},\zeta_{12}) \otimes S(\zeta_2,\zeta_{22})$.
\item \ replace $\zeta_{12}$ to $\zeta_{12}^{(1)}$.
\end{enumerate}
A linear map
\begin{align*}
\iota_{2 \otimes 1 }:{\mathcal B} \longrightarrow
S(\zeta_2,\zeta_{22},\zeta_{12}^{(2)}) \otimes S(\zeta_1,\zeta_{11})
\end{align*}
is defined in the same way.
One can show that
\begin{align*}
{\mathcal U}({\mathfrak X}) & \;\cong\; {\mathcal U}({\mathbf C}\{Z_1,Z_{11},Z_{12}\})\otimes {\mathcal U}({\mathbf C}\{Z_2,Z_{22}\}) \\
& \;\cong\; {\mathcal U}({\mathbf C}\{Z_2,Z_{22},Z_{12}\})\otimes {\mathcal U}({\mathbf C}\{Z_1,Z_{11}\})
\end{align*}
and that ${\mathcal B}$ is a dual Hopf algebra of ${\mathcal U}({\mathfrak X})$.
Through this isomorphism and the duality, one can show the following proposition:
\begin{prop}\label{prop:b}
The maps $\iota_{1 \otimes 2}$ and $\iota_{2 \otimes 1}$ are $\sh$-isomorphisms.
\end{prop}
(Such an isomorphism is also obtained by \cite{B}.)
Let ${\mathcal B}^0$ be the subspace of ${\mathcal B}$ spanned by elements
which have no terms ending with $\zeta_1$ and $\zeta_2$, and
$S^0(\zeta_1,\zeta_{11},\zeta_{12}^{(1)})$ (resp. $S^0(\zeta_2,\zeta_{22})$)
the subspace spanned by elements which have no terms ending with $\zeta_1$
(resp. $\zeta_2$), and so on. They are shuffle algebras. One can show the following
isomorphism:
\begin{prop}\label{prop:b0}
By $\iota_{1 \otimes 2}$ and $\iota_{2 \otimes 1}$,
\begin{align*}
{\mathcal B}^0 \, \cong \, S^0(\zeta_1,\zeta_{11},\zeta_{12}^{(1)})\otimes S^0(\zeta_2,\zeta_{22})
\, \cong \, S^0(\zeta_2,\zeta_{22},\zeta_{12}^{(2)})\otimes S^0(\zeta_1,\zeta_{11}).
\end{align*}
\end{prop}
The free shuffle algebra $S(\zeta_1,\zeta_{11},\zeta_{12}^{(1)})$ is a polynomial
algebra over $S^0(\zeta_1,\zeta_{11},\zeta_{12}^{(1)})$ of the variable $\zeta_1$ as a
shuffle algebra \cite{R}:
\begin{align*}
S(\zeta_1,\zeta_{11},\zeta_{12}^{(1)}) \, \cong \,
S^0(\zeta_1,\zeta_{11},\zeta_{12}^{(1)})[\zeta_1].
\end{align*}
Likewise, we have
\begin{align*}
S(\zeta_2,\zeta_{22}) \, \cong \, S^0(\zeta_2,\zeta_{22})[\zeta_2]
\end{align*}
as a shuffle algebra. Applying these isomorphisms to Proposition \ref{prop:b0}, we
have
\begin{prop}\label{prop:structure_b}
The reduced bar algebra ${\mathcal B}$ is a polynomial algebra over ${\mathcal B}^0$ of the variables
$\zeta_1, \zeta_2$ as a shuffle algebra:
\begin{align*}
{\mathcal B} \, \cong \, {\mathcal B}^0 [\zeta_1, \zeta_2].
\end{align*}
\end{prop}
Assume that $0<|z_1|,|z_2|<1$ and
define the following two contours $C_{1 \otimes 2},\; C_{2 \otimes 1}$:
\setlength{\unitlength}{1.5cm}
\begin{picture}(0,5.4)(-0.2,-0.6)
\put(0,0){\scalebox{0.75}{\includegraphics{m05.eps}}}
\put(-0.5,0){$(0,0)$}
\put(3.1,0){$(1,0)$}
\put(-0.5,3.1){$(0,1)$}
\put(3.1,3.1){$(1,1)$}
\put(4.4,0.3){$z_1$}
\put(0.3,4.4){$z_2$}
\put(2.2,2.65){$(z_1,z_2)$}
\put(1.3,0.6){$C_{2\otimes1}^{(1)}$}
\put(1.9,1.3){$C_{2\otimes1}^{(2)}$}
\put(0.6,1.3){$C_{1\otimes2}^{(2)}$}
\put(1.3,2.2){$C_{1\otimes2}^{(1)}$}
\put(2.5,0.8){$C_{2\otimes 1}=C_{2\otimes1}^{(2)}\circ C_{2\otimes1}^{(1)}$}
\put(-0.4,2.6){$C_{1\otimes 2}=C_{1\otimes2}^{(1)}\circ C_{1\otimes2}^{(2)}$}
\thicklines
\put(0.55,0.5){\vector(1,0){2}}
\put(2.55,0.5){\vector(0,1){2}}
\thinlines
\put(0.55,0.5){\vector(0,1){2}}
\put(0.55,2.5){\vector(1,0){2}}
\put(4.5,2.8){\begin{minipage}{6.8cm}
$C_{1\otimes 2}=C_{1\otimes2}^{(1)}\circ C_{1\otimes2}^{(2)}$,\\ $\qquad C_{1\otimes2}^{(2)}\!:(0,0)
\to(0,z_2)$,\\ $\qquad C_{1\otimes2}^{(1)}:(0,z_2)\to(z_1,z_2)$.\\
$C_{2\otimes 1}=C_{2\otimes1}^{(2)}\circ C_{2\otimes1}^{(1)}$,\\
$\qquad C_{2\otimes1}^{(1)}:(0,0)\to(z_1,0)$,\\ $
\qquad C_{2\otimes1}^{(2)}:(z_1,0)\to(z_1,z_2) $.\\
\end{minipage}}
\end{picture}
\setlength{\unitlength}{1cm}
\noindent
The composition of paths $C \circ C'$ is defined by connecting $C$ after $C'$.
For $\psi_1 \otimes \psi_2 \in
S^0(\zeta_1,\zeta_{11},\zeta_{12}^{(1)})\otimes S^0(\zeta_2,\zeta_{22}) $, we set
\begin{align*}
\int_{C_{1\otimes 2}}\psi_1 \otimes \psi_2 := \int_{z_1=0}^{z_1}\psi_1 \int_{z_2=0}^{z_2}\psi_2
\end{align*}
and for $\psi_1 \otimes \psi_2 \in
S^0(\zeta_2,\zeta_{22},\zeta_{12}^{(2)})\otimes S^0(\zeta_1,\zeta_{11})$,
\begin{align*}
\int_{C_{2 \otimes 1}}\psi_1 \otimes \psi_2 := \int_{z_2=0}^{z_2}\psi_1 \int_{z_1=0}^{z_1}\psi_2.
\end{align*}
Since the map $\iota_{1 \otimes 2}$ (resp. $\iota_{2 \otimes 1}$) picks up the terms of ${\mathcal B}^0$
whose iterated integral along $C_{1 \otimes 2}$ (resp. $C_{2 \otimes 1}$) does not vanish,
we have
\begin{align*}
\int_{(0,0)}^{(z_1,z_2)} \!\!\varphi&=\int_{C_{1 \otimes 2}} \!\! \varphi
=\int_{C_{1 \otimes 2}} \!\!\iota_{1 \otimes 2}(\varphi) \\
&=\int_{C_{2 \otimes 1}} \!\! \varphi
=\int_{C_{2 \otimes 1}} \!\!\iota_{2 \otimes 1}(\varphi)
\end{align*}
for $\varphi \in {\mathcal B}^0$.
\subsection{The fundamental solution of \eqref{2KZeq}}
We consider the fundamental solution ${\mathcal L}(z_1,z_2)$ of \eqref{2KZeq} normalized at
the origin $(z_1,z_2)=(0,0)$. It is a solution satisfying the following conditions:
\begin{align*}
{\mathcal L}(z_1,z_2)={\hat{\mathcal L}}(z_1,z_2) z_1^{Z_1}z_2^{Z_2}
\end{align*}
where
\begin{align*}
{\hat{\mathcal L}}(z_1,z_2)=\sum_{s=0}^{\infty}{\hat{\mathcal L}}_s(z_1,z_2), \quad {\hat{\mathcal L}}_s(z_1,z_2)\in{\mathcal U}_s({\mathfrak X}), \quad
{\hat{\mathcal L}}_s(0,0)=0 \ (s > 0),
\end{align*}
and ${\hat{\mathcal L}}_0(z_1,z_2)={\mathbf I}$. We put
\begin{align*}
\varOmega_0 &=\zeta_1 Z_1+\zeta_2 Z_2, \\
\varOmega' &=\varOmega-\varOmega_0=\zeta_{11} Z_{11}+\zeta_{22} Z_{22}+\zeta_{12} Z_{12}.
\end{align*}
It is easy to see that ${\hat{\mathcal L}}_s(z_1,z_2)$ satisfies the following recursive equation:
\begin{align*}
d{\hat{\mathcal L}}_{s+1}(z_1,z_2)=[\varOmega_0,{\hat{\mathcal L}}_s(z_1,z_2)]+\varOmega'{\hat{\mathcal L}}_s(z_1,z_2).
\end{align*}
Hence we have
\begin{align}\tag{IISOL} \label{iisol}
{\hat{\mathcal L}}_s(z_1,z_2)=\int_{(0,0)}^{(z_1,z_2)} \left(\operatorname{ad}(\varOmega_0)+\mu(\varOmega')\right)^s
({\mathbf 1} \otimes {\mathbf I}).
\end{align}
Here we use the following convention of notations:
\begin{align*}
\operatorname{ad}(\omega \otimes X)(\varphi\otimes F)&=(\omega\circ\varphi)\otimes \operatorname{ad}(X)(F),\\
\mu(\omega \otimes X)(\varphi\otimes F)&=(\omega\circ\varphi)\otimes \mu(X)(F)
\end{align*}
for $\varphi \otimes F \in S(A) \otimes {\mathcal U}({\mathfrak X}),\; \omega \otimes X \in {\mathcal B}_1 \otimes {\mathfrak X}$.
This says that \textbf{the fundamental solution normalized at $(z_1,z_2)=(0,0)$
exists and is unique}. Moreover we can show that
\begin{align*}\tag{IIFORM} \label{iiform}
\Big(\operatorname{ad}(\varOmega_0)+\mu(\varOmega')\Big)^s({\mathbf 1} \otimes {\mathbf I}) \
\in {\mathcal B}^0 \otimes {\mathcal U}_s({\mathfrak X}).
\end{align*}
\section{Decomposition theorem and hyperlogarithms}
\subsection{The decomposition theorem of the normalized fundamental solution}
We consider the following four formal (generalized) 1KZ equation.
In the following $d_{z_1}$ (resp. $d_{z_2}$) stands for the exterior differentiation
by the variable $z_1$ (resp. $z_2$):
\begin{alignat*}{4}
d_{z_1}G(z_1,z_2)&=\varOmega_{1\otimes2}^{(1)} G(z_1,z_2), \quad
&\varOmega_{1\otimes2}^{(1)}&=\zeta_1 Z_1+\zeta_{11} Z_{11}+\zeta_{12}^{(1)} Z_{12}, \\
d_{z_2}G(z_2)&=\varOmega_{1\otimes2}^{(2)} G(z_2), \quad
&\varOmega_{1\otimes2}^{(2)}&=\zeta_2 Z_2+\zeta_{22} Z_{22},\\
d_{z_2}G(z_1,z_2)&=\varOmega_{2\otimes1}^{(2)}G(z_1,z_2), \quad
&\varOmega_{2\otimes1}^{(2)}&=\zeta_2 Z_2+\zeta_{22} Z_{22}+\zeta_{12}^{(2)} Z_{12},\\
d_{z_1}G(z_1)&=\varOmega_{2\otimes1}^{(1)} G(z_1), \quad
&\varOmega_{2\otimes1}^{(1)}&=\zeta_1 Z_1+\zeta_{11} Z_{11}.
\end{alignat*}
The fundamental solution normalized at the origin to each equation satisfies the conditions
\begin{align*}
& {\mathcal L}_{i_1 \otimes i_2}^{(i_k)} = {\hat{\mathcal L}}_{i_1 \otimes i_2}^{(i_k)} \, z_{i_k}^{Z_{i_k}},\\
{\hat{\mathcal L}}_{i_1 \otimes i_2}^{(i_k)} =\sum_{s=0}^{\infty} {\hat{\mathcal L}}_{i_1 \otimes i_2,s}^{(i_k)},
\quad & {\hat{\mathcal L}}_{i_1\otimes i_2,s}^{(i_k)}\Big|_{z_{i_k}=0}=0 \quad (s > 0), \quad
{\hat{\mathcal L}}_{i_1\otimes i_2,0}^{(i_k)}= {\mathbf I}.
\end{align*}
\begin{prop}\label{prop:decomposition}
\begin{enumerate}
\item The fundamental solution ${\mathcal L}(z_1,z_2)$ of \eqref{2KZeq} normalized at the origin
decomposes to product of the normalized fundamental solutions of the (generalized)
formal 1KZ equations as follows:
\begin{align*}
{\mathcal L}(z_1,z_2)&={\mathcal L}_{1 \otimes 2}^{(1)}{\mathcal L}_{1 \otimes 2}^{(2)}
= {\hat{\mathcal L}}_{1 \otimes 2}^{(1)}{\hat{\mathcal L}}_{1 \otimes2 }^{(2)}z_1^{Z_1}z_2^{Z_2} \\
&= {\mathcal L}_{2 \otimes 1}^{(2)}{\mathcal L}_{2 \otimes 1}^{(1)}
= {\hat{\mathcal L}}_{2 \otimes 1}^{(2)}{\hat{\mathcal L}}_{2 \otimes 1}^{(1)}z_1^{Z_1}z_2^{Z_2}.
\end{align*}
\item If the decomposition
\begin{align*}
{\mathcal L}(z_1,z_2)=G_{i_1\otimes i_2}^{(i_1)}G_{i_1\otimes i_2}^{(i_2)}
\end{align*}
holds, where $G_{i_1\otimes i_2}^{(i_k)}=\hat{G}_{i_1\otimes i_2}^{(i_k)} \, z_{i_k}^{Z_{i_k}}$
satisfies the same conditions as ${\mathcal L}_{i_1 \otimes i_2}^{(i_k)}$ does, we have
$G_{i_1\otimes i_2}^{(i_k)}={\mathcal L}_{i_1 \otimes i_2}^{(i_k)}$.
\end{enumerate}
\end{prop}
\subsection{The iterated integral solution along the contours $C_{1 \otimes 2}$ and
$C_{2 \otimes 1}$}
From \eqref{iiform}, we can choose $C_{1 \otimes 2}$ as the integral contour
in \eqref{iisol}. Hence we have
\begin{align*}
{\hat{\mathcal L}}_s(z_1,z_2)&=\int_{C_{1 \otimes 2}}
\left(\operatorname{ad}(\varOmega_0)+\mu(\varOmega')\right)^s({\mathbf 1} \otimes {\mathbf I})\\
&=\int_{C_{1 \otimes 2}} (\iota_{1 \otimes 2}\otimes \operatorname{id}_{{\mathcal U}({\mathfrak X})})
\left(\left(\operatorname{ad}(\varOmega_0)+\mu(\varOmega')\right)^s({\mathbf 1} \otimes {\mathbf I})\right) \\
&=\sum_{s'+s''=s} \ \sum_{W',W''}
\int_0^{z_1} \theta^{(1)}_{1 \otimes 2}(W')
\int_0^{z_2} \ \theta^{(2)}_{1 \otimes 2}(W'') \ \alpha(W')\alpha(W'')({\mathbf I}).
\end{align*}
Here $W'$ runs over ${\mathcal W}_{s'}^0(Z_1,Z_{11},Z_{12})$,
$W''$ runs over ${\mathcal W}_{s''}^0(Z_2,Z_{22})$. \
(${\mathcal W}^0_s({\mathfrak A})={\mathcal W}^0({\mathfrak A}) \cap {\mathcal U}_s({\mathfrak X})$, and ${\mathcal W}^0({\mathfrak A})$
stands for the set of words of the letters ${\mathfrak A}$ which do not end with $Z_1,Z_2$.)
$\alpha: {\mathcal U}({\mathfrak X}) \to \operatorname{End}({\mathcal U}({\mathfrak X}))$ is an algebra homomorphism
\begin{align*}
\alpha: (Z_1,Z_{11},Z_2,Z_{22},Z_{12}) \mapsto
(\operatorname{ad}(Z_1),\mu(Z_{11}),\operatorname{ad}(Z_2),\mu(Z_{22}),\mu(Z_{12})),
\end{align*}
and $\theta^{(1)}_{1 \otimes 2}:{\mathcal U}({\mathbf C}\{Z_1,Z_{11},Z_{12}\}) \to S(\zeta_1,\zeta_{11},\zeta_{12}^{(1)})$ and $\theta^{(2)}_{1 \otimes 2}:{\mathcal U}({\mathbf C}\{Z_2,Z_{22}\}) \to S(\zeta_2,\zeta_{22})$ are linear maps
defined by replacing
\begin{align*}
\theta^{(i)}_{1 \otimes 2}(Z_i)=\zeta_i, \ \theta^{(i)}_{1 \otimes 2}(Z_{ii})=\zeta_{ii} \ \ (i=1,2),
\ \theta^{(1)}_{1 \otimes 2}(Z_{12})=\zeta_{12}^{(1)}.
\end{align*}
\noindent
In the same way, we have
\begin{align*}
{\hat{\mathcal L}}_s(z_1,z_2) &=\int_{C_{2 \otimes 1}}
\left(\operatorname{ad}(\varOmega_0)+\mu(\varOmega')\right)^s({\mathbf 1} \otimes {\mathbf I})\\
& =\int_{C_{2 \otimes 1}} (\iota_{2 \otimes 1}\otimes \operatorname{id}_{{\mathcal U}({\mathfrak X})})
\left(\left(\operatorname{ad}(\varOmega_0)+\mu(\varOmega')\right)^s({\mathbf 1} \otimes {\mathbf I})\right)\\
&=\sum_{s'+s''=s} \ \sum_{W',W''}
\int_0^{z_2} \theta^{(2)}_{2 \otimes 1}(W')
\int_0^{z_1} \ \theta^{(1)}_{2 \otimes 1}(W'') \ \alpha(W')\alpha(W'')({\mathbf I}).
\end{align*}
Here $W'$ runs over ${\mathcal W}_{s'}^0(Z_2,Z_{22},Z_{12})$, and $W''$ runs over
${\mathcal W}_{s''}^0(Z_1,Z_{11})$.
$\theta^{(2)}_{2 \otimes 1}:{\mathcal U}({\mathbf C}\{(Z_2,Z_{22},Z_{12}\}) \to S(\zeta_2,\zeta_{22},\zeta_{12}^{(2)})$ and $\theta^{(1)}_{2 \otimes 1}:{\mathcal U}({\mathbf C}\{Z_1,Z_{11}\}) \to S(\zeta_1,\zeta_{11})$ are linear maps
defined by replacing
\begin{align*}
\theta^{(i)}_{2 \otimes 1}(Z_i)=\zeta_i, \ \theta^{(i)}_{2 \otimes 1}(Z_{ii})=\zeta_{ii} \ \ (i=1,2),
\ \theta^{(2)}_{2 \otimes 1}(Z_{12})=\zeta_{12}^{(2)}.
\end{align*}
Since $[Z_1, Z_2]=[Z_1,Z_{22}]=0$, we have
\begin{align*}
{\hat{\mathcal L}}(z_1,z_2)= \left( \sum_{W'}\int_0^{z_1} \theta^{(1)}_{1 \otimes 2}(W')\alpha(W') ({\mathbf I}) \right)
\, \left( \sum_{W''}\int_0^{z_2} \theta^{(2)}_{1 \otimes 2}(W'')\alpha(W'') ({\mathbf I}) \right).
\end{align*}
This says that \textbf{each decomposition
in Proposition \ref{prop:decomposition} corresponds to
the choice of the integral contours} $C_{1 \otimes 2}, \, C_{2 \otimes 1}$.
\subsection{Hyperlogarithms of the type ${\mathcal M}_{0,5}$}
In \eqref{hlog}, let $m=2, a_1=1, a_2=z_2$, replace $\xi_0, \xi_1, \xi_2$ by $\zeta_1, \zeta_{11},
\zeta_{12}^{(1)}$ respectively, and put $\zeta(a_i)=\xi_i \ (i=1,2)$.
Then \eqref{hlog} reads as
\begin{align*}
L({}^{k_1} a_{i_1} \cdots{}^{k_r} a_{i_r} \, ;z_1)
&=\int_0^{z_1} \zeta_1^{k_1-1}\circ \zeta(a_{i_1}) \circ
\zeta_1^{k_2-1}\circ \zeta(a_{i_2})
\circ \cdots \circ \zeta_1^{k_r-1} \circ \zeta(a_{i_r}) \\
& = \sum_{n_1>n_2>\cdots>n_r>0} \ \frac{a_{i_1}^{n_1-n_2} a_{i_2}^{n_2-n_3}
\cdots a_{i_r}^{n_r}}{n_1^{k_1}\cdots n_r^{k_r}} z_1^{n_1},
\end{align*}
which is referred to as a \textbf{hyperlogarithm of the type ${\mathcal M}_{0,5}$}.
If $a_{i_1}=\cdots=a_{i_r}=1$, it is a multiple polylogarithm of one variable \eqref{1mpl}
\begin{align*}
\operatorname{Li}_{k_1,\ldots,k_{r}}(z_1)=L({}^{k_1}1\cdots{}^{k_r}1;z_1),
\end{align*}
and
\begin{align}\tag{2MPL} \label{2mpl}
\operatorname{Li}_{k_1,\ldots,k_{i+j}}(i,j;z_1,z_2)
:= L({}^{k_1}1\cdots{}^{k_i}1{}^{k_{i+1}}z_2 \cdots {}^{k_{i+j}}z_2;z_1)
\end{align}
is called a \textbf{multiple polylogarithm of two variables}. They constitute a subclass of
hyperlogarithms of the type ${\mathcal M}_{0,5}$.
We should note that, in the previous subsection, the iterated integral
\begin{align*}
L\big( \theta^{(1)}_{1 \otimes 2}(W') \,;\, z_1 \big) :=
\int_0^{z_1} \theta^{(1)}_{1 \otimes 2}(W') \quad (W' \in {\mathcal W}_{s'}^0(Z_1,Z_{11},Z_{12}))
\end{align*}
is a hyperlogarithm of the type ${\mathcal M}_{0,5}$,
and the iterated integral
\begin{align*}
L\big( \theta^{(2)}_{1 \otimes 2}(W'') \,;\, z_2 \big) :=
\int_0^{z_2} \ \theta^{(2)}_{1 \otimes 2}(W'') \quad ( W''\in {\mathcal W}_{s''}^0(Z_2,Z_{22}))
\end{align*}
is a multiple polylogarithm of one variable. Thus, \textbf{the normalized fundamental solution
${\mathcal L}(z_1,z_2)$ is a generating function of hyperlogarithms of the type ${\mathcal M}_{0,5}$}.
\section{Relations of multiple polylogarithms}
\subsection{Generalized harmonic product relations of hyperlogarithms}
From Proposition \ref{prop:b0}, one can define
\begin{align*}
\varphi(W',W'')=\iota_{1\otimes2}^{-1}(\theta_{1\otimes2}^{(1)}(W')
\otimes\theta_{1\otimes2}^{(2)}(W'')) \in {\mathcal B}^0
\end{align*}
for $W' \in {\mathcal W}^0(Z_1,Z_{11},Z_{12}), \ W'' \in {\mathcal W}^0(Z_2,Z_{22})$.
Then we have
\begin{align*}
\int_{C_{1\otimes2}}\!\!\iota_{1\otimes2}(\varphi(W',W''))
=L(\theta_{1\otimes 2}^{(1)}(W');z_1)L(\theta_{1\otimes 2}^{(2)}(W'');z_2),
\end{align*}
and
\begin{align*}
{\hat{\mathcal L}}_s(z_1,z_2)=\sum_{s'+s''=s}\
\sum_{\substack{W' \in {\mathcal W}^0_{s'}(Z_1,Z_{11},Z_{12})\\
W'' \in {\mathcal W}^0_{s''}(Z_2,Z_{22})}}
\int_{(0,0)}^{(z_1,z_2)} \ \varphi(W',W'') \ \alpha(W')\alpha(W'')({\mathbf I}).
\end{align*}
\noindent
Since $\{\alpha(W')\alpha(W'')({\mathbf I}) \,|\, W' \in {\mathcal W}^0(Z_1,Z_{11},Z_{12}), \
W'' \in {\mathcal W}^0(Z_2,Z_{22})\}$ is a linearly independent set, we obtain the following proposition:
\begin{prop}
We have
\begin{align}\tag{GHPR} \label{ghpr}
L(\theta^{(1)}_{1\otimes 2}(W');z_1)L(\theta^{(2)}_{1\otimes 2}(W'');z_2)
=\int_{C_{2\otimes 1}} \ \iota_{2\otimes 1}(\varphi(W',W''))
\end{align}
for $W' \in {\mathcal W}^0(Z_1,Z_{1 1},Z_{12}), W'' \in {\mathcal W}^0(Z_2,Z_{22})$.
\end{prop}
We call \eqref{ghpr} the \textbf{generalized harmonic product relations of hyperlogarithms}.
\begin{rem}
We have actually
\begin{align*}
\left(\operatorname{ad}(\varOmega_0)+\mu(\varOmega')\right)^s({\mathbf 1} \otimes {\mathbf I})
=\sum_{s'+s''=s} \ \sum_{W',W''} \
\varphi(W',W'') \otimes \alpha(W')\alpha(W'')({\mathbf I}).
\end{align*}
For the proof, see \cite{OU1}.
\end{rem}
\subsection{Harmonic product of multiple polylogarithms}
For $W'=Z_1^{k_1-1}Z_{11}\cdots Z_1^{k_i-1}Z_{11} Z_1^{k_{i+1}-1}Z_{12}
\cdots Z_1^{k_{i+j}-1}Z_{12}, \ W''={\mathbf I}$, we have
\begin{align*}
\int_{C_{1 \otimes 2}}\varphi(W',{\mathbf I})=\operatorname{Li}_{k_1,\ldots,k_{i+j}}(i,j;z_1,z_2).
\end{align*}
Hence \eqref{ghpr} for this case reads as
\begin{align*}
\operatorname{Li}_{k_1,\ldots,k_{i+j}}(i,j;z_1,z_2)=\int_{C_{2 \otimes 1}} \varphi(W',{\mathbf I}).
\end{align*}
Moreover, by induction, one can prove that
\textbf{the generalized harmonic product relations properly contain
the harmonic product of multiple polylogarithms} such as \eqref{hpmpl}.
Taking the limit, we have harmonic product of multiple zeta values.
Thus \textbf{we can interpret the harmonic product of multiple zeta values
as a connection problem for the formal KZ equation} such as \eqref{hpmzv}.
\section{The five term relation for the dilogarithm}
We define the action of ${\mathfrak S}_n$ on ${\mathcal M}_{0,n}$ by $\sigma(x_i)=x_{\sigma(i)}$.
For $n=5$, the action of $\sigma=(23)(45) \in {\mathfrak S}_5 $ is given, in the cubic coordinates,
by a birational transformation on ${\mathbf P}^1\times{\mathbf P}^1$ such as
\begin{align*}
\sigma(z_1,z_2) = \left(\frac{-z_1(1-z_2)}{1-z_1}, \, \frac{-z_2(1-z_1)}{1-z_2} \right).
\end{align*}
It satisfies $\sigma^2=\operatorname{id}$ and preserves the divisors $D({\mathcal M}^{cubic}_{0,5})$.
Let $\sigma^* : {\mathcal B} \to {\mathcal B}$ be the pull back induced by $\sigma$,
\begin{align*}
\begin{array}{l}
\sigma^*\zeta_1=\zeta_1+\zeta_{11}-\zeta_{22},
\quad \sigma^*\zeta_{11}=-\zeta_{11}+\zeta_{12}, \\
\sigma^*\zeta_2=-\zeta_{11}+\zeta_2+\zeta_{22},
\quad \sigma^*\zeta_{22}=-\zeta_{22}+\zeta_{12}, \quad \sigma^*\zeta_{12}=\zeta_{12}.
\end{array}
\end{align*}
and define an automorphism $\sigma_* : {\mathcal U}({\mathfrak X}) \to {\mathcal U}({\mathfrak X})$ by
\begin{align*}
(\sigma^* \otimes \operatorname{id}) \varOmega = (\operatorname{id} \otimes \sigma_*) \varOmega.
\end{align*}
Hence we have
\begin{align*}
\begin{array}{l}
\sigma_*Z_1 = Z_1, \quad \sigma_*Z_{11} = Z_1 - Z_{11} -Z_2, \\
\sigma_*Z_2=Z_2, \quad \sigma_*Z_{22} = - Z_1 + Z_2 - Z_{22},
\quad \sigma_*Z_{12} = Z_{11} + Z_{22} + Z_{12}.
\end{array} \notag
\end{align*}
Since $(\operatorname{id} \otimes \sigma_*)^{-1} (\sigma^* \otimes \operatorname{id}) \varOmega =
(\sigma^* \otimes \sigma_*^{-1}) \varOmega = \varOmega$, the function
\begin{align*}
{\tilde{\mathcal L}}(z,w)=(\sigma^* \otimes \sigma_*^{-1}){\mathcal L}(z_1,z_2)={\mathcal L}(\sigma(z_1,z_2))
\Big|_{Z \to \sigma_*^{-1}Z,\ (Z=Z_1,Z_{11},Z_2,Z_{22},Z_{12})}
\end{align*}
is also a fundamental solution of the KZ equation of two variables
which has the asymptotic behavior
\begin{align*}
{\tilde{\mathcal L}}(z_1,z_2) \sim {\mathbf I} \left(\frac{-z_1(1-z_2)}{1-z_1}\right)^{Z_1}
\left(\frac{-z_2(1-z_1)}{1-z_2}\right)^{Z_2} \qquad (z_1,z_2) \to (0,0).
\end{align*}
Therefore the \textbf{connection formula} for ${\mathcal L}(z_1,z_2)$ and ${\tilde{\mathcal L}}(z_1,z_2)$ is written as
\begin{align*}
{\tilde{\mathcal L}}(z_1,z_2) &= \mathcal{L} (z_1,z_2) \exp(-\mathrm{sgn}(\mathrm{Im} z_1) \, \pi i Z_1)
\, \exp(-\mathrm{sgn}(\mathrm{Im} z_2)\, \pi i Z_2).
\end{align*}
For the later use, it is convenient to rewrite this as follows:
\begin{align*}
(\sigma^*{\mathcal L})(z_1,z_2) &= (\sigma_*{\mathcal L})(z_1,z_2)
\exp(-\mathrm{sgn}(\mathrm{Im} z_1)\, \pi i Z_1)
\, \exp(-\mathrm{sgn}(\mathrm{Im} z_2)\, \pi i Z_2).
\end{align*}
\noindent
The terms $[Z_1,Z_{11}]$ and $[Z_2,Z_{22}]$ in the both sides above
appear in $\sigma^*{\hat{\mathcal L}}_2(z_1,z_2)$ and $\sigma_*{\hat{\mathcal L}}_2(z_1,z_2)$.
Comparing the coefficients of $[Z_1,Z_{11}]$, we have
\begin{align}
\operatorname{Li}_2\left(\frac{-z_1(1-z_2)}{1-z_1}\right)
= \operatorname{Li}_{1,1}(1,1;z_1,z_2)-\operatorname{Li}_2(z_1)-\operatorname{Li}_{1,1}(z_1)+\operatorname{Li}_2(0,1;z_1,z_2),
\tag{L1} \label{landen1}
\end{align}
and comparing the coefficients of $[Z_2,Z_{22}]$,
\begin{equation}
\operatorname{Li}_2\left(\frac{-z_2(1-z_1)}{1-z_2}\right)
= -\operatorname{Li}_{1,1}(1,1;z_1,z_2)-\operatorname{Li}_2(z_2)-\operatorname{Li}_{1,1}(z_2)+\operatorname{Li}_1(z_2)\operatorname{Li}_1(z_1).
\tag{L2}\label{landen2}
\end{equation}
We should observe that \eqref{landen1} is regarded as a \textbf{``two-variables''
analogue of the Landen formula for the dilogarithm} \cite{Le}.
Since $\operatorname{Li}_2(0,1;z_1,z_2)=\operatorname{Li}_2(z_1z_2)$ and
$\operatorname{Li}_{1,1}(z)=\frac{1}{2}\log^2(1-z)$, $\eqref{landen1} + \eqref{landen2}$
gives the \textbf{five term relation for the dilogarithm} \eqref{5term}:
\begin{align*}
\operatorname{Li}_2(z_1z_2)= \operatorname{Li}_2 & \left(\frac{-z_1(1-z_2)}{1-z_1}\right) +
\operatorname{Li}_2\left(\frac{-z_2(1-z_1)}{1-z_2}\right) \notag \\
& \qquad + \operatorname{Li}_2(z_1)+\operatorname{Li}_2(z_2)+\frac{1}{2}\log^2\left(\frac{1-z_1}{1-z_2}\right).
\end{align*}
\vspace{5mm}
|
\section{Introduction}
The most stringent constraints on a primordial non-Gaussianity come
from the temperature anisotropy measurement by WMAP 5-year observations \cite{k09},
which show a small effect of non-Gaussianity on large scales.
On the other hand, the probe of a positive non-Gaussianity on smaller scales
seems to have been detected recently in WMAP 3-year observations \cite{ym08}.
Extragalactic dust emission can be also associated with primordial deviations from Gaussianity
at sub-degree scales \cite{rsph09}.
From the theoretical point of view, the assumption that primordial density fluctuations
generating large scale structure formation can be modeled by a Gaussian distribution
is supported by the inflationary models. However, distinct variants of the inflationary models
incorporate a primordial non-Gaussianity, as shown by \cite{bkmr04}.
We present a series of N-body numerical simulations of large scale structure formation
in order to investigate the effect of a primordial non-Gaussianity on the
probability distribution function and the two-point correlation
function of the density fluctuation field of the dark matter distribution.
The non-Gaussianity is quantified by the usual parameter $f_{NL}$,
which modifies the gravitational potential at a primordial epoch.
Higher order modifications of the gravitational potential
have been recently considered as well \cite{ds09}.
\section{Non-Gaussian model}
For a large class of models of generation of the initial seeds
for structure formation, including single-field \cite{m03} and
multi-field inflation \cite{bu04}, or the curvaton \cite{kk06}, the non-Gaussianity
of the initial density fluctuations can be modeled
through a quadratic term in the Bardeen's gauge-invariant
potential $\Phi$, namely \cite{ks01}
\begin{equation}
\label{FNL}
\Phi (\mathbf{x}) = \Phi_{\rm L} (\mathbf{x}) + f_{\rm{NL}} \left[\Phi_{\rm L}^2 (\mathbf{x}) -
\langle\Phi_{\rm L}^2 (\mathbf{x})\rangle \right] \;,
\label{NG}
\end{equation}
where $\Phi_{\rm L}$ is a Gaussian random field,
with $\langle\Phi_{\rm L} (\mathbf{x})\rangle = 0$, and the specific value
of the dimensionless non-linearity parameter $f_{\rm{NL}}$ depends on
the assumed scenario. This form of non-Gaussianity can be obtained
from a truncated expansion of the effective inflaton potential \cite{sb90}.
From the latest WMAP5 temperature data \cite{k09}
\begin{equation}
-151 < f_{\rm{NL}} < +253 \; \rm with \; 95\% \; CL
\end{equation}
With this convention positive (negative) $f_{\rm NL}$ corresponds to
positive (negative) skewness of the probability distribution function of density fluctuations.
The Gaussianity tests show that the primordial fluctuations are Gaussian
to the 0.1$\%$ level.
Since the modification introduced in \eqref{NG} is of second order in the potential,
the linear two-point correlation function is not modified for any value of $f_{NL}$
in the chosen interval $|f_{NL}| \le 5,000$. In fact, by order of magnitude
$\Phi ({\mathbf x}) \le 10^{-5}-10^{-6}$; so even for $f_{NL} \sim 10^3$,
the non-Gaussian correction gives a contribution of the order of $1\%$.
It should be noted that distinct realizations of primordial non-Gaussianity
in the density fluctuation field has been proposed, see e.g. \cite{csz07}.
\section{N-body simulations}
The procedure of generation of primordial non-Gaussian initial condition
has been implemented in the MPI pa\-rallelized numerical code MPGRAFIC \cite{p08}.
The N-body simulations have been performed using the AMR
parallelized numerical code RAMSES-3.0 \cite{t02}. We have followed only the evolution
of the collisionless dark matter particles.
We performed simulations in computational boxes of physical sizes
$(162 h^{-1}$ Mpc)$^3$, $(300 h^{-1}$ Mpc)$^3$,
$(324 h^{-1}$ Mpc)$^3$, $(500 h^{-1}$ Mpc)$^3$
with a grid resolution of $256^3$. The series of simulations presented here
aims to detect the non-linear scales effect of the primordial non-Gaussianity.
\section{Cosmological Parameters}
We have adopted the following set of cosmological parameters, within the framework
of the so-called ``concordance'' $\Lambda$-CDM model, compatible with WMAP 5-year observations \cite{k09}:
$\Omega_b = 0.044$, $\Omega_m = 0.26$, $\Omega_\Lambda = 0.74$, $H=72$ km s$^{-1}$Mpc$^{-1}$.
The spectral index of the initial power spectrum is $n=0.951$.
The rms of density fluctuations in spheres of radius $R_0$, $\sigma^2(R_0)$, is defined as
\begin{equation}
\sigma^2(R_0) = \int d^3 \mathbf{k} P(k;z) W^2(kR_0)
\end{equation}
in absence of non-linear perturbations, i.e. $P(k;z)$ is the linear power spectrum.
Here $W^2(kR_0)$ is an appropriate window function.
At the scale $R_0 = 8 h^{-1}$ Mpc, we imposed $\sigma^2(R_0 )\sim 0.792$ \cite{k09}.
\section{Initial conditions}\label{ini}
We first briefly review the procedure of generation of 3D Gaussian initial conditions \cite{b01,p08}.
In the Gaussian case, the 3D initial density and displacement fields
are completely defined by the two-point correlation function,
or power spectrum, $P(k)$ of the density fluctuation field $\delta({\mathbf x})$,
defined as $P(k) \delta_D({\mathbf k} - {\mathbf k'}) = \langle \delta({\mathbf k}) ^*\delta({\mathbf k'}) \rangle$,
where $\delta_D({\mathbf x})$ is the Dirac delta function and
\begin{equation}
\delta({\mathbf k}) = \frac{1}{(2\pi)^3} \int d{\mathbf x} e^{-i {\mathbf k}{\mathbf x}} \delta({\mathbf x})
\label{fourier}
\end{equation}
The 3D density fluctuation field is generated as convolution
of a Gaussian white noise $n({\mathbf x})$ with the CDM transfer function $T(k) = \sqrt{P(k)}$.
In the position space, the Gaussian field $n({\mathbf x})$ is generated with the standard technique
of the Gaussian white noise, using a uniform phase
and an amplitude extracted from the Rayleigh distribution.
The field $n({\mathbf x})$ has zero average value and unitary variance, i.e. $\langle |n({\mathbf x})|^2 \rangle = 1$,
or, equivalently, constant power spectrum: $\langle |n({\mathbf k})|^2 \rangle = 1$.
As a second step, the Gaussian white noise $n({\mathbf x})$ is convolved with a CDM transfer function $T(x)$
to obtain the density fluctuation and displacement fields according to the Zel'dovich approximation \cite{z70}, i.e.
\begin{equation}
\delta({\mathbf x}) = \int d{\mathbf k} e^{i {\mathbf k}{\mathbf x}} n({\mathbf k}) T(k) , \quad
s_j ({\mathbf x}) = \int d{\mathbf k} e^{i {\mathbf k}{\mathbf x}} \frac{n({\mathbf k}) T(k)}{k^2} i k_j.
\label{fourier2}
\end{equation}
with $j=1,2,3$.
Therefore, the primordial fields $\delta({\mathbf x})$ and $s_j ({\mathbf x})$
are made up by two decoupled contributions: the field $n({\mathbf x})$
represents the statistics at early epoch, which is commonly assumed to be Gaussian,
providing the phase of the fluctuation. On the other hand,
the unknown physics of plasma processes leading to the CMB epoch
is contained in the transfer function $T(k)$,
completely determined by the two-point correlation function, for which
the parametrization of \cite{eh98,p08} is used. The $T(k)$ provides
the amplitude of the fluctuation and depends on the cosmological model.
In order to disentangle primordial non-Gaussianity
from non-Gaussianity generated by the non-linear evolution of structures
under the gravitational potential, the statistics has been modified at early times,
i.e. before the convolution with the transfer function, by using the \eqref{NG}.
Once the Gaussian white noise $n({\mathbf x})$ is generated,
the primordial density fluctuations field in the Fourier space $\delta^p({\mathbf k}) = n({\mathbf k})$ is related
through the Poisson equation to the primordial gravitational potential $\Phi^p (\mathbf{x})$:
\begin{equation}
k^2 \Phi^p (\mathbf{k}) = \frac{3 \Omega_{m_0}}{2 a r^2_{H_0}} \delta^p({\mathbf k})
\label{Poisson}
\end{equation}
where $\Omega_{m_0}$, i.e. the matter density in units of critical density,
and $r_{H_0} = c/H_0$ are evaluated at the present epoch.
The potential $\Phi^p (\mathbf{k})$ is first transformed back to the position space,
and converted in a non-Gaussian potential according to the prescription in \eqref{NG}.
The new gravitational potential $\Phi^p_{NG} (\mathbf{x})$ corresponding to a
non-Gaussian primordial statistics is converted back to the Fourier space
in which the new density fluctuation and displacement fields are computed.
The initial condition for the N-body simulation is obtained by using the new
non-Gaussian gravitational potential to perturb the initially uniform density
distribution according to the Zel'dovich approximation.
The initial redshift, as input for the code RAMSES, is $z_{start} \sim 77$.
In Fig.~\ref{PDF} the probability distribution function PDF$(\delta)$ is shown for different values of $f_{NL}$.
The PDF$(\delta)$ of the fluctuation density field $\delta$ at a given scale R is computed
by sampling the simulation box at a given redshift with a number of independent spheres of radius R,
fully covering the simulation box.
In Fig.~\ref{PDF} the PDF with comoving sphere radius R $= 8 h^{-1}$ Mpc is shown
in a simulation box of size L = 162 $h^{-1}$ Mpc and initial grid resolution 256$^3$,
with a cell grid comoving size of $\Delta L = 0.63 h^{-1}$ Mpc.
Significant discrepancies with respect to the Gaussian case
are found in the tail of the PDF only for values of $f_{NL}$
beyond the constraints inferred from WMAP 5-years \cite{k09}, at variance with \cite{g08,p09}.
The second and third order momenta of the $\delta$ distribution
have been estimated at the initial redshift $z_{start} \sim 77$, corresponding
to the initial condition existing prior the gravitational evolution under
the hydrodynamics equations. We find that at $z_{start}$, for
$f_{NL} = 5000$ and $f_{NL} = -5000$ the third order momentum is
$\delta^3 = 4.61\times 10^{-8}$ and $\delta^3 = -3.93\times 10^{-8}$, respectively.
As a check the numerical reliability of our computation,
the third momentum of the non-Gaussian density fluctuation $\delta$
can be easily estimated, as pointed out by \cite{ddhs08}.
From the Laplacian of \eqref{NG},
$\nabla^2 \Phi = \nabla^2 \Phi_{\rm L} + 2f_{\rm{NL}} \left[|\nabla\Phi_{\rm L}|^2 +
\Phi_{\rm L}\nabla^2\Phi_{\rm L} \right] $,
by using the Poisson equation, the last relation implies
$\langle \delta^3 \rangle = 6 f_{NL} \langle \delta_L^3 \Phi_L \rangle$,
at the lowest order in the product $f_{NL} \Phi_L$. This relation
has been verified with good accuracy.
\begin{figure}
\includegraphics[height=.4\textheight]{fnl_2.pdf}
\caption{PDF for different values of $f_{NL}$ at redshift $z_{start} \sim 77$.}
\label{PDF}
\end{figure}
\section{Definition of $f_{\rm NL}$}
As already pointed out by several authors, see e.g. \cite{k09,p08},
the definition of $f_{\rm NL}$ depends on the cosmic epoch
(i.e. on the value of the FRW scale factor $a$) at which \eqref{FNL}
is applied, because both the potentials Gaussian and non-Gaussian,
$\Phi_{\rm L}$ and $\Phi$ respectively, evolve in time with
the factor $g(a) = D(a)/a$, where $D(a)$ is the linear growth factor
of density fluctuations. Therefore the relation between the $f_{{NL}_\infty}$
introduced at $z \rightarrow \infty$ and $f_{{NL}_0}$ introduced at present epoch
is $f_{{NL}_\infty} = f_{{NL}_0} g(0)/g(\infty)$, where $g(\infty)/g(0) \sim 1.34$.
In the simulations presented here,
the gravitational potential is modified directly at primordial epoch $z \rightarrow \infty$,
before the convolution with the CDM transfer function $T(k)$. This choice allows
for a direct modification of the primordial statistics (see Sect. above).
Therefore, no linear extrapolation back to $z_{start}$ of the gravitational potential is performed.
The parameter $f_{NL}$ introduced in \eqref{NG} affects only the statistics of the matter distribution and
does not affect the physics pre-CMB epoch which is included
by the further convolution with the transfer function.
\section{Results}
The formalism summarized above allowed us to numerically explore
the imprint of the primordial non-Gaussianity of $f_{NL}$-type
on the dimensionless power spectrum of the CDM density field $\Delta^2 (k) = P(k)k^3/(2\pi^2)$ at $z=0$,
see left panel of Fig.~\ref{Delta}.
The computational box has a size of 162 $h^{-1}$ Mpc,
with an initial coarse graining of $256^3$ cells.
Higher resolution simulations are currently underway.
The chosen resolution allowed us to resolve the small scale contribution of the non-Gaussian perturbation to the
structure formation.
\begin{figure}
\includegraphics[height=.27\textheight]{comp_Delta2.pdf}
\includegraphics[height=.27\textheight]{pk_7new.pdf}
\caption{{\it Left panel}: Redshift evolution of the dimensionless power spectrum $\Delta^2(k)$
toward $z=0$, for different values of $f_{NL}$. {\it Right panel}:
Comparison at $z=0$ of the power spectrum of matter density $P(k)$
in the Gaussian and non-Gaussian case. We find a discrepancy of the
order of $15\%$ ($2-3\%$) at $k=0.7 (h$ Mpc$^{-1}$), for $f_{NL} = \pm 5000$ ($\pm 750$).}
\label{Delta}
\end{figure}
At $z > 5$ the discrepancy between the Gaussian and non-Gaussian
case decreases because non-linear effects due to primordial non-Gaussianity
on the scale of interest become negligible at large redshifts.
In the right panel of Fig.~\ref{Delta} the ratio of power spectra $P(k; f_{NL})/P(k; f_{NL} = 0)$ is extracted
from the simulations at redshift $z=0$. The matter power spectrum of non-Gaussian
models appears to deviate by a few per cent at $k = 0.1 h Mpc^{-1}$, with $f_{NL} = \pm5000$.
Moreover, deviations from Gaussian case are enhanced at non-linear scale.
\section{Conclusions and perspectives}
The formalism of $f_{NL}$-type allowed us to numerically explore
the imprint of the primordial non-Gaussianity
on a number of observables quantities, i.e. the initial probability distribution function
of the density fluctuation field $PDF(\delta)$ and the non-linear power spectrum
at present epoch.
The amplitude of departure of power spectrum from the Gaussian case
could be hardly detectable, also due to a progressive coalescence
of power spectra as $z$ decreases.
The detectability of third order momentum of $\delta$, regardeless the
very small amplitude, can lead to observable differences at the present epoch
in the case of initial predominance of sub-density regions.
A series of simulations with higher space resolution will be performed
to analyse the mass function and the bias of dark matter haloes within the same
model for the primordial non-Gaussianity at the leading perturbative order
in the gravitational potential, with parameter $f_{NL}$.
On the other end some simulation tests involving the
successive term in the Taylor expansion of the potential have been carried out,
with the parameter $g_{NL}$ \cite{k09}, leading to not-observable deviations
from Gaussian case in the $PDF$ and $P(k;z=0)$ up to $g_{NL}=10^6$.
\begin{theacknowledgments}
The numerical simulations have been performed with the
Horizon cluster, for which the technical assistance is gratefully
acknowledged. The work of FF was supported by CNRS fellowship.
FF acknowledges the Institut f\"ur Theoretische Physik of Ruhr-Universit\"at Bochum,
where part of this work has been done.
\end{theacknowledgments}
\bibliographystyle{aipproc}
|
\section{Introduction}
Recently some theoretical and numerical studies were published which investigate the metal-insulator transitions and the transition between the two insulator phases of the ionic Hubbard model \cite{Byczuk1,Craco1,Bouadim1,Paris1,Kancharla1}. The numerical results are obtained with the DMFT (at zero temperature) and the determinant quantum Monte Carlo method. The existence of an intermediate metallic phase between the band and the Mott insulators seems confirmed by all the authors but the nature of this phase and of the metal-insulator transitions are still under debate.
\\ In this paper, we present results on the two-dimensional ionic Hubbard model obtained with a method based on the direct-space proposed by Suzuki and al.\cite{Suzuki1,Suzuki2} and Hirsch and al.\cite{Hirsch1,Hirsch2}. This quantum Monte Carlo method is presented in references \cite{martinie1,martinie2}. At fixed temperature, this method allows to generate some of the most representative occupation number basis states of the model. These states are used to compute average values of energy, molar specific heat, occupancy, structure factor and rough static electric conductivity.
\section{Ionic Hubbard model}
The Hamiltonian of the ionic Hubbard model can be written
\begin{eqnarray}
\label{hamiltonian1}
H =&&-t\sum_{\left\langle i,j\right\rangle,\sigma }\left( c^{\dagger}_{i,\sigma}c_{j,\sigma}+hc\right) +U\sum_{i}n_{i\downarrow}n_{i\uparrow}\nonumber\\&&+ \Delta \sum_{i\in A}n_{i} -\Delta \sum_{i\in B}n_{i}
\end{eqnarray}
The square lattice is a bipartite lattice with two sublattices $ A $ and $ B $. $ c^{\dagger}_{i,\sigma} $ and $ c_{i,\sigma} $ are the fermion creation and destruction operators at the lattice site $ i $ with spin $ \sigma $. $ n_{i,\sigma}=c^{\dagger}_{i,\sigma}c_{i,\sigma} $ is the number operator. $ t $ is the hopping term between nearest-neighbor sites, $ U $ denote the on-site Coulomb repulsion, $ \Delta $ is the staggered potential between the $ A $ and $ B $ sublattices.
\section{Simulation parameters}
The square lattice contains $6\times6$ sites with periodic boundary conditions. Each elementary cell contains two sites A and two sites B. There are $ 18 $ spin up and $ 18 $ spin down (half filling). The hopping parameter $ t $ is fixed at a value $ 1 $ except for the simulations of the atomic limit where $ t=0 $. The model is decomposed in sub-systems which contain four sites. These sub-systems are grouped together in two sub-hamiltonians. The imaginary-time interval is divided into twenty slices. For each simulation five decreasing-increasing temperature cycles or decreasing-increasing interaction U cycles were programmed.There are one hundred points by curves.
\section{The atomic limit ($ t=0 $)}
Simulations were performed for the simple case of the atomic limite where $ t=0 $. For this value the Hamiltonian is diagonal, so there is no problem due to the non-commutativity, in consequence the number of slides can be one, and there is no sign problem. Cycles of increasing and decreasing values of $ U $ were programmed at the fixed temperature $ kT=0.01 $ for different values of the staggered potentiel $ \Delta $. The conductivity is always zero, so the model is an insulator whatever the values of the interactions. The Figs. \ref{fig:energyt0}, \ref{fig:chalspect0} and \ref{fig:occupanciest0} display the results. As it is expected, one remarks that the electronic transition happens for $ U_{c}=2\Delta $ without hysteresis phenomenon. There is one spin by site for $ U\gtrsim2\Delta $ whereas only the sites of the sublattices $ B $ are occupied for $ U\lesssim2\Delta $. In this domain the energy of the model is $ E\simeq18\left( U-2\Delta\right) $, while it is zero for $ U\gtrsim2\Delta $. There is not magnetic order.
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-1.eps}
\end{center}
\caption{(Color online). The energy at $ kT=0.01 $ for different values of $ \Delta $.}
\label{fig:energyt0}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=7cm]{figure-2.eps}
\end{center}
\caption{(Color online). The molar specific heat at $ kT=0.01 $ for different values of $ \Delta $.}
\label{fig:chalspect0}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-3.eps}
\end{center}
\caption{(Color online). The site occupancies at $ kT=0.01 $ for $ \Delta=0.5 $, $ \Delta=2 $ and $ \Delta=4 $.The solid lines corespond to sites A and the dashed lines correspond to sites B.}
\label{fig:occupanciest0}
\end{figure}
The Figs. \ref{fig:energy-T-t0}, \ref{fig:chalspec-T-t0} and \ref{fig:comparoccupation-T-t0} show the influence of the temperature on the energy curves, the molar specific heat curves and the occupancies curves versus interaction $ U $. The occupancies curves for the different temperatures have very little error bars and cross almost exactly at $ U=2 \Delta $. This is in good agreement with the zero value of the specific heat for this value of $ U $. At this point the site A occupancy is near $ 0.66 $ while the site B occupancy is about $ 1.32 $. In this special state $ 12 $ sites A are each occupied by one spin, $ 12 $ sites B are occupied equally by one spin and $ 6 $ sites B are occupied by two spins. The energy of this state is $ E=6U-12 \Delta=0 $. Indeed, the three energy curves in Fig. \ref{fig:energy-T-t0} pass through the same point $ \left( U=1, E=0\right) $ so $ \partial E/\partial T =0 $.\\
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-4.eps}
\end{center}
\caption{(Color online). Energy in the atomic limit $ \left( t=0, \Delta=0.5 \right) $ for different temperatures.}
\label{fig:energy-T-t0}
\end{figure}
One remarks that all the specific heat curves of the Fig. \ref{fig:chalspec-T-t0} match exactly for the abscisse $ U/kT $. The specific heat curves for four sizes of the model, at half filling, are shown in Fig. \ref{fig:chalspec-N-t0}. One remarks that these curves are similar.
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-5.eps}
\end{center}
\caption{(Color online). Molar specific heat in the atomic limit $ \left( t=0, \Delta=0.5 \right) $ for different temperatures.}
\label{fig:chalspec-T-t0}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-6.eps}
\end{center}
\caption{(Color online). Site occupancies in the atomic limit $ \left( t=0, \Delta=0.5 \right) $ for different temperatures. Solid line (black) correspond to $ kT=0.005 $, the dashed line (red) correspond to $ kT=0.01 $, the dotted line (green) correspond to $ kT=0.02 $ and the dot-dash line (blue) correspond to $ kT=0.05 $.}
\label{fig:comparoccupation-T-t0}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-7.eps}
\end{center}
\caption{(Color online). Molar specific heat in the atomic limit $ \left( t=0, \Delta=0.5 \right) $ for four model sizes.}
\label{fig:chalspec-N-t0}
\end{figure}
\section{Model with hopping interaction $ \left( t=1 \right) $}
\subsection{Simulations at $ T=\texttt{constant} $}
The Figs. \ref{fig:conductivity-hight-U}, \ref{fig:conductivity-low-U}, \ref{fig:factantiferro-U} and \ref{fig:paire-U} show the DC conductivity, the structure factor and the double occupancy curves for different values of the staggered potential $ \Delta $ at $ kT=0.01 $.
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-8.eps}
\end{center}
\caption{(Color online). Conductivity for low and hight values of $ U $ and different values of $ \Delta $. $ \left( t=1, kT=0.01 \right) $ .}
\label{fig:conductivity-hight-U}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-9.eps}
\end{center}
\caption{(Color online). DC Conductivity for low values of $ U $ and different values of $ \Delta $. $ \left( t=1, kT=0.01 \right) $ .}
\label{fig:conductivity-low-U}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-10.eps}
\end{center}
\caption{(Color online). Structure factor for different values of $ \Delta $. $ \left( t=1, kT=0.01 \right) $ .}
\label{fig:factantiferro-U}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-11.eps}
\end{center}
\caption{(Color online). Double occupancy for different values of $ \Delta $. $ \left( t=1, kT=0.01 \right) $ .}
\label{fig:paire-U}
\end{figure}
For $ \Delta \gtrsim 1 $ the model undergoes a transition between two insulator states. During the transition the system becomes conductor. In the low $ U $ insulator state, only the $ B $ sites are occupied by two spins. It is a band insulator (BI). All the sites are occupied by one spin in the hight $ U $ insulator state. This last insulator state presents an antiferromagnetic structure, it is a Mott insulator (MI). This transition between the two insulator states with an intermediate metallic phase was already observed \cite{Byczuk1,Craco1,Bouadim1,Paris1,Kancharla1} with other simulation methods at $ T=0 $ and $ T\neq0 $. Our results are in good agreement with those obtained within the other methods.\\
Fig. \ref{fig:conductivity-U-MD} shows the DC conductivity for $ \Delta=1 $ and $ \Delta=2 $ for decreasing and increasing values of $ U $. An hysteresis phenomenon appears for the transition from the metal to the Mott insulator while the the curves fit for the metal-band insulator transition. This behaviour is observed for all the values of $ \Delta $. One can deduce that the MI-to-metal phase transition is a first order transition, while the BI-to-metal phase transition is continuous. This is in good agreement with the result of reference \cite{Craco1}.\\
One observes that the structure factor decreases for large $ \Delta $ while the behaviour of the double occupancy is similar for all the values of $ \Delta \gtrsim 1.0$. The caracteristcs of the BI phase ( null structure factor and double occupancy $\approx 0.5 $) are the same for all values of $ \Delta $ whereas the insulator phase induce by increasing value of $ U $ is not a purely MI. Moreover, the conductivity of this phase is not null for $ \Delta \gtrsim 5 $.
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-12.eps}
\end{center}
\caption{(Color online). DC conductivity for $ \Delta =1 $ and $ \Delta =2 $ for decreasing and increasing values of $ U $ at $ kT=0.01 $.}
\label{fig:conductivity-U-MD}
\end{figure}
\subsection{Simulations at $ U=\texttt{constant} $}
The Figs. \ref{fig:conductivity-hight-T} and \ref{fig:conductivity-low-T} show the conductivity curves at low and hight temperatures for $ \Delta =0.5 $ and different values of $ U $. One observes metallic behaviour for $ kT\gtrsim 0.1 $. At low temperature, for decreasing temperature (Fig \ref{fig:conductivity-low-T}), the system becomes insulator with behaviour change for $ U\simeq 1.75 $. This behaviour change can be observed equally on the double occupancy curves of Fig. \ref{fig:doubleoccupancy-low-T}. For $ U\gtrsim 1.75 $ the metal-insulator transition occurs with hysteresis phenomenon (Fig. \ref{fig:hysteresis-low-T}). One can deduce that this transition is a first order transition. On the contrary, for $ U\lesssim 1.75 $ the conductivity curves for increasing and decreasing temperature are similar.
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-13.eps}
\end{center}
\caption{(Color online). Conductivity curves at hight temperatures for $ \Delta=0.5 $ and different values of the coulombian repulsion $ U $.}
\label{fig:conductivity-hight-T}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-14.eps}
\end{center}
\caption{(Color online). Conductivity curves at low temperatures for $ \Delta=0.5 $ and different values of the coulombian repulsion $ U $.}
\label{fig:conductivity-low-T}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-15.eps}
\end{center}
\caption{(Color online). Double occupancy curves at low temperatures for $ \Delta=0.5 $ and different values of the coulombian repulsion $ U $.}
\label{fig:doubleoccupancy-low-T}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-16.eps}
\end{center}
\caption{(Color online). Conductivity curves at low temperatures for increasing and decreasing temperature for $ U=2.0 $, $ U=2.5 $ and $ U=3.0 $ ($ \Delta =0.5 $). One observes an hysteresis phenomenon.}
\label{fig:hysteresis-low-T}
\end{figure}
\subsection{Phase diagram}
The conductivity curves of Figs. \ref{fig:conductivity-hight-U} and \ref{fig:conductivity-low-U} can be used to drawn the phase diagram at the constant temperature $ kT=0.01 $. For large $ \Delta $ one can consider that the metallic region shrinks to a single metallic point. For each value of $ \Delta\lesssim 5.0 $, the coulombian interactions $ U_{c1} $ and $ U_{c2} $ which correspond with the metal-insulator transitions are determined at mid-height of the maximum conductivity. The phase diagram of the ionic Hubbard model is shown in Fig. \ref{fig:diagram}.\\
The behaviour change observed on the double occupancy curves of Fig. \ref{fig:doubleoccupancy-low-T} corresponds to the point $ \left( \Delta =0.5, U\simeq 1.75\right) $ in the phase diagram. This suggests that a cross-over line exists in the metallic region. This line can correspond approximatly to the set of points for which the double occupancy is $ 0.25 $. For this double occupancy value $ 9 $ sites A and $ 9 $ sites B are occupied by one spin and $ 9 $ sites B are occupied by two spins.\\
One remarks, on the phase diagram, that the transition lines for $ t=0 $ and $ t=1 $ are parallel for large $ \Delta $. The gap between these two lines is $ \triangle U \approx 2t $.
\begin{figure}
\begin{center}
\includegraphics[width=0.9\linewidth]{figure-17.eps}
\end{center}
\caption{ Phase diagram of the 2D IHM at $ kT=0.01 $. The dot-dash line is the transition line at the atomic limit ($ U_{c}=2\Delta $). The symbol plus is the point at which behaviour changes. The dotted line correspond to the set of points for which double occupancy is $ 0.25 $.}
\label{fig:diagram}
\end{figure}
\section{Conclusion}
\label{sec:conclusion}
Most results presented in references \cite{Byczuk1,Craco1,Bouadim1,Paris1,Kancharla1} are obtained for $ T=0 $ whereas, by principle, our simulation method works for not null temperature. However the results and the phase diagram are similar. Our results confirm the existence of a metallic region between Mott and band insulator phases. The natures of the metal-insulator transitions are different. The MI-metal transition is discontinuous while the BI-metal transistion is continuous like it is told in reference \cite{Craco1}. The metallic phase shrinks to a line for large coulombian interaction $ U $, but the BI-MI transition is not direct. Moreover, the insulator phase for $ U>U_{c} $ is not a purely Mott insulator phase. Studies with increasing and decreasing temperature show that there is a behaviour change in the metallic region which divides it into two regions. These two regions correspond to the precursor phases of MI and BI phases.
|
\section{Introduction}
Over the past 15 years, well over 500 brown dwarf members of the late-M, L and T dwarf spectral classes have been identified in various Galactic environments, encompassing a broad diversity in color, spectral properties, and physical characteristics (\citealt{2005ARA&A..43..195K} and references therein). Sustained effort has been made to identify the coldest of these sources, which include the low-mass extreme of star formation and primordial relics of the Galactic halo population. The most recent discoveries made with the Two Micron All Sky Survey (2MASS; \citealt{2003yCat.2246....0C,2006AJ....131.1163S}), the UKIRT Infrared Deep Sky Survey (UKIDSS; \citealt{2007MNRAS.379.1599L}) and the Canada France Hawaii Telescope Legacy Survey (CFHTLS; \citealt{2008A&A...484..469D}) have extended the known population down to and below effective temperatures {T$_{eff}$} $\approx$ 600~K
(e.g., \citealt{2007MNRAS.381.1400W,2008MNRAS.391..320B, 2009MNRAS.395.1237B, 2008ApJ...689L..53B,2008A&A...482..961D,2009ApJ...695.1517L}). This has raised the question as to where the currently coldest class of brown dwarfs---the T dwarfs---ends and the next cooler class---the Y dwarfs---might begin. Such exceedingly dim and cold sources are predicted to encompass several major chemical transitions in brown dwarf atmospheres, including the disappearance of Na and K into salt condensates, the emergence of strong {NH$_3$} absorption across the infrared band, and the formation of photospheric water ice clouds (e.g., \citealt{1999ApJ...513..879M, 1999ApJ...519..793L, 2002Icar..155..393L, 2003ApJ...596..587B}). Accordingly, there is considerable interest and controversy as to how to delineate this putative class;
see discussions by \citet{2007ApJ...667..537L,2008MNRAS.391..320B} and \citet{2008A&A...482..961D}.
A promising low-temperature brown dwarf candidate was recently identified by \citet{2010arXiv1001.4393B} and \citet{2010arXiv1001.2743S}
as a co-moving companion to
the nearby blue L6 dwarf SDSS~J141624.08+134826.7 (hereafter SDSS~J1416+1348; \citealt{2010ApJ...710...45B, 2009arXiv0912.3565S, kirkpatrick2010}). The object, {ULAS~J141623.94+134836.3} (hereafter {ULAS~J1416+1348}), was identified in UKIDSS as a faint ($J$ = 17.35$\pm$0.02) and unusually blue ($J-K = -1.58~\pm$~0.17) near-infrared source separated by 9$\farcs$8 from the L dwarf. Using astrometry from 2MASS,
UKIDSS, the Sloan Digital Sky Survey Data Release 7 (SDSS DR7; \citealt{2000AJ....120.1579Y,2009ApJS..182..543A}), and follow-up imaging, both \citet{2010arXiv1001.4393B} and
\citet{2010arXiv1001.2743S} were able to confirm common proper motion of this pair. \citet{2010arXiv1001.2743S} also determined an astrometric distance to the primary of 7.9$\pm$1.7~pc, consistent with spectrophotometric estimates from \citet[6.5--10.7~pc]{2010ApJ...710...45B} and \citet[6.4--9.6~pc]{2009arXiv0912.3565S}. At this distance, the (poorly constrained) absolute magnitudes of
{ULAS~J1416+1348}, $M_J$ = 17.8$\pm$0.5 and $M_K$ = 19.4$\pm$0.5, are
equivalent to or fainter than those of the latest-type brown dwarfs with measured distances, Wolf 940B
($M_J$ = 17.68$\pm$0.28 and $M_K$ = 18.37$\pm$0.28; \citealt{2009MNRAS.395.1237B}) and ULAS~J003402.77$-$005206.7
($M_J$ = 17.65$\pm$0.11 and $M_K$ = 17.98$\pm$0.12; \citealt{2007MNRAS.381.1400W,2009arXiv0912.3163S}).
\citet{2010arXiv1001.4393B} report a 1.0--2.5~$\micron$ spectrum of {ULAS~J1416+1348}, identifying it
as a T7.5 brown dwarf with highly suppressed $K$-band flux.
Indeed, {ULAS~J1416+1348} is the bluest T dwarf in $J-K$ color identified to date, matching the unusually blue nature
of its L dwarf companion.
{\it Spitzer} photometry reported in \citet{2010arXiv1001.4393B} further
suggest an exceptionally low-temperature ({T$_{eff}$} $\sim$ 500~K), metal-poor ([M/H] $\approx$ -0.3) and high surface gravity atmosphere ({$\log{g}$} $\approx$ 5.0--5.3~cgs).
In this article, we report our measurement of the near-infrared spectrum of {ULAS~J1416+1348} obtained with the NASA Infrared Telescope Facility (IRTF) SpeX spectrograph \citep{2003PASP..115..362R}. This spectrum encompasses the 0.8--2.4~$\micron$ region, including the metallicity-sensitive $Y$-band peak. The unusual shape of this and the $K$-band flux peak, along with fits to spectral models, affirm the interpretation of this source as a metal-poor, high surface gravity T7.5 brown dwarf, albeit with a {T$_{eff}$} that is significantly warmer than that reported
by \citet{2010arXiv1001.4393B}.
In Section~2 we describe our observations and discuss the spectral properties of {ULAS~J1416+1348}, including its classification, spectral anomalies and possible indications of {NH$_3$} absorption in the 1.0--1.3~$\micron$ region. In Section~3 we present our spectral model fits to the data and corresponding atmospheric parameters, as well as a model-dependent spectroscopic distance that is in accord with the astrometric distance of the primary. We discuss the relevance of this system with regard to the nature of blue L and T dwarfs in Section~4. Results are summarized in Section~5.
\section{Near Infrared Spectroscopy}
\subsection{Observations and Data Reduction}
Low resolution near-infrared spectral data for {ULAS~J1416+1348} were
obtained with SpeX on 2010 January 23 (UT) in mostly clear skies with some light cirrus and
0$\farcs$8 seeing. We used the SpeX prism mode with the 0$\farcs$5 slit aligned to the parallactic angle
for all observations, providing 0.7--2.5~$\micron$
coverage in a single order with resolution R~$\equiv$ {$\lambda/{\Delta}{\lambda}$} $\approx 120$
and dispersion of 20--30~{\AA}~pixel$^{-1}$.
{ULAS~J1416+1348} was acquired with the slit viewing camera using the $J$ filter
and guiding was performed on the nearby primary.
A total of 34 exposures of 180~s each were obtained in ABBA dither pairs,
nodding along the slit. The first 16 exposures of the source were obtained over
an airmass range of 1.28--1.41. We then observed the A0~V star HD~121880 ($V$ = 7.59)
at an airmass of 1.12 for flux calibration and telluric absorption correction, as well as
internal flat field and argon arc lamps for pixel response and wavelength calibration.
{ULAS~J1416+1348} was then reacquired and 18 more exposures made over an airmass range of 1.05--1.13.
Data were reduced with the IDL SpeXtool package, version 3.4
\citep{2004PASP..116..362C}, using standard settings.
Due to the faintness of {ULAS~J1416+1348} and its highly structured spectral morphology, individual spectra
were optimally extracted using a trace of HD~121880 as a template. These
spectra were then combined using a robust weighted average after scaling each to the median flux at the $J$-band peak. Telluric absorption and instrumental response corrections were determined from the A0~V spectrum following the method of \citet{2003PASP..115..389V},
with line-shape kernels derived from the arc lines and adjustments made to the H I line strengths and wavelength scale, as outlined in \citet{2004PASP..116..362C}.
\subsection{The Spectrum of {ULAS~J1416+1348}}
The reduced spectrum of {ULAS~J1416+1348} is shown in Figure~\ref{fig_nirspec},
compared to equivalent data for the T8 dwarfs 2MASS~J04151954$-$0935066 (hereafter 2MASS~J0415$-$0935; \citealt{2002ApJ...564..421B, 2004AJ....127.2856B})
and 2MASS~J09393548$-$2448279 (hereafter 2MASS~J0939$-$2448; \citealt{2005AJ....130.2326T, 2006ApJ...637.1067B}).
{ULAS~J1416+1348} exhibits the unambiguous signatures of a T dwarf, with
strong {H$_2$O} and {CH$_4$} absorption and a blue spectral energy distribution. The 1.6~$\micron$ {CH$_4$} band in the spectrum of {ULAS~J1416+1348} is slightly weaker than those of the T8 comparison sources, although the breadth of the $J$- and $H$-band peaks (both shaped by the wings of {H$_2$O} and {CH$_4$} bands) are equivalent
to the spectrum of 2MASS~J0415$-$0935. T dwarf classification indices \citep{2006ApJ...637.1067B} indicate a spectral type of T7.5$\pm$0.5 for this source, which is also consistent with its {NH$_3$}-H and $W_J$ indices (Table~\ref{tab_indices}; \citealt{2007MNRAS.381.1400W,2008MNRAS.391..320B, 2008A&A...482..961D}). This classification and most of the spectral indices are in agreement with those determined by \citet{2010arXiv1001.4393B}. However, we find a significant disagreement in our measurement of the {CH$_4$}-J index.\citet{2010arXiv1001.4393B} specifically note this index as anomalousm whereas our value is consistent with the overall spectral classification of {ULAS~J1416+1348}. As both spectra were obtained at roughly the same resolution ({$\lambda/{\Delta}{\lambda}$} $\approx$ 100), and signal-to-noise of the SpeX data in the $Y$-, $J$- and $H$-band peaks is good ($\sim$40--70), the origin of this anomaly is unclear.
What is most remarkable about the spectrum of {ULAS~J1416+1348} is the breadth of its 1.07~$\micron$ $Y$-band flux peak and strongly suppressed 2.10~$\micron$ $K$-band flux peak. The latter feature was also noted in the spectral data of \citet{2010arXiv1001.4393B}, and explains the very blue $J-K$ color of the source; we calculate a spectrophotometric color of $J-K = -1.71\pm0.23$\footnote{This value was determined by calculating the colors of 100 realizations of the spectral data, with fluxes varied following a normal distribution of the noise spectrum. We report here the mean and standard deviation of these measures.} from our SpeX spectrum, consistent with both the UKIDSS photometry and measurements by \citet{2010arXiv1001.4393B}. The broadened $Y$-band peak in the spectrum of {ULAS~J1416+1348}
is readily apparent, and similar to but more extreme than the broadened peak seen in the spectrum of 2MASS~J0939$-$2448 (Figure~\ref{fig_nirspec}; see also Figure~2 in \citealt{2006ApJ...639.1095B}). The origins of both features are discussed below.
\subsection{{NH$_3$} Absorption in the 1.0--1.3~$\micron$ Region?}
In addition to these broad spectral anomalies, we identified several intriguing absorption features around the $Y$-, $J$- and $H$-band flux peaks in the spectrum of {ULAS~J1416+1348}. As shown in Figure~\ref{fig_nirfeatures}, these features are at 0.997, 1.039, 1.072, 1.232, 1.249, 1.292, 1.302 and 1.570~$\micron$, none of which are present in the late-type T dwarfs with comparable SpeX data.\footnote{See \url{http://www.browndwarfs.org/spexprism}.} Among these features, the 1.072, 1.232 and 1.570~$\micron$ features are also seen in the absorption spectrum of Jupiter \citep{2009ApJS..185..289R}. Given tentative suggestions of the onset of {NH$_3$} absorption in the near-infrared spectra of the latest-type T dwarfs \citep{2000ApJ...541..374S, 2007ApJ...667..537L,2008A&A...482..961D}, we examined whether any of these features might be coincident with {NH$_3$} opacity. Figure~\ref{fig_nirfeatures} overlays the laboratory transmission function of {NH$_3$} from \citet{1999JQSRT..62..193I}, measured at temperatures of 200--300~K and pressures of 0.01--1~bar. Structure in the
{NH$_3$} spectrum appears to be coincident with
some of the features, most notably the 0.997, 1.039 and prominent 1.072~$\micron$ dips in the $Y$-band, and the weaker 1.292 and 1.302~$\micron$ dips in the $J$-band. However, strong {NH$_3$} opacity features, such as the 1.01--1.05 and 1.19--1.23~$\micron$ bands, are not seen in the data.
There are important caveats to such comparisons of opacity measurements to low-resolution brown dwarf spectra. First, opacity from several species, most notably {H$_2$O} and {CH$_4$} gas, blankets the entire near-infrared region, and the resolution of the SpeX data makes it impossible to separate narrow features from these species from coincident absorption arising from to {NH$_3$}. Second, the laboratory measurements of \citet{1999JQSRT..62..193I} were obtained in very different gas conditions than those that characterize the warmer photospheres of T dwarfs, and are not likely to include the higher angular momentum states expected in to be present in brown dwarf spectra. Indeed, \citet{2007ApJ...667..537L} have shown that current brown dwarf models incorporating the \citet{1999JQSRT..62..193I} opacities predict {NH$_3$} bands that are much stronger than observed, even when nonequilibrium abundances due to vertical mixing are considered \citep{2006ApJ...647..552S, 2007ApJ...669.1248H}.
In summary, while these features are notable, they cannot be conclusively associated with {NH$_3$} absorption. Higher resolution spectra coupled with better opacity data are needed to verify their origin.
\section{The Physical Properties of {ULAS~J1416+1348}}
\subsection{Qualitative indicators of High Surface Gravity and Subsolar Metallicity}
The broadened $Y$-band and suppressed $K$-band peaks in the spectrum of {ULAS~J1416+1348} are similar in nature to those seen in previously identified, unusually blue T dwarfs, and are indicative of pressure effects related to surface gravity and metallicity \citep{2002ApJ...564..421B, 2006ApJ...639.1095B, 2004AJ....127.3553K,2006AJ....131.2722C, 2007ApJ...667..537L, 2009ApJ...702..154S}.
$K$-band flux is regulated by collision-induced H$_2$ absorption \citep{1969ApJ...156..989L, 1994ApJ...424..333S, 2002A&A...390..779B}, which is sensitive to both photospheric gas temperature and pressure.
The short wavelength slope of the $Y$-band peak is shaped by the red wing of the pressure-broadened 0.77~$\micron$ \ion{K}{1} doublet, which is also modulated by temperature (affecting the K abundance) and pressure (affecting the pressure-broadened wings; \citealt{2003A&A...411L.473A,2003ApJ...583..985B}).
The $Y$-band peak has been specifically noted as being metallicity-sensitive in comparison of synthetic atmosphere
models, becoming both broadened and blue-shifted for lower metallicities \citep{2006ApJ...639.1095B, 2007ApJ...667..537L}.
The archetype blue T dwarf, 2MASS~J09373487+2931409 (hereafter 2MASS~J0937+2931; \citealt{2002ApJ...564..421B}) exhibits the same $Y$-band and $K$-band anomalies as {ULAS~J1416+1348} and, importantly, is consistently well-matched to models with subsolar metallicities ([M/H] = $-$0.1 to $-$0.4) and high surface gravities ({$\log{g}$} = 5.2 to 5.5; \citealt{2006ApJ...639.1095B,2009ApJ...695..844G}).
2MASS~J0939$-$2448 also exhibits these peculiarities (Figure~\ref{fig_nirspec}), and its near- and mid-infrared spectrum is well-matched to subsolar metallicity models as well, although it is additionally suspected of being an unresolved binary \citep{2008ApJ...689L..53B,2009ApJ...695.1517L}.
Importantly, the $Y$-band and $K$-band anomalies are more pronounced in the spectrum of {ULAS~J1416+1348} than in those of 2MASS~J0937+2931 and 2MASS~J0939$-$2448.
Our measure of the $K/J$ index---the relative flux between the $J$- and $K$-band peaks---is the smallest reported to date: 0.037$\pm$0.004 compared to 0.059 for 2MASS~J0939$-$2448 (see Table~6 in \citealt{2006ApJ...637.1067B} and Table~6 in \citealt{2009MNRAS.395.1237B}).
These measures suggest that {ULAS~J1416+1348} is a true outlier in terms of its physical properties.
\subsection{Comparison to Spectral Models}
To quantify these properties, we compared the spectrum of {ULAS~J1416+1348} to the
atmosphere models of \citet{2008ApJ...689.1327S}. We followed the prescriptions detailed in \citet{2008ApJ...678.1372C} and \citet{2008ApJ...689L..53B}, comparing our SpeX spectrum, flux-calibrated to the $J$-band photometry reported in \citet{2010arXiv1001.4393B}, to models spanning temperatures
{T$_{eff}$} = 500--1000~K (50~K steps); surface gravities {$\log{g}$} = 4.0--5.5~cgs (0.5~dex steps); metallicities
[M/H] = $-$0.3, 0 and +0.3 dex relative to Solar; and vertical diffusion coefficients {$K_{zz}$} = 0 and 10$^4$~{cm$^2$~s$^{-1}$} (see \citealt{2006ApJ...647..552S}). The models were smoothed to the resolution of the SpeX data
using a Gaussian kernel, and interpolated onto a common wavelength scale.
Fits were made exclusively to the 0.9--2.4~$\micron$ region. The goodness-of-fit statistic $G_k$ \citep{2008ApJ...678.1372C} was used to gauge the agreement between models and data, and we followed the same weighting scheme employed by those authors in which each pixel is weighted by its breadth in wavelength space. Model surface fluxes were scaled by the factor $C_k = (R/d)^2$ which minimizes $G_k$ (Equation~2 in \citealt{2008ApJ...678.1372C}), where $R$ is the radius of the brown dwarf and $d$ its distance from the Earth.
Fits were made to all 264 models. Distributions of the fit parameters were generated following a weighting scheme similar to that described in \citet{2008ApJ...689L..53B}, in which each model's parameters were incorporated into the distributions with a weight proportional to\footnote{Note that in \citet{2008ApJ...689L..53B}, the weighting function was e$^{-0.1G_k}$, a conservative choice that favored poorer-fitting models more highly. The 0.5 coefficient used here is more consistent with the probability distribution function of the $\chi^2$ statistic, for which $G_k$ is a close analog.} e$^{-0.5G_k}$. To examine the robustness of our fits to observational uncertainties, we also performed a Monte Carlo simulation similar to that described in \citet{2008ApJ...678.1372C} and \citet{2009ApJ...706.1114B}, generating 1000 realizations of the spectrum with fluxes randomly varied about the measured values following a normal distribution of the observational noise; the overall scaling of the spectrum was also varied following a normal distribution tied to the uncertainty in the $J$-band photometry. These spectra were compared to the 20 models that best fit the original spectrum, and distributions of the resulting best-fit parameter sets and $C_k$ scale factors were determined.
Table~\ref{tab_fits} summarizes the parameters of the ten best-fitting spectral models, while Figure~\ref{fig_modelfit} shows the overall best-fit model overlaid on the spectrum of {ULAS~J1416+1348}: {T$_{eff}$} = 650~K, {$\log{g}$} = 5.0~cgs, [M/H] = $-$0.3 and {$K_{zz}$} = 10$^4$~{cm$^2$~s$^{-1}$}. This model was the best fit for all of the spectra in our Monte Carlo simulation; i.e., its Monte Carlo fraction $f_{MC}$ = 1.000 \citep{2008ApJ...678.1372C}. It is a reasonably good match to the spectral
energy distribution, qualitatively reproducing the strong absorption bands, broadened $Y$-band peak and suppressed $K$-band peak, although the $H$-band peak flux is $\sim$10-15\% underestimated.
The parameter distributions from all of the model fits are also shown in Figure~\ref{fig_modelfit}.
Gaussian fits to these distributions yield optimal parameters of {T$_{eff}$} = 650$\pm$60~K and {$\log{g}$} = 5.2$\pm$0.4~cgs.
The metallicity distribution clearly favors a metal-poor atmosphere; the five best-fitting models all have a subsolar metallicity. In fact, the one-sided distribution in our limited model set means that we
cannot rule out metallicities less than $-0.3$. The model fits also indicate some vertical mixing is present, favoring {$K_{zz}$} = 10$^4$~{cm$^2$~s$^{-1}$} over 0~{cm$^2$~s$^{-1}$}, although a strict constraint cannot be made.
With respect to surface gravity and metallicity, our fits to the spectrum of {ULAS~J1416+1348} are in agreement with
the estimates of \citet{2010arXiv1001.4393B}, indicating that this unusually blue T dwarf is likely to be old, massive and metal-poor. The derived {T$_{eff}$} and {$\log{g}$} parameters and their uncertainties correspond to
an age of 2--10~Gyr and a mass of 0.021--0.045~{M$_{\sun}$} according to the evolutionary models of \citet{2003A&A...402..701B}. This age is consistent with membership in the Galactic disk population, as previously suggested by the system's kinematics \citep{2010ApJ...710...45B, 2009arXiv0912.3565S}. The subsolar metallicity favored by the model fits is in quantitative agreement with spectral analyses of other blue T dwarfs, as well as characterization of the L dwarf companion, SDSS~J1416+1348, which does not appear to be a full-fledged L subdwarf (\citealt{2010ApJ...710...45B}; however, see \citealt{kirkpatrick2010}). As such, these fits support our qualitative analysis of the spectrum: the spectral peculiarities and blue color of {ULAS~J1416+1348} appear to be the result of a high pressure atmosphere arising from high surface gravity and subsolar metallicity.
\subsection{The {T$_{eff}$} of {ULAS~J1416+1348}}
Our inferred {T$_{eff}$} for {ULAS~J1416+1348} is somewhat low for T7--T8 dwarfs, which typically have {T$_{eff}$} $\approx$ 700--900~K \citep{2004AJ....127.3516G,2004AJ....127.2948V, 2009ApJ...702..154S}. This is likely to be a surface gravity and/or metallicity effect.
\citet{2006ApJ...639.1095B} have previously found that late-type T dwarfs with higher surface gravities tend to have lower {T$_{eff}$}s for a given spectral type.
\citet{2010arXiv1001.4393B}, on the other hand, derive an even lower temperature for {ULAS~J1416+1348}, {T$_{eff}$} $\approx$ 500~K, based on this source's uniquely red $H$-[4.5] color. The link between {T$_{eff}$} and $H$-[4.5] color for brown dwarfs cooler than $\sim$1000~K was originally established by \citet{2007MNRAS.381.1400W}, and has been shown to provide increased sensitivity for the latest-type T dwarfs \citep{2009ApJ...702..154S,2010arXiv1001.0762L}. However, \citet{2009ApJ...695.1517L} have noted that metallicity effects are relevant and can shift $H$-[4.5] to the red by roughly 0.1~mag for every 0.1~dex decrement in metallicity (see also Figure~6 in \citealt{2010arXiv1001.4393B}). If {ULAS~J1416+1348} has a metallicity significantly below [M/H] = $-$0.3---not ruled out by the present model fits---then this characteristic may have as much to do with its extreme color as its low temperature.
It is relevant to note that the [3.6]-[4.5] color of {ULAS~J1416+1348}, another {T$_{eff}$} indicator \citep{2006ApJ...651..502P}, is not an extremum; this source is in fact bluer than 2MASS~J0939$-$2448. This may be an indication that the 3.3~$\micron$ {CH$_4$} band, like the 1.6~$\micron$ band, is relatively weak compared to other T8-T9 dwarfs, consistent with a warmer {T$_{eff}$}. However, metallicity and/or surface gravity effects may again complicate a strict correlation.
The \citet{2008ApJ...689.1327S} spectral model based on the atmosphere parameters favored by \citet{2010arXiv1001.4393B}---{T$_{eff}$} = 500~K, {$\log{g}$} = 5.0~cgs, [M/H] = $-$0.3~dex and {$K_{zz}$} = 10$^4$~{cm$^2$~s$^{-1}$---is also shown in Figure~\ref{fig_modelfit}. The cooler model actually provides a better match to the relative flux between the $J$- and $H$-band peaks and the width of the $J$-band peak; but predicts stronger 1.6~$\micron$ {CH$_4$} absorption, a far more distorted $Y$-band peak and a more suppressed $K$-band flux peak than observed. These deviations make this model a 2$\sigma$ outlier compared to the best-fit model for our data.
We emphasize that the differences between these fits do not explicitly rule out either set of parameters It is well known that incomplete opacity tables, inaccurate treatment of \ion{K}{1} pressure broadening and the influence of distributed condensate opacity (``cloud tops'') can result in poor fits to T dwarf near-infrared spectra \citep{2006ApJ...639.1095B, 2007ApJ...656.1136S, 2008ApJ...678.1372C,2009ApJ...702..154S}. However, to the limits of the accuracy of the current spectral models, our analysis favors a warmer temperature for {ULAS~J1416+1348} than indicated by its $H$-[4.5] color.
\subsection{Spectroscopic Distance}
Following \citet{2009ApJ...706.1114B}, we calculated a spectroscopic parallax for {ULAS~J1416+1348} using the model-to-data flux scaling factor $C_k$ derived from the spectral modeling. The mean value and uncertainty of this factor (based on the same $G_k$ weighting scheme used for the parameter distributions) yields a distance-to-radius ratio d/R = 12.8$\pm$3.0~pc/{R$_{Jup}$}. Based on the inferred {T$_{eff}$} and {$\log{g}$} range, the evolutionary models of \citet{2008ApJ...689.1327S} predict a radius $R$ = 0.83$^{+0.14}_{-0.10}$~{R$_{Jup}$}, corresponding to a distance of 10.6$^{+3.0}_{-2.8}$~pc. This is larger than but within 1$\sigma$ of the astrometric distance of the primary from \citet{2010arXiv1001.2743S}, 7.9$\pm$1.7~pc. In contrast, the 500~K model shown in Figure~\ref{fig_modelfit} requires d/R = 5.8~pc/{R$_{Jup}$}, and the corresponding $R$ = 0.73~{R$_{Jup}$} implies a distance of only 4.2~pc, significantly smaller than both spectrophotometric and astrometric estimates for SDSS~J1416+1348. Hence, to the limits of the accuracy of the spectral and evolutionary models of \citet{2008ApJ...689.1327S}, our atmospheric parameter determinations for {ULAS~J1416+1348} are commensurate with this source being cospatial with its co-moving L dwarf companion.
\section{The Nature of Blue L and T Dwarfs}
The SDSS~J1416+1348/{ULAS~J1416+1348} system provides a unique opportunity to explore the underlying physical properties that distinguish blue L and T dwarfs. While surface gravity and metallicity effects have long
been acknowledged as contributors to the peculiarities of blue T dwarfs,
condensate cloud properties have been seen as playing a more important role in
shaping the spectra of blue L dwarfs. Several studies have argued that thin and/or patchy condensate clouds in the photospheres of blue L dwarfs adequately explain their unique photometric and spectroscopic characteristics \citep{2004AJ....127.3553K, 2007AJ....133..439C,2008ApJ...674..451B, 2009ApJ...702..154S}. However, thin clouds cannot be responsible for the colors and spectra of late-type blue T dwarfs---such as {ULAS~J1416+1348}---since clouds are buried deep below the visible photosphere in these low-temperature objects \citep{2001ApJ...556..872A}.
The distinct empirical characteristics shared by SDSS~J1416+1348 and {ULAS~J1416+1348} must have an origin that is common to both sources; this argues for age and/or metallicity. Older ages for blue L and T dwarfs are supported by their collective kinematics; \citet{2009AJ....137....1F} and \citet{kirkpatrick2010} have shown that this subgroup exhibits
a much broader range of tangential velocities ($\sigma_{V} \approx$ 50~{km~s$^{-1}$}) than L and T dwarfs with ``normal'' colors ($\sigma_{V} \approx$ 22~{km~s$^{-1}$}). The high surface gravities inferred from spectral model fits to blue L and T dwarfs further support older ages for these sources (e.g., \citealt{2008ApJ...689L..53B, 2008ApJ...678.1372C, 2009ApJ...695..844G}). Subsolar metallicities are also supported by spectral model fits to blue T dwarfs, and the fact that blue L dwarfs exhibit spectral characteristics that are intermediate between normal field L dwarfs and halo L subdwarfs \citep{2004ApJ...614L..73B,kirkpatrick2010}. However, the discovery of a blue L5 dwarf companion to the solar-metallicity field M4.5 star G~203-50 \citep{2008ApJ...689..471R} suggests that metallicity does not play a consistent role in shaping these spectra.
We argue that the common photometric and spectroscopic properties of SDSS~J1416+1348 and {ULAS~J1416+1348} favors old age, and possibly subsolar metallicity, as the physical trait that characterizes the blue L and T dwarf populations.
Thin condensate clouds may still be common for blue L dwarf atmospheres, with higher surface gravities and subsolar metallicities contributing to increased sedimentation rates and a reduced supply of condensate species, respectively.
However, our conjecture predicts that these cloud properties are simply a consequence of the high-pressure photospheres characterizing old, high surface gravity and---in some cases---metal-poor brown dwarfs.
\section{Summary}
We have measured the 0.8--2.4~$\micron$ spectrum of {ULAS~J1416+1348}, the common proper motion companion
to the blue L dwarf SDSS~J1416+1348.
These data confirm the T7.5 spectral type determined by \citet{2010arXiv1001.4393B},
show possible {NH$_3$} features in the 1.0--1.3~$\micron$ region,
and reveal broadened $Y$-band and highly suppressed $K$-band peaks consistent with a
high surface gravity and/or subsolar metallicity. Spectral model fits based on calculations by
\citet{2008ApJ...689.1327S} indicate atmospheric
parameters {T$_{eff}$} = 650$\pm$60~K, {$\log{g}$} = 5.2$\pm$0.4~cgs, [M/H] $\leq$ -0.3 and {$K_{zz}$} = 10$^4$~{cm$^2$~s$^{-1}$}. The metallicity and surface gravity are consistent with the analysis by \citet{2010arXiv1001.4393B}, but our {T$_{eff}$} is $\sim$150~K (2.5$\sigma$) warmer. If correct, it suggests that the extreme $H$-[4.5] color of this source
may be due to metallicity and/or surface gravity effects, rather than an exceedingly low {T$_{eff}$}. Our fit parameters for {ULAS~J1416+1348} imply a model-dependent spectroscopic distance that is formally consistent with the astrometric distance of SDSS~J1416+1348 measured by \citet{2010arXiv1001.2743S}, and further strengthens the case that this pair is a coeval system of
unusually blue brown dwarfs. We argue that the common peculiarities of the SDSS~J1416+1348/{ULAS~J1416+1348} system implies that most unusually blue L and T dwarfs derive their unique properties from old age, and possibly subsolar metallicity, with the thin clouds of blue L dwarfs being a secondary effect.
Despite the substantial amount of follow-up already done for this fairly recent discovery, its benchmark role in understanding temperature, surface gravity, metallicity and cloud effects in L and T dwarf spectra motivates further observational study of both components. These include independent parallax measurements to verify absolute fluxes; higher-resolution near-infrared spectroscopy and mid-infrared spectroscopy of the secondary to validate potential {NH$_3$} features and discern the origin of its unusual mid-infrared colors; broad-band spectral energy distribution measurements of both components to measure luminosities and constrain {T$_{eff}$}s; high-resolution imaging to search for additional components; and improved model fits to better constrain atmospheric parameters.
In addition, the $\sim$100~AU projected separation of this system---wider than any L dwarf/T dwarf pair identified to date---raises questions as to the formation of it and other widely-separated, very low-mass stellar/brown dwarf multiples (e.g., \citealt{2004ApJ...614..398L,2005A&A...440L..55B, 2007ApJ...660.1492C}). Coupled with its proximity to the Sun, the SDSS~J1416+1348/{ULAS~J1416+1348} system is a target of opportunity
for studies of cold brown dwarf atmospheres and origins.
\acknowledgements
The authors acknowledge telescope operator Dave Griep
at IRTF for his assistance with the observations;
B.\ Burningham for providing an electronic version of the Irwin et al.\ {NH$_3$} opacity spectrum,
and D.\ Saumon for providing electronic copies of the spectral and evolutionary models used in the analysis.
This research has benefitted from the M, L, and T dwarf compendium housed at DwarfArchives.org and maintained by Chris Gelino, Davy Kirkpatrick, and Adam Burgasser;
the SpeX Prism Spectral Libraries, maintained by Adam Burgasser at
\url{http://www.browndwarfs.org/spexprism}; and
the VLM Binaries Archive maintained by Nick Siegler at \url{http://www.vlmbinaries.org}.
The authors recognize and acknowledge the
very significant cultural role and reverence that
the summit of Mauna Kea has always had within the
indigenous Hawaiian community. We are most fortunate
to have the opportunity to conduct observations from this mountain.
Facilities: \facility{IRTF~(SpeX)}
\clearpage
\begin{figure}
\epsscale{0.8}
\plotone{f1.eps}
\caption{SpeX prism spectrum of {ULAS~J1416+1348} (black line) compared
to equivalent data for the T8 dwarfs
2MASS~J0415$-$0935 (red line; \citealt{2004AJ....127.2856B})
and 2MASS~J0939$-$2448 (blue line; \citealt{2006ApJ...637.1067B}).
All three spectra
are normalized at 1.27~$\micron$, and the corresponding noise spectrum for {ULAS~J1416+1348}
is indicated by the light grey line.
Prominent {H$_2$O} and {CH$_4$} absorption features are labeled, as well as the region influenced by the pressure-broadened \ion{K}{1} doublet wing ($\lambda \lesssim 1$~$\micron$) and collision-induced H$_2$ opacity ($\lambda \gtrsim 1.75$~$\micron$).
\label{fig_nirspec}}
\end{figure}
\clearpage
\begin{figure}
\epsscale{1.0}
\plotone{f2.eps}
\caption{Close-up views of the 1.07~$\micron$ ($Y$-band, left), 1.27~$\micron$ ($J$-band, middle) and 1.58~$\micron$ ($H$-band, right) flux peaks in the spectra of {ULAS~J1416+1348} (black line) and
2MASS~J0415$-$0935 (red line).
Data are normalized in each panel to the peak flux in the given band.
Also shown is the normalized transmission spectrum of {NH$_3$}
from \citet{1999JQSRT..62..193I}. The transmission is magnified by a factor of two in the $Y$-band panel to highlight weaker bands.
The absorption features in the spectrum of {ULAS~J1416+1348} noted in the text are indicated by dashed lines.
\label{fig_nirfeatures}}
\end{figure}
\clearpage
\begin{figure}
\epsscale{1.0}
\includegraphics[width=0.95\textwidth]{f3a.eps}
\includegraphics[width=0.24\textwidth]{f3b.eps}
\includegraphics[width=0.24\textwidth]{f3c.eps}
\includegraphics[width=0.24\textwidth]{f3d.eps}
\includegraphics[width=0.24\textwidth]{f3e.eps}
\caption{(Top panel): Best-fitting model spectrum (red line) to our SpeX spectrum of {ULAS~J1416+1348} (black line): {T$_{eff}$} = 650~K, {$\log{g}$} = 5.0~cgs, [M/H] = -0.3 and {$K_{zz}$} = 10$^4$~{cm$^2$~s$^{-1}$}; and the model corresponding to the parameters of \citet{2010arXiv1001.4393B}: {T$_{eff}$} = 500~K, {$\log{g}$} = 5.0~cgs, [M/H] = -0.3 and {$K_{zz}$} = 10$^4$~{cm$^2$~s$^{-1}$}. The data are scaled to $J$-band photometry from \citet{2010arXiv1001.4393B} and the models are scaled to minimize $G_k$ values (indicated).
The noise spectrum for {ULAS~J1416+1348} is indicated by the grey line.
(Bottom panels): Parameter distributions of (left to right) {T$_{eff}$}, {$\log{g}$}, [M/H] and {$K_{zz}$} based on the weighting scheme described \citet{2008ApJ...689L..53B} and in the text. Means and uncertainties for {T$_{eff}$} and {$\log{g}$} are indicated in the first two panels and are based on Gaussian fits to the distributions. The metallicity distribution is such that we can only conclude that [M/H] $\leq$ -0.3, while the {$K_{zz}$} distribution indicates a slight preference for {$K_{zz}$} = 10$^4$~{cm$^2$~s$^{-1}$}.
\label{fig_modelfit}}
\end{figure}
\clearpage
\begin{deluxetable}{ccccc}
\tabletypesize{\footnotesize}
\tablecaption{Spectral Indices for {ULAS~J1416+1348}. \label{tab_indices}}
\tablewidth{0pt}
\tablehead{
\colhead{Index} &
\colhead{Value\tablenotemark{a}} &
\colhead{SpT} &
\colhead{Value B10\tablenotemark{b}} &
\colhead{Reference} \\
}
\startdata
{H$_2$O}-J & 0.053$\pm$0.008 & T8 & 0.07$\pm$0.01 & 1 \\
{CH$_4$}-J & 0.268$\pm$0.006 & T7 & 0.34$\pm$0.01 & 1 \\
$W_J$ & 0.376$\pm$0.005 & T7 & 0.34$\pm$0.01 & 2,3 \\
{H$_2$O}-H & 0.181$\pm$0.011 & T8 & 0.20$\pm$0.01 & 1 \\
{CH$_4$}-H & 0.197$\pm$0.010 & T7 & 0.20$\pm$0.01 & 1 \\
{NH$_3$}-H & 0.675$\pm$0.014 & \nodata & 0.61$\pm$0.01 & 4 \\
{CH$_4$}-K & 0.085$\pm$0.144 & T7 & 0.29$\pm$0.02 & 1 \\
K/J & 0.037$\pm$0.004 & \nodata & \nodata & 1 \\
\enddata
\tablenotetext{a}{Spectral index values were measured for 1000 realizations of the spectrum, each with a normal distribution of random values scaled by the noise spectrum added to the original fluxes. The reported values are the means and standard deviations of these measurements.}
\tablenotetext{b}{Spectral index values reported in \citet{2010arXiv1001.4393B} based on {$\lambda/{\Delta}{\lambda}$} $\approx$ 100 near-infrared spectral data.}
\tablerefs{(1) \citet{2006ApJ...637.1067B}; (2) \citet{2007MNRAS.381.1400W}; (3) \citet{2008MNRAS.391..320B}; (4) \citet{2008A&A...484..469D}.}
\end{deluxetable}
\begin{deluxetable}{lcccccc}
\tabletypesize{\footnotesize}
\tablecaption{Ten Best-Fitting \citet{2008ApJ...689.1327S} Spectral Models to SpeX Data for {ULAS~J1416+1348}. \label{tab_fits}}
\tablewidth{0pt}
\tablehead{
\colhead{Rank} &
\colhead{T$_{eff}$~(K)} &
\colhead{$\log{g}$~(cgs)} &
\colhead{[M/H]~(dex)} &
\colhead{$K_{zz}$~({cm$^2$~s$^{-1}$})} &
\colhead{$G_k$} &
\colhead{$d/R$ (pc/{R$_{Jup}$})}
\\
}
\startdata
1\tablenotemark{a} & 650 & 5.0 & -0.3 & 10$^4$ & 5.70 & 12.8 \\
2 & 700 & 5.0 & -0.3 & 10$^4$ & 5.97 & 15.5 \\
3 & 600 & 5.0 & -0.3 & 10$^4$ & 6.10 & 10.2 \\
4 & 700 & 5.0 & -0.3 & 0 & 6.53 & 15.4 \\
5 & 650 & 5.0 & -0.3 & 0 & 6.58 & 12.7 \\
6 & 600 & 5.5 & 0.0 & 10$^4$ & 6.85 & 10.5 \\
7 & 650 & 5.5 & 0.0 & 10$^4$ & 6.88 & 13.3 \\
8 & 700 & 5.5 & 0.0 & 10$^4$ & 7.31 & 15.7 \\
9 & 600 & 5.0 & -0.3 & 0 & 7.37 & 10.2 \\
10 & 650 & 5.5 & 0.0 & 0 & 8.01 & 13.3 \\
\cline{1-7}
Avg.\tablenotemark{b} & 650$\pm$60 & 5.2$\pm$0.4 & $\leq$-0.3 & $\sim10^4$ & \nodata & 12.8$\pm$3.0 \\
\enddata
\tablenotetext{a}{Best-fitting model for 1000 synthesized spectra in Monte Carlo simulation; i.e., $f_{MC} = 1.000$ (see \citealt{2008ApJ...678.1372C}).}
\tablenotetext{b}{Based on the weighted parameter distributions shown in Figure~\ref{fig_modelfit}. Each model contributes its parameters to the distributions scaled by the factor $e^{-0.5G_k}$. The means and uncertainties of {T$_{eff}$} and {$\log{g}$} were determined by Gaussian fits to their respective distributions (see \citealt{2008ApJ...689L..53B}). The [M/H] distribution peaked at the lower limit of the sampled parameter space, while the models slightly favor {$K_{zz}$} = 10$^4$~{cm$^2$~s$^{-1}$} over 0~{cm$^2$~s$^{-1}$}.}
\end{deluxetable}
\clearpage
|
\section{Introduction}
D$_p$-brane is a $(p+1)$-dimensional hypersurface in space-time
defined by the property that open strings can end on it. They have
been an active area of study and many remarkable properties of
them have been discussed \cite{Polchinski}. In
particular, they have provided a useful tool for the study of black
holes in string theory \cite{Strominger}; AdS/CFT correspondence
is another offspring of D$_p$-branes \cite{magoo}. Moreover, it is
well-known that IIA and IIB string theories have odd and even
BPS D$_p$-branes, respectively, and T-duality changes D$_p$-branes in
these theories to each other (or in other words it replaces a
scalar field with a gauge field and vice versa). D$_p$-branes are
sources of RR (P+1)-form field in IIA and IIB string theories.
There is an interesting candidate for DLCQ of M-theory
in terms of D$_{0}$-branes, the BFSS matrix model
\cite{Banks}. This conjecture tells us that all D$_p$-branes in IIA
superstring theory can be described in terms of D$_0$-branes and
their bound states. This is a requirement for the BFSS matrix model
as it is supposed that
the theory of D$_0$-branes describes the whole theory of DLCQ of
M-theory. In particular, e.g. a flat D$_2$-brane may be understand as
a bound state of N(N$\rightarrow\infty$) D$_0$-branes.
Therefore, understanding of connection among various
D$_p$-branes seems to be significant. It is well known that the low energy limit
of the D-brane action reduce to the Yang-Mills(YM) action. In YM limit, This
connection has already been addressed in the literature e.g. see
\cite{Susskind:2001fb} but here we use another way introduced in \cite{Ho} and
explain it more precisely. In fact in YM theory we consider
fluctuations around two directions which are transverse to multiple
D$_p$-branes and it will be shown that these two directions play the
role of components of gauge field in the world volume of
D$_{p+2}$-brane. We will then extend the prescription found at the
level of YM theory to terms, for both BPS and non-BPS
Dirac-Born-Infeld(DBI) action and their Chern-Simons actions as
well. The D$_{p+2}$-brane action is consistent with what is
expected.
The multiple D$_p$-branes theory has three parameters the string
length, $l_s$, string coupling constant, $g_s$, and the number of
D$_p$-branes, $N$. Then our prescription can be explained as follow.
Two transverse directions can be considered as a \emph{fuzzy}
torus\footnote{Although, other configurations can be considered.}
(compact transverse direction on fuzzy torus) or on the other
hand, for finite $N$, D$_p$-branes are uniformly distributed on these two
directions. Now two new parameters are added to our theory which are
radii of fuzzy torus, $R_1,R_2$. In order to take geometric
interpretation, there is a consistent limit which is when the number
of multiple D$_p$-branes and radii of fuzzy torus go to infinity
like $\theta=\frac{R_1R_2}{l_s^2N}\rightarrow 0$. As we will see in
this limit D$_p$-branes have been dissolved in the world volume of a
D$_{p+2}$-brane.\footnote{The other limits are when $\theta$ is constant or
goes to infinity. In both cases the final theories are non-commutative.}
In comparison to two T-dualities replacing two transverse directions with two
components of gauge field(or vice versa) on D$_p$-branes, in our
prescription, we consider \emph{multiple} D$_p$-branes and show
that two transverse directions replace with two components of gauge
filed on D$_{p+2}$-brane. In opposite way, one can argue that a
D$_{p+2}$-brane becomes \emph{multiple} D$_p$-branes if two
components of the gauge field are substituted by transverse scalar
fields.
This paper is organized as following.
In the two next sections the prescription is introduced and applied to YM theory
and DBI actions. We then show that the non-BPS D-branes
are consistent with above prescription and in the section five Chern-Simons action
is considered. The last section is devoted to conclusion.
\section{A large N limit of SYM theory}
The bosonic part of a $p+1$ dimensional U(N) SYM theory is described by the action %
\begin{equation}\label{action} %
S=-\frac{1}{2g_{YM}^2}\int d^{p+1}x{\rm{Tr}}\left((D_\mu
X^I)^2-\frac{1}{2}[X^I,X^J]^2+\frac12F_{\mu\nu}^2\right),
\end{equation} %
where\footnote{In terms of D-brane and string theory parameters note that
$g_{YM}^2=(\lambda^2T_p)^{-1}=(2\pi\lambda)^{\frac{p-1}{2}}\lambda^{-1}g_s,
\ \ \lambda=2\pi l_s^2$ where $T_p$ is the brane tension.} %
\begin{equation}\begin{split} %
F_{\mu\nu}&=\partial_\mu A_\nu-\partial_\nu A_\mu+i[A_\mu,A_\nu],\cr
D_\mu\Phi&=\partial_\mu\Phi+i[A_\mu,\Phi].
\end{split}\end{equation} %
$\mu(=0,...,p)$ and $I(=p+1,..,9)$. This action is also the action for
N coincident D$_p$-branes in the $\alpha'\rightarrow 0$ limit and in this picture
$\mu$ and $I$ denote world volume and transverse indices respectively.
The above action has following U(N) gauge symmetry %
\begin{subequations}\begin{align} %
\label{xtransformatiom}\delta_gX^I&=i[X^I,\alpha],\\
\delta_gA_\mu&=\partial_\mu\alpha+i[A_{\mu},\alpha].
\end{align}\end{subequations} %
In order to describe a D$_{p+2}$-brane, our prescription two steps. %
\begin{itemize}
\item
The $9-p$ transverses directions are decomposed as
\begin{equation} %
X^I\sim(y^{{\dot{\mu}}=1,2},X^{i=p+3,..,9}),
\end{equation} %
and the following replacement will then be done
\begin{subequations}\label{prescription}\begin{align} %
\label{prescription1}[\Phi(x),\Psi(x)]&\rightarrow
i\lambda\theta\{\Phi(x,y),\Psi(x,y)\},\\
\label{prescription2}\int{d^{p+1}x \rm{Tr}(\star) }&\rightarrow
\frac{1}{\lambda}\int{d^{p+1}xd^2y(\star)},
\end{align}\end{subequations}%
where $\{\Phi,\Psi\}=\epsilon^{{\dot{\mu}}{\dot{\nu}}}\partial_{\dot{\mu}} \Phi\partial_{\dot{\nu}}\Psi$(it is fixed such that
$\epsilon_{{\dot{\mu}}{\dot{\nu}}}\{y^{\dot{\mu}},y^{\dot{\nu}}\}=2$). $\Phi$ and $\Psi$ are arbitrary
fields and $\theta$ is a dimensionless constant. Note that, in new action, after above replacement all matrices in the YM theory
have been changed with fields. Furthermore, these fields are functions of $x,y$ and hence
$D_{\dot{\mu}} X^i$ is no longer zero. Another point is that
\eqref{prescription1} is right replacement when the number of
D$_p$-branes are large.
\item
Fluctuations around two transverse directions, $X^{\dot{\mu}}$, is
considered as %
\begin{equation}\label{newx} %
X^{\dot{\mu}}\equiv\frac{1}{\lambda\theta}y^{\dot{\mu}}-\epsilon^{{\dot{\mu}}{\dot{\nu}}}A_{\dot{\nu}}.
\end{equation} %
The form of above equation indicates that fluctuations play the
role of extra components of the world volume D$_{p+2}$-brane gauge
field which we will show it in detail.
\end{itemize}
Let's start with variation of scalar fields under gauge
transformation \eqref{xtransformatiom} which leads to %
\begin{equation}\label{xtransformatiomnew} %
\delta_gX^I=-(\delta_gy^{\dot{\mu}})\partial_{\dot{\mu}}X^I,
\end{equation} %
where %
\begin{equation} %
\delta_gy^{\dot{\mu}}=\lambda\theta\epsilon^{\dot{\mu}\dot{\nu}}\partial_{\dot{\nu}}\alpha.
\end{equation} %
Variation of the new components of the gauge field can be found by using
\eqref{newx} and \eqref{xtransformatiomnew} which is %
\begin{equation} %
\delta_gA_{\dot{\mu}}=\partial_{\dot{\mu}}\alpha+\lambda\theta\epsilon_{{\dot{\mu}}{\dot{\nu}}}\{X^{\dot{\nu}},\alpha\}.
\end{equation} %
The first term in the YM action can be simplified by using
\eqref{newx} and we then have %
\begin{equation} %
(D_\mu X^I)^2=(D_\mu X^i)^2+(F_{\mu{\dot{\nu}}})^2,
\end{equation} %
The second term gives $\{X^i,X^j\}^2$ and %
\begin{equation}\begin{split}\label{comrelation} %
\{X^{\dot{\mu}},X^{\dot{\nu}}\}^2&=\frac{2}{(\lambda\theta)^4}+\frac{1}{(\lambda\theta)^2}(F^{{\dot{\mu}}{\dot{\nu}}})^2+\ {\rm{total\ derivative\ terms}},\cr
\{X^{\dot{\mu}},X^i\}^2&=\frac{1}{(\lambda\theta)^2}(D^{\dot{\mu}} X^i)^2,
\end{split}\end{equation} %
where %
\begin{equation}\begin{split}\label{comrelation} %
\{X^{\dot{\mu}},X^{\dot{\nu}}\}&=\frac{\epsilon^{{\dot{\mu}}{\dot{\nu}}}}{(\lambda\theta)^2}-\frac{1}{\lambda\theta}\epsilon^{{\dot{\mu}}{\dot{\alpha}}}F_{{\dot{\alpha}}{\dot{\beta}}}\epsilon^{{\dot{\beta}}{\dot{\nu}}},\cr
\{X^{\dot{\mu}},X^i\}&=\frac{\epsilon^{{\dot{\mu}}{\dot{\alpha}}}}{\lambda\theta}D_{\dot{\alpha}} X^i ,
\end{split}\end{equation} %
and %
\begin{equation}\begin{split}%
F_{{\dot{\mu}}\nu}&=\partial_{\dot{\mu}} A_\nu-\partial_\nu A_{\dot{\mu}}-\lambda\theta\{A_{\dot{\mu}},A_\nu\},\cr
F_{{\dot{\mu}}{\dot{\nu}}}&=\partial_{\dot{\mu}} A_{\dot{\nu}}-\partial_{\dot{\nu}} A_{\dot{\mu}}-\lambda\theta\{A_{\dot{\mu}},A_{\dot{\nu}}\},\cr
D_{\dot{\mu}}\Phi&\equiv
-\lambda\epsilon_{{\dot{\mu}}{\dot{\nu}}}\{X^{\dot{\nu}},\Phi\}=\partial_{\dot{\mu}}\Phi-\lambda\theta\{A_{\dot{\mu}},\Phi\}.
\end{split}\end{equation} %
By plugging the above expressions in \eqref{action}, we arrive at a
\emph{non-commutative} D$_{p+2}$-brane action is appeared %
\begin{equation} %
S=-\frac{1}{2g^2_{YM}\lambda}\int d^{p+3}x \left((D_{\hat{\mu}}
X^i)^2+\frac{(\lambda \theta)^2}{4}\{X^i,X^j\}^2+\frac12 F_{\hat{\mu}\hat{\nu}}^2+\frac{1}{2(\lambda
\theta)^2}\right),
\end{equation} %
where %
\begin{equation}\begin{split} %
F_{\hat{\mu}\hat{\nu}}&=\partial_{\hat{\mu}} A_{\hat{\nu}}
-\partial_{\hat{\nu}} A_{\hat{\mu}}+\lambda \theta\{A_{\hat{\mu}},A_{\hat{\nu}}\},\cr
D_{\hat{\mu}}\Phi&=\partial_{\hat{\mu}}\Phi+\lambda
\theta\{A_{\hat{\mu}},\Phi\}.
\end{split}\end{equation} %
$\hat{\mu}=(\mu,{\dot{\mu}})$ denotes the world volume index of
D$_{p+2}$-brane and the above action is invariant under %
\begin{equation}\begin{split} %
\delta_gA_\mu&=\partial_\mu\alpha+\lambda \theta\{A_{\mu},\alpha\},\cr
\delta_gA_{\dot{\mu}}&=\partial_{\dot{\mu}}\alpha+\lambda
\theta\epsilon_{{\dot{\mu}}{\dot{\nu}}}\{X^{\dot{\nu}},\alpha\}.
\end{split}\end{equation} %
As we will see in next section this action is a low energy limit of
\emph{non-commutative} D$_{p+2}$-brane action \eqref{nonaction}
where the B-field turns on as
$B_{{\dot{\mu}}{\dot{\nu}}}=\frac{1}{\theta}\epsilon_{{\dot{\mu}}{\dot{\nu}}}$. Hence the arbitrary
constant, $\theta$, plays the role of \emph{non-commutative
parameter} and it is acceptable to recover the commutative action
when $\theta\rightarrow0$. Obviously, the U(1) gauge theory action
for D$_{p+2}$-brane is recovered although there is an extra term
going to infinity in the action. This term can be considered as zero point energy
of multiple D$_p$-branes dissolved in the D$_{p+2}$-brane world
volume.
\section{ DBI theory}
One of the remarkable properties of D$_p$-brane is that the U(1)
gauge symmetry is enhanced to U(N) gauge symmetry for N coincident
D$_p$-branes. The suitable action for multiple
D$_p$-branes was introduced in \cite{Myers}. The action is %
\begin{equation}\begin{split}\label{mdbiactin} %
S&=-T_p\int d^{p+1}x\cr
&\times{\rm{STr}}\Bigg(
\sqrt{-\det\left({\rm{P}}[E_{mn}
+E_{m I}(Q^{-1}-\delta)^{IJ}E_{Jn}]+\lambda
F_{\mu\nu}\right)\det(Q^I_J)}\Bigg),
\end{split}\end{equation} %
where $E_{mn}=G_{mn}+B_{mn},\
Q^I_J\equiv\delta^I_J+i\lambda[X^I,X^K]E_{KJ}$. The brane tension is
$T_p=\frac{2\pi}{g_s(2\pi l_s)^{p+1}}$($l_s$ and $g_s$ are string
length and coupling) and "P" denotes pull-back of background metric
and NSNS two-form($m,n=0,..,9$). $F_{\mu\nu}$ is field strength of
gauge field living on the D$_p$-brane. STr(...) denotes that
one takes a symmetrized average over all ordering of various fields appearing
inside the trace.
Let us again consider that two longitudinal directions of D-branes
fluctuate such that
$X^{\dot{\mu}}=\frac{1}{\lambda\theta}y^{\dot{\mu}}-\epsilon^{{\dot{\mu}}{\dot{\nu}}}A_{\dot{\nu}}$.
Working, in static gauge \textit{i.e.} $\lambda X^{\mu}=x^{\mu}$, in
flat space background and noting that the $B$ is turned off in
D$_p$-brane theory, the determinant in the action can be
decomposed as
\begin{equation} %
\tilde{D}=\det\left(%
\begin{array}{ccc}
\eta_{\mu\nu}+\lambda F_{\mu\nu} & \lambda D_\mu X^{\dot{\nu}} & \lambda D_\mu X^j \\
-\lambda D_\nu X^{\dot{\mu}} & \delta^{{\dot{\mu}}{\dot{\nu}}}+i\lambda X^{{\dot{\mu}}{\dot{\nu}}} & i\lambda X^{{\dot{\mu}} j} \\
-\lambda D_\nu X^i & i\lambda X^{i{\dot{\nu}}} & Q^i_k\delta^{kj} \\
\end{array}%
\right). %
\end{equation} %
Using \eqref{comrelation} and
\begin{equation} \label{fields}%
D_{\mu}X^{{\dot{\nu}}}=F_{\mu{\dot{\alpha}}}\epsilon^{{\dot{\alpha}}{\dot{\nu}}},
\end{equation} %
one can rewrite the determinant as
\begin{equation} %
\tilde{D}=\det\left(%
\begin{array}{ccc} \label{nondbi}
\eta_{\mu\nu}+\lambda F_{\mu\nu} & \lambda F_{\mu{\dot{\alpha}}}\epsilon^{{\dot{\alpha}}{\dot{\mu}}} & \lambda D_\mu X^i \\
-\lambda F_{\nu{\dot{\alpha}}}\epsilon^{{\dot{\alpha}}{\dot{\mu}}} & \delta^{{\dot{\mu}}{\dot{\nu}}}-\frac{1}{\theta}\epsilon^{{\dot{\mu}}{\dot{\nu}}}+\lambda \epsilon^{{\dot{\mu}}{\dot{\alpha}}}F_{{\dot{\alpha}}{\dot{\beta}}}\epsilon^{{\dot{\beta}}{\dot{\nu}}} & -\lambda \epsilon^{{\dot{\mu}}{\dot{\alpha}}}D_{\dot{\alpha}} X^i \\
-\lambda D_\nu X^i & \lambda \epsilon^{{\dot{\nu}}{\dot{\alpha}}}D_{\dot{\alpha}} X^i & Q^i_k\delta^{kj} \\
\end{array}%
\right). %
\end{equation} %
Using the standard trick for recombining the determinant\cite{Myers}
\begin{equation} %
\tilde{D}=\det\left(%
\begin{array}{ccc} \label{nondbi}
A & B \\
C & D
\end{array}%
\right) %
=\det{(A-BD^{-1}C)\det{(D)}},
\end{equation}%
one can rearrange the action in a compact form similar to
\eqref{mdbiactin}. As it can be seen, in the new action a
constant $B$-field is turned on in new directions ${\dot{\mu}}$'s. Due to
the facts that the covariant derivatives $D_{\hat{\mu}}$ and
field strength $F_{\hat{\mu}\hat{\nu}}$ have now a non-vanishing first
order $\theta$ term and the $B$ field is present, the new action is the $\theta\rightarrow
0$ limit of a \emph{non-commutative} $U(1)$ D$_{p+2}$ brane action\cite{Ardalan}.
The general form of such an action is
\begin{equation}\begin{split}\label{nonaction} %
\hat{S}&=-T_p\int d^{p+1}x\Bigg(e^{-\phi(\hat{A})}\cr
&\times\sqrt{-\det\left({\rm{P}}_\theta[E_{mn}(\hat{A})
+E_{m I}(\hat{A})(Q^{-1}-\delta)^{IJ}E_{Jn}(\hat{A})]+\lambda
\hat{F}_{mn}\right)\det(Q^I_J)}\Bigg)*_{N}
\end{split}\end{equation} %
where
$Q^I_J\equiv\delta^I_J+i\lambda[\hat{X}^I,\hat{X}^K]_ME_{KJ}(\hat{A})$ and "$\ \hat{}\ $"
denotes non-commutative fields. Here the $*_N$ is the multiplication rule
among various non-commutative fields \cite{Liu2}.
One can show that in
the limits where the non-commutativity parameter $\theta$ goes to
zero this action reduces to DBI action.
\section{Non-BPS DBI action}
Besides of BPS branes, which are charges of RR fields, non-BPS branes may
exist in the IIA(B) theory. The existence of these branes may cause the instability
and breaks the supersymmetry. From the field theory of world volume of D-brane point of view,
the existence of a tachyonic field T is the reason of such phenomena.
In general, the action of non-BPS D$_p$-branes can be written as the
product of the action of BPS D$_p$-branes and contribution which comes from the tchyon field such as
\begin{eqnarray}%
S_{Non-BPS}=T_p\int{d^{p+1}x\ S_{BPS}F(T)},
\end{eqnarray}%
where $F(T)$ is the tachyonic contribution of the action of
D$_p$-brane.
The form of this action can be obtained by using the
non-BPS D$_9$(or D$_8$)-brane action and then performing the
T-duality transformations\cite{Kluson}. Its most general form up to
first order derivative of $T$ may be written as
\newpage
\begin{eqnarray}%
F(T)=V(T)&-&\Sigma_{n=1}^{\infty}f_n(T)\lambda^n(( E^{\mu\nu}-E^{\mu
I}E_{IJ}E^{J\nu})D_{\mu}TD_{\nu}T\cr &+&2i E^{\mu
K}E_{KJ}[X^J,T]D_{\mu}T-E_{IJ}[X^I,T][X^J,T])^n,
\end{eqnarray}%
and $V(T)$ is the potential for the tachyon. Functions $f_n(T)$
are some even functions of the tachyon field \cite{Sen}, and although their explicit
form are not known, there
are some conjectures about it \cite{Garousi,Bergshoeff}.
Using the prescription \eqref{prescription1} in flat space
background the second term vanishes and so $F(T)$ reads as
\begin{eqnarray}%
F(T)=V(T)-\Sigma_{n=1}^{\infty}f_n(T)\lambda^n(( E^{\mu\nu}-E^{\mu
I}E_{IJ}E^{J\nu})D_{\mu}TD_{\nu}T\cr
+(\lambda\theta)^2E_{IJ}\{X^I,T\}\{X^J,T\})^n.
\end{eqnarray}%
Applying
$X^{{\dot{\mu}}}=\frac{1}{\lambda\theta}y^{{\dot{\mu}}}-\epsilon^{{\dot{\mu}}{\dot{\nu}}}A_{{\dot{\nu}}}$
and noting that
$$\epsilon_{{\dot{\mu}}{\dot{\nu}}}D^{{\dot{\mu}}}TD^{{\dot{\nu}}}T=0,$$ one easily finds that the
tachyonic part of Non-BPS multiple D$_p$-branes action changes to
tachyonic part of a Non-BPS D$_{p+2}$-brane action.
\section{Chern-Simons term}
In this section we examine that the Chern-Simons term of N D$_p$-brane action also
reproduce the Chern-Simons term of a single D$_{p+2}$-brane in the large N limits
discussed in previous sections. As it is well known, the world volume of D$_p$
branes in IIA(B) string theory can couple to RR fields of rank lower
than the dimension of brane and also can couple to RR fields with
higher rank due to Myers terms\cite{Myers}. In fact, the matrix representation
of fields and then the non-commutativity of such fields in non-Abelian theories
allows one to couple a combination of form fields of higher rank and commutators of
scalar fields with world volume of D-brane in a covariant manner.
On the other hand, the Chern-Simons action is given by \cite{Myers} %
\begin{eqnarray} %
S_{CS}=\mu_p\int{{\rm{STr}}\left({\rm{P}}[e^{i\lambda{\rm{i}_X}{\rm{i}_X}}
(\Sigma C^{(n)}e^{B})]e^{\lambda F}\right),}
\end{eqnarray} %
where $C^{(n)}$ denotes $(p+1)$-forms and $\mu_p$ is the RR charge
of the brane and $\rm{i}_X$ is the exterior derivative in $X$ direction which acts on RR
form fields. This special form of the action is necessary for
compatibility with the T-duality transformations of various IIA and
IIB fields. Note that all fields are in adjoint representation of
the U(N) gauge group.
The general couplings of these form fields with world volume of
D$_p$-branes have many terms so, for simplicity, we study
D$_1$-D$_3$ transition and only consider the couplings of all form
fields up to the first power of fields.
Before we proceed, we mention that due to the
non-commutativity of two longitudinal directions of D$_1$-branes, one
may define \cite{Mukhi,Liu} a new two form field $Q^{(2)}$ which its
components are
\begin{equation} %
Q^{(2)}\equiv-i\lambda^3\theta [X_{{\dot{\mu}}},X_{{\dot{\nu}}}]dX^{{\dot{\mu}}}\wedge
dX^{{\dot{\nu}}}.
\end{equation}%
So, one should consider the couplings of this two form field with
world volume of D$_1$-branes. We will see that this form fields has
important role in this story.
Finally, we will do our computations in the regime where $\theta
\rightarrow 0$ and
also consider the constant $C_{{\dot{\mu}}{\dot{\nu}}}$ and $F_{{\dot{\mu}}{\dot{\nu}}}$ fields such
that
\begin{equation}%
C_{{\dot{\mu}}{\dot{\nu}}}=\lambda^2\epsilon_{{\dot{\mu}}{\dot{\nu}}},\;\;\;\;\;F_{{\dot{\mu}}{\dot{\nu}}}=\frac{\epsilon_{{\dot{\mu}}{\dot{\nu}}}}{\lambda}.
\end{equation}%
By the above assumptions we have
\begin{eqnarray}%
\lambda^2\theta\{X^{{\dot{\mu}}},X^{{\dot{\nu}}}\}&\rightarrow&
\epsilon^{{\dot{\mu}}{\dot{\nu}}}(\frac{1}{\theta}-1),\;\;\;\;\;\;Q_{{\dot{\mu}}{\dot{\nu}}}\rightarrow
\lambda^2\epsilon_{{\dot{\mu}}{\dot{\nu}}},\cr
\lambda^2\theta\{X^{{\dot{\mu}}},X^{i}\}&\rightarrow& \lambda
\epsilon^{{\dot{\mu}}{\dot{\alpha}}}D_{{\dot{\alpha}}}X^{i},\;\;\;\;\;\;\theta\{X^i,X^j\}\rightarrow
0,
\end{eqnarray}%
After all, recalling that in D$_1$-branes we have $B=0$ then the
Chern-Simons term is equal to
\begin{eqnarray} \label{csd1}%
S_{CS}&=&\frac{1}{\lambda}\int{\rm{STr}}\left({\rm{P}}[e^{i\lambda \textbf{i}_X\textbf{i}_X}(
C^{(2)}(1+Q^{(2)}))]e^{\lambda F}\right)\cr
&\rightarrow &\frac{1}{\lambda^2}
\int{d^4x\left(\frac{1}{2}C_{mn}D_{\mu} X^mD_{\nu} X^n\right)\epsilon^{\mu\nu}}\cr
&+&\frac{1}{\lambda^2}\int{d^4x\left(\frac{-\lambda^2\theta}{2\lambda^2}\{X^{m},X^{n}\}C_{nm}\right)\lambda F_{\mu\nu}\epsilon^{\mu\nu}}\cr
&+&\frac{1}{\lambda^2}\int{d^4x\left(\frac{-\lambda^2\theta}{2\lambda^2}\{X^{m},X^{n}\}C_{nm}\right)
Q_{{\dot{\mu}}{\dot{\nu}}}D_{\mu} X^{{\dot{\mu}}}D_{\nu} X^{{\dot{\nu}}}\epsilon^{\mu\nu}}\cr
&+&\frac{1}{\lambda^2}\int{d^4x\left(\frac{-\lambda^2\theta}{2\lambda^2}\{X^{{\dot{\mu}}},X^{{\dot{\nu}}}\}Q_{{\dot{\nu}}{\dot{\mu}}}\right)
C_{mn}D_{\mu} X^{m}D_{\nu} X^{n}\epsilon^{\mu\nu}}\cr
&+&\frac{1}{\lambda^2}\int{d^4x\left(\frac{-\lambda^2\theta}{2\lambda^2}\{X^{{\dot{\mu}}},X^{m}\}C_{[m[n}Q_{{\dot{\mu}}]{\dot{\nu}}]}\right)
D_{\mu} X^{n}D_{\nu} X^{{\dot{\nu}}}\epsilon^{\mu\nu}},
\end{eqnarray} %
where the last term is antisymmetric for the pairs $(m,{\dot{\mu}})$ and
$(n,{\dot{\nu}})$. Note also that due to dimensions of form fields any
contraction with these fields leaves a factor $\frac{1}{\lambda}$ in
the action.
Using \eqref{comrelation} and \eqref{fields} one obtains
\begin{eqnarray}%
C_{mn}D_{\mu} X^mD_{\nu} X^n&=&
\frac{1}{\lambda^2}C_{\mu\nu}-\frac{1}{\lambda}C_{i[\mu}D_{\nu]}X^i+C_{ij}D_{\mu}X^iD_{\nu}X^j\cr
&-&\frac{1}{\lambda}C_{{\dot{\mu}}[\mu}F_{\nu]{\dot{\alpha}}}\epsilon^{{\dot{\alpha}}{\dot{\mu}}} -C_{{\dot{\mu}}
i}D_{[\mu}X^iF_{\nu]{\dot{\alpha}}}\epsilon^{{\dot{\alpha}}{\dot{\mu}}}+{\cal{O}}(F^2),
\end{eqnarray}%
and one may rewrite the integrand of the above expression term by
term as
\begin{eqnarray}%
&+&\frac{1}{2\lambda^2}\left(C_{mn}D_{\mu}
X^mD_{\nu}X^n\right)\epsilon^{\mu\nu}\cr
&-&\frac{1}{2\lambda^2}\left(2\lambda(\frac{1}{\theta}-1)+2C_{i{\dot{\mu}}}D_{{\dot{\alpha}}}X^i\epsilon^{{\dot{\mu}}{\dot{\alpha}}}+{\cal{O}}(\theta)\right)F_{\mu\nu}\epsilon^{\mu\nu}\cr
&-&\frac{1}{2\lambda^2}\left({\cal{O}}(F^2)\right)\cr
&-&\frac{1}{2\lambda^2}\left(2(\frac{1}{\theta}-1)C_{mn}D_{\mu}
X^mD_{\nu} X^n\right)\epsilon^{\mu\nu}\cr
&-&\frac{1}{2\lambda^2}\left(2(\frac{1}{\theta}-1)(C_{\mu{\dot{\nu}}}+C_{i{\dot{\nu}}}D_{\mu}X^i)F_{\nu{\dot{\beta}}}\epsilon^{{\dot{\beta}}{\dot{\nu}}}+{\cal{O}}(F^2)\right)\epsilon^{\mu\nu}\cr
&-&\frac{1}{2\lambda^2}\left((C_{\mu
i}+C_{ji}D_{\mu}X^j)F_{\nu{\dot{\beta}}}D_{{\dot{\nu}}}X^i\epsilon^{{\dot{\beta}}{\dot{\nu}}}+{\cal{O}}(F^2)\right)\epsilon^{\mu\nu}
\end{eqnarray}%
where the two last terms come from the last term of \eqref{csd1}.
Now, for D$_3$-brane, the Chern-Simons term up to first power of
fields is equal to
\begin{eqnarray} \label{chd3}%
S_{CS}&=&\frac{1}{\lambda^2}\int{\rm{STr}}\left(P[e^{i \lambda \textbf{i}_X\textbf{i}_X}(\Sigma
C^{(n)}e^B)]e^{\lambda F}\right)\cr
&\rightarrow&\frac{1}{2\lambda^2}\left(\int d^2x\left(
C_{mn}B_{pq}
D_{\hat{\mu}} X^mD_{\hat{\nu}} X^nD_{\hat{\lambda}} X^p D_{\hat{\theta}}
X^q\right)\epsilon^{\hat{\mu}\hat{\nu}\hat{\lambda}\hat{\theta}}+\rm{antisymm\;part}\right)\cr
&+&\frac{1}{2\lambda^2}\left(\int{d^2x\left(C_{mn}D_{\hat{\mu}}
X^mD_{\hat{\nu}} X^n\lambda F_{\hat{\lambda}\hat{\theta}}\right)
\epsilon^{\hat{\mu}\hat{\nu}\hat{\lambda}\hat{\theta}}}+\rm{antisymm\;part}\right),
\end{eqnarray} %
where by \textit{antisymm part} we mean that, for example, the terms
of $C_{mn}B_{pq}$ are antisymmetric under the exchange of indices of
$C$ and $B$ fields. Noting that in D$_3$-brane we have found
$P(B)=\frac{\epsilon_{{\dot{\mu}}{\dot{\nu}}}}{\theta}$, we rewrite the $S_{CS}$
of D$_3$ brane as\newpage
\begin{eqnarray} %
S_{CH}&=&\frac{1}{2\lambda^2}\int{d^2x
\left(C_{\mu\nu}-2C_{i\mu}D_{\nu}X^i+C_{ij}D_{\mu}X^id_{\nu}X^j\right)\epsilon_{{\dot{\mu}}{\dot{\nu}}}\epsilon^{\mu\nu{\dot{\mu}}{\dot{\nu}}}}\cr
&+&\frac{1}{2\lambda}\int{d^2x\left(C_{mn}D_{\hat{\mu}}X^mD_{\hat{\nu}}X^{n}F_{\hat{\lambda}\hat{\theta}}\right)
\epsilon^{\hat{\mu}\hat{\nu}\hat{\lambda}\hat{\theta}}}.
\end{eqnarray} %
It is not hard to show that the above D$_3$-brane Chern-Simons terms
reproduce the D$_1$-branes Chern-Simons terms. We see the role of
the two form field $Q^{(2)}$ in this D$_1$-D$_3$ transition in which
some terms in D$_3$ such as $C_{\mu
i}F_{\nu{\dot{\beta}}}D_{{\dot{\nu}}}X^i\epsilon^{{\dot{\beta}}{\dot{\nu}}}\epsilon^{\mu\nu}$ come from
the coupling of this form field with $C^{(2)}$ and $F^{(2)}$ in
\eqref{csd1}.
The general form of Chern-Simons action in non-commutative theories
in the presence of
non-zero constant B field may be written as\cite{Mukhi,Liu,Mukhi2}%
\begin{equation} %
S_{CS}=\mu_p\int\frac{\rm{Pf}\hat{Q}}{\rm{Pf}\theta}\left(P[e^{i(i_X*i_X)}(\Sigma
C^{(n)})]e^{B+\lambda F}\right), %
\end{equation} %
where
$\hat{Q}^{mn}=\theta^{mn}-\theta^{m\hat{\alpha}}\hat{F}_{\hat{\alpha}\hat{\beta}}\theta^{\hat{\beta}n}$
and $\theta^{mn}=(B^{-1})^{mn}$, the Pf denotes the Pfaffian of an antisymmetric matrix and
all products are understood as $\ast$ product. It can be seen that in the $\theta\rightarrow 0$ limit
this action and the action \eqref{chd3} coincide with each other.
\section{Conclusion}
In this paper we have shown that in flat space multiple
D$_p$-branes can be considered as a D$_{p+2}$-brane for large number
of D$_p$-branes. In fact matrix valued $p+1$ dimension fields go
to $p+3$ dimension fields. The point is that two transverse
directions to multiple D$_p$-branes appear as two components of gauge
field living on the D$_{p+2}$-brane. One can run this prescription
in opposite way and start with a D$_{p+2}$-brane and
it finally leads to multiple D$_p$-branes. In other words, one exchanges two
scalar fields with two components of gauge fields and at the end we
have a D$_{p+2}$-brane or multiple D$_{p-2}$-branes.
Although, this setup has been done in flat space the above idea
can be generalized to curved background. In well-known curved background such as
pp-wave our half BPS D$_p$-branes have spherical symmetry and we may
then expect multiple D$_p$-branes lead to a spherical D$_{p+2}$-brane.
Moreover, by using the above prescription, we expect a relation
between mN D$_{p}$-branes and m D$_{p+2}$-branes in large N limit. Such idea
is useful to extract new understanding of the theory of multiple non-commutative D-branes.
Understanding of three dimensional conformal field theory(CFT) was an open
issue for years \cite{Schwarz}. Symmetries of this theory are consistent with
symmetries of multiple M2-branes. Recently, a groundbreaking three dimensional CFT was presented in
\cite{Bagger} known as BLG theory.
In BLG theory, there are a lot of attempts to show that a M5-brane action
leads to multiple M2-branes action or vice versa. We hope that uplifting the above
results teach us more about M2-M5 relation. In the other hand in comparison to
non-commutative D$_p$-brane, it seems that one should know about "\emph{non-commutative}"
M5-brane to explain the relation correctly, although the geometry of M5-brane
is not known by now \cite{Chen}.
\section*{ Acknowledgment}
We would like to specially thank M. M. Sheikh-Jabbari for careful
reading of the manuscript and valuable discussion.
|
\section{I) Nonlinear response measurements through third harmonics detection}
As in the main paper, we consider the nonlinear response of a dielectric system to a time dependent electric field $E(t)$. The most general relationship between the polarisation $P(t)$ and the excitation $E(t)$ is a series expansion in $E$ \cite{SThibierge}. For a purely ac field, the even terms are forbidden because of the $E(t) \to -E(t)$ symmetry, which yields :
\begin{equation}
\frac {P(t)}{\epsilon_0} = \int_{-\infty}^{\infty}\chi_1(t-t')E(t')dt' + \int_{-\infty}^{\infty}\chi_3(t-t'_1,t-t'_2,t-t'_3)
\times E(t'_1)E(t'_2)E(t'_3)dt'_1dt'_2dt'_3 + ...
\label{eqchidet}.
\end{equation}
In this equation $\epsilon_0$ is the dielectric constant of vacuum, $\chi_1$ the linear susceptibility and $\chi_3$ the cubic nonlinear susceptibility in the time domain. The dots in Eq.~\ref{eqchidet} indicate an infinite sum involving higher order nonlinear susceptibilities $\chi_5$, etc. Note that causality implies $\chi_{i}(t<0) =0$. The Fourier transform of Eq.~\ref{eqchidet} for a purely ac field $E(t)=E \cos(\omega t)$ gives:
\begin{eqnarray}
\frac {P(\omega')}{\epsilon_0} &=& \frac{E}{2}\left[ \chi_1(\omega)+ \frac{3E^2}{4}\chi_3(-\omega,\omega,\omega)+... \right] \delta(\omega'-\omega) \nonumber \\
&+& \frac{E}{2}\left[ \chi_1(-\omega)+ \frac{3E^2}{4}\chi_3(\omega,-\omega,-\omega)+... \right] \delta(\omega'+\omega) \nonumber \\
&+& \frac{E^3}{8}\chi_3(\omega,\omega,\omega) \delta(\omega'-3\omega) \nonumber \\
&+& \frac{E^3}{8}\chi_3(-\omega,-\omega,-\omega) \delta(\omega'+3\omega)+...
\label{eqchideomega} ,
\end{eqnarray}
where the polarization $P$ and the susceptibilities $\chi_i$ are now taken in the frequency domain and the dots indicate again infinite sums involving higher order terms.
The response $P(t)$ to $E(t)$ = $E \cos(\omega t)$ can thus be written
\begin{equation}
P(t)/\epsilon_0 = Re \left[(E \chi_1(\omega)+3/4E^3 \chi_{\bar{3}}(\omega)+ ...) e^{-i\omega t}\right]
+ Re \left[1/4E^3 \chi_3(\omega) e^{-i3\omega t} + ...\right]+ ....
\label{eqchideomegapract1}
\end{equation}
To obtain Eq.~(\ref{eqchideomegapract1}), we have used the fact that because $\chi_1$ and $\chi_3$ are real in the time domain, their Fourier transform verify
$\chi_1^*(\omega)$ = $\chi_1(-\omega)$ and $\chi_3^*(\omega_1,\omega_2,\omega_3)$ = $\chi_3(-\omega_1,-\omega_2,-\omega_3)$ (the star denotes the complex conjugate), and the invariance of $\chi_3$ by permutation of its arguments. For simplicity, we write $\chi_3(\omega)$ = $\chi_3(\omega,\omega,\omega)$ and $\chi_{\bar{3}}(\omega)$ = $\chi_3(-\omega,\omega,\omega)$. Eq.~(\ref{eqchideomegapract1}) can be written
\begin{multline}
P(t)/\epsilon_0 = E(\chi'_1 \cos\omega t + \chi''_1 \sin\omega t) + 3/4 E^3(\chi'_{\bar{3}} \cos\omega t + \chi''_{\bar{3}} \sin\omega t)+... \\
+1/4 E^3(\chi'_3 \cos3\omega t + \chi''_3 \sin3\omega t)+...,
\label{eqchideomegapract2}
\end{multline}
where the susceptibilities $\chi_i$ are given as a function of their real and imaginary parts $\chi_i'$ and $\chi_i''$. For practical applications, the modulii and arguments $\left|\chi_i\right|$ and $\delta_i$ are rather used:
\begin{multline}
P(t)/\epsilon_0 = E\left|\chi_1\right| \cos(\omega t - \delta_1) + 3/4 E^3\left|\chi_{\bar{3}}\right| \cos(\omega t - \delta_{\bar{3}}) + \\
+... + 1/4E^3\left|\chi_3\right| \cos(3\omega t - \delta_3)+...
\label{eqchideomegapract3}
\end{multline}
We see in the rhs in Eqs~\ref{eqchideomegapract1}-\ref{eqchideomegapract3} that the nonlinear susceptibility of interest, namely $\chi_3$ appearing in Eq. 1 of the letter, is \textit{directly given} by the measurement of the third harmonics of the polarisation.
\section{II) Third harmonics for a liquid without glassy correlations.}
We briefly summarize here what can be expected for the third harmonics in the case of a liquid without any glassy correlations. We use the work of D\'ejardin and Kalmykov \cite{SDejardin} which studies the nonlinear dielectric relaxation of an assembly of rigid polar symmetric particles. The particles are assumed to be diluted enough in a non polar solvent, so that any interaction between them can be safely neglected. A strong electric field is applied to the system, which biases the (noninertial) rotational Brownian motion of the polar particles and yields a net dielectric response. This response is expressed by expanding the relaxation functions as a Fourier series in the time domain. Then the infinite hierarchy of recurrence equations for the Fourier components is obtained in terms of a matrix continued fraction.
We focus on the case where the field is purely ac, namely $E(t)=E \cos \omega t$. In the following equations, the subscript ``D" refers to the calculation of D\'ejardin and Kalmykov. By using their Eq. (21), one gets :
\begin{equation}
\frac{\chi_{3,D}(\omega)}{\chi_{3,D}(\omega=0)} = -3 \frac{3-17\omega^2 \tau_D^2+i \omega \tau_D(14-6 \omega^2 \tau_D^2)}{(1+\omega^2 \tau_D^2)(9+4\omega^2 \tau_D^2)(1+9\omega^2 \tau_D^2)} \ ,
\label{chi3normDejardin}
\end{equation}
where $\tau_D$ is the usual Debye time since one finds $\chi''_{1,D} \propto \frac{\omega \tau_D}{1+\omega^2 \tau_D^2}$, i.e. $\chi''_1$ is maximum for the frequency $f_D = 1/(2\pi \tau_D)$. By using Eqs. (7) and (19) of D\'ejardin and Kalmykov, one can calculate all the prefactors and obtain $\chi_{3,D}(\omega =0)$:
\begin{equation}
\chi_{3,D}(\omega=0) = \frac{\epsilon_0 (\Delta \chi_1)^2}{k_BT N_0} \frac{1}{5}\ ,
\end{equation}
where $N_0$ is the volumic concentration of polar particles, and where, as in the letter, $\Delta \chi_1 = \chi'_{1,D}(0)-\chi'_{1,D}(\infty)$. To get a benchmark for the behavior of $\chi_{3, trivial}$ for the `no glassy correlations' case, we set $f_D=f_{\alpha}$ as well as $N_0=1/a^3$ with $a^3$ the molecular volume in glycerol. We thus obtain from the above equations:
\begin{equation}
\chi_{3, trivial}(\omega) = \frac{\epsilon_0 (\Delta \chi_1)^2 a^3}{k_BT} \left(\frac{1}{5}\frac{\chi_{3,D}(\omega)}{\chi_{3,D}(\omega=0)} \right)\ .
\label{eqintermediaire}
\end{equation}
\begin{figure*}
\hskip -11mm
\includegraphics[scale=1.27,angle=0]{FigS1-Crauste-EPAPS}
\caption{\label{figS1} The quantity $\vert \mathnormal{X}_{trivial}\vert = \vert \frac{ \chi_{3, trivial}(\omega) k_BT}{\epsilon_0 (\Delta \chi_1)^2 a^3} \vert$ is plotted as a function of the ratio $f/f_{D}$. Inset: same graph in log-log plot showing that $\mathnormal{X}_{trivial}$ decreases as $f^{-3}$ when $f \gg f_{D}$.}
\end{figure*}
\begin{figure*}
\hskip -11mm
\includegraphics[scale=1.27,angle=0]{FigS2-Crauste-EPAPS}
\caption{\label{figS2} The phase of $\mathnormal{X}_{trivial}$ is plotted as a function of the ratio $f/f_{D}$.}
\end{figure*}
The last term in the rhs of the latter equation depends only on $f/f_D$, i.e. $f/f_{\alpha}$. Thus, the latter equation and Eq. (1) of the letter are formally similar: The quantity $\mathnormal{X}_{trivial} = \frac{ \chi_{3, trivial}(\omega) k_BT}{\epsilon_0 (\Delta \chi_1)^2 a^3}$ plays, for the `no glassy correlations' case, the same role as $\mathnormal{X}= \frac{ \chi_{3}(\omega) k_BT}{\epsilon_0 (\Delta \chi_1)^2 a^3}$ in the main letter. The behavior of $\mathnormal{X}_{trivial}$ is shown on Figs S1-S2. Three main features are noticeable:
(i) On Fig. S1, one sees that $\vert \mathnormal{X}_{trivial} \vert$ has no peak in frequency, at variance with what is observed on Figs (1),(2),(4) of the main letter. This strongly support that, as claimed in the letter, the peak in frequency of $\vert \mathnormal{X} \vert$ is directly linked to the effect of glassy correlations.
(ii) From Eqs. (\ref{chi3normDejardin})-(\ref{eqintermediaire}), one sees that the magnitude of $\vert \mathnormal{X}_{trivial} \vert$ does not depend on the temperature $T$. The $T$ dependence of $\mathnormal{X}_{trivial}$ is only through $\tau_D(T)$, since we have set $f_{D}=f_{\alpha}$. This absence of temperature dependence of $\vert \mathnormal{X}_{trivial} \vert$ strongly supports our claim that the increase of the maximum value of $\vert \mathnormal{X} \vert$ seen on Figs (1),(3) of the letter comes from the variation of the correlation volume when the temperature is decreased towards $T_g$.
(iii) On Fig. S1, one sees that $\vert \mathnormal{X}_{trivial} \vert$ is maximum for $\omega =0$. The trivial part of $\chi_3$ thus dominates over the singular part in the limit of low frequencies $f \ll f_{\alpha}$. Of course the role of the trivial part of $\chi_3$ is larger when the non trivial part is lower, i.e. when the temperature is high. This naturally accounts for the departure to the scaling of $\mathnormal{X}$ as a function of $f/f_{\alpha}$ when both the temperature is high and the frequency $f$ is much lower than $f_{\alpha}$.
\section{III) More information about our experimental setup}
The principle of the experiment was that of a bridge including two
capacitors in which the dielectric layer was the supercooled liquid under study (here, glycerol) as described in section V of ref.~\cite{SThibierge}.
These two glycerol samples with different inter-electrode distances were placed in a closed cell filled with high purity glycerol (purity $> 99.6$ \%) in a dessicated atmosphere to prevent water absorption. The temperature of the cell was regulated by a LakeShore$^{\hbox{\textregistered}}$ temperature controller with a 50 Ohms heater fixed on the cell. The cell and its heater were placed in a cryostat in which the cold stage remained at $\sim$ 20 K and was connected to the cell through a thermal impedance. This device allowed a temperature regulation with a precision of $\sim 50$ mK, and a temperature quench from 300 to 200 K in $\sim$ 1.5 hour.
The electrodes of the two capacitors are $6$ mm thick rods of high purity copper on which $200$ nm of gold was evaporated to prevent the formation of any oxydized layer. A highly specialised surface treatment reduced any surface defects: The final roughness was about 10 nm, while the planeity, i.e. the difference between the highest point of the electrodes and the mean plane, was better than $1\,\mu$m.
Most of the results presented in the letter were obtained with Mylar$^{\hbox{\textregistered}}$ spacers depicted in the letter, giving inter-electrode distances of $19$ $\mu$m and $41$ $\mu$m for the two samples. For completeness, we have added a few results obtained very recently by using spacers made of resin patterned by photolithography (with a thickness $L$ $= 12$ $\mu$m for the thin sample and $25$ $\mu$m for the thick sample).
The bridge was fed by a low distortion voltage source (Stanford Research Systems$^{\hbox{\textregistered}}$ DS360) which gave a rms excitation voltage $\leq$ 14V at a frequency $f$ = $\omega/2\pi$ ranging from 10 mHz to 200 kHz.
The nonlinear susceptibility $\chi_{3}$ was determined after having carefully checked the cubic dependence of the measured signal at $3\omega$ with respect to the excitation at $1\omega$. Last, for a significant fraction of the frequencies and temperatures investigated here, we have checked the quantitative consistency between the $3 \omega$ signal measured in our ``two sample bridge'' and the one found with the ``twin T notch filter'' depicted in Ref.~\cite{SThibierge}, which allows to measure $\chi_3$ by a different method.
\section{IV) Heating effects calculations}
Let us consider a sample made of a supercooled liquid excited by an oscillating field $E \cos(\omega t)$. For small enough {\it E} values, the resulting polarisation $P_{lin}$ is linear and reads :
\begin{equation}\label{S1}
\frac{P_{lin}(t)}{\epsilon_0 E}= \chi_{1}' \cos(\omega t) +\chi_{1}'' \sin(\omega t).
\end{equation}
When $E$ is increased, a part of the nonlinear response comes from the dissipated electrical power, the volumic density of which is $p(t)=\frac{1}{2} \epsilon_0 \chi_{1}'' \omega E^2\left(1+\cos(2 \omega t - \phi) \right)$, with $\phi$ = $-\pi + 2\arctan(\chi_1''/\chi_1')$
(see e.g. Appendix of Ref.~\cite{SRichert08}). The resulting heat propagates towards a ``thermostat'' which in our case can be considered to be the copper electrodes (thickness 6 mm, diameter 20 mm). The resulting average sample temperature increase $\delta T(t)$ is
\begin{equation}\label{S2}
\delta T(t) = \delta T_{0} +\delta T_{2} \cos(2\omega t - \phi_2),
\end{equation}
where the mean dc temperature increase $\delta T_0$ is larger or equal than the ac one $\delta T_2$, thus at any time $\delta T(t) \ge 0$.
$\phi_2$ is a phase shift related to heat transport. As our measurements give the nonlinear dielectric response averaged over the sample volume, $\delta T(t)$ in equation~(\ref{S2}) is the temperature increase averaged over the same volume. Using equations~(\ref{S1})-(\ref{S2}), we thus obtain for the nonlinear part of the susceptibility due to heating effects:
\begin{equation}\label{S3}
\frac{P(t) - P_{lin}(t)}{\epsilon_0 E}= \left( \frac{\partial \chi_{1}'}{\partial T} \delta T(t) \right) \cos(\omega t) +\left( \frac{\partial \chi_{1}''}{\partial T} \delta T(t) \right) \sin(\omega t).
\end{equation}
This is an upper limit on the contribution of heating to the nonlinear response, since we have assumed that $\delta T(t)$ induces instantaneously a modification of the susceptibility. As already advocated in Ref. \cite{SRichert08}, this is questionable, specially in what concerns the contribution of $\delta T_2(t)$ which should be averaged out because of the finite relaxation time $\tau$ of the system (see after eq. \ref{S7}). By using equations~(\ref{S2}-\ref{S3}), one gets :
\begin{eqnarray}
\frac{P(t) - P_{lin}(t)}{\epsilon_0 E} & = & \left[\left(\delta T_0 + \frac{1}{2}\delta T_2 \cos(\phi_2)\right)\frac{\partial \chi_{1}'}{\partial T} + \frac{1}{2}\delta T_2 \sin(\phi_2)\frac{\partial \chi_{1}''}{\partial T}\right] \cos(\omega t)\nonumber \\
& & + \left[\left(\delta T_0 - \frac{1}{2}\delta T_2 \cos(\phi_2)\right)\frac{\partial \chi_{1}''}{\partial T} + \frac{1}{2}\delta T_2 \sin(\phi_2)\frac{\partial \chi_{1}'}{\partial T}\right] \sin(\omega t)\nonumber \\
& & + \left[ \frac{1}{2}\delta T_2 \cos(\phi_2)\frac{\partial \chi_{1}'}{\partial T} - \frac{1}{2}\delta T_2 \sin(\phi_2)\frac{\partial \chi_{1}''}{\partial T}\right] \cos(3\omega t)\nonumber \\
& & + \left[ \frac{1}{2}\delta T_2 \sin(\phi_2)\frac{\partial \chi_{1}'}{\partial T} + \frac{1}{2}\delta T_2 \cos(\phi_2)\frac{\partial \chi_{1}''}{\partial T}\right] \sin(3\omega t).
\label{S4}
\end{eqnarray}
The four terms in the right hand side of equation~(\ref{S4}) give the nonlinear response of the system due to heating that we define in general by
\begin{equation}\label{S5}
\frac{P(t) - P_{lin}(t)}{\epsilon_0 E} = (\delta \chi_{1}')_h \cos(\omega t) + (\delta \chi_{1}'')_h \sin(\omega t) + \frac{E^2}{4} \chi_{3,h}' \cos(3 \omega t)+ \frac{E^2}{4} \chi_{3,h}'' \sin(3 \omega t).
\end{equation}
In the right hand side of equation~(\ref{S5}) each of the four terms are proportional to $E^2$. $\chi_{3,h}'$ and $\chi_{3,h}''$ do not depend on $E$, while $(\delta \chi_{1}')_h$ and $(\delta \chi_{1}'')_h$ are proportional to $E^2$. Our notations have been chosen in order to remain as close as possible to the notations of refs~\cite{SRichert08, SRichert06, SRichert07a}.
We now calculate the expression of $\delta T(t)$ that one has to introduce in equation~(\ref{S4}) in order to obtain the nonlinear response in equation~(\ref{S5}). We just consider the simple heating effect, when the dynamical heterogeneities are not considered:
The supercooled liquid is thus just characterized by its thermal conductivity $\kappa_{th}$ and its total specific heat $c$. As in ref.~\cite{SBirge86}, we consider that $c$ is frequency dependent due to the fact that the slow degrees of freedom cannot contribute to $c$ for frequencies much larger than $f_{\alpha}=1/\tau$. Let us define $(x,y)$ as the plane of our copper electrodes, with $z=0$ for the lower electrode and $z=L$ for the upper one. Due to their very high thermal conductivity and to their large thickness ($6$ mm), the two electrodes can be considered, to a very good approximation, as a thermostat. The temperature increase $\delta \theta (x,y,z,t)$ of the supercooled liquid at point $(x,y,z)$ and time $t$ thus vanishes for $z=0$ and $z=L$. As the diameter $D$ = 2 cm of the electrodes is typically one thousand times larger than $L$, we may consider that $\delta \theta$ does not depend on $(x,y)$. We obtain $\delta \theta(z,t)$ by solving the heat propagation equation:
\begin{equation}\label{S6}
c \frac{\partial \delta \theta(z,t)}{\partial t} = \kappa_{th} \frac{\partial^2 \delta \theta(z,t)}{\partial z^2} + p(t).
\end{equation}
By averaging spatially the solution of equation~(\ref{S6}), we obtain the $\delta T(t)$ to be used in equation~(\ref{S4}):
\begin{eqnarray}
\delta T_0 & = & \frac{\epsilon_{0} \chi_{1}'' \omega E^2 L^2}{24 \kappa_{th}} \nonumber \\
\delta T_{2}(t) & = & \delta T_0 \frac{\cos(2 \omega t-\phi_2)}{\sqrt{1+(2 \omega \tau_{th})^2}},
\label{S7}
\end{eqnarray}
where $\tau_{th}= cL^2/(\kappa_{th} \pi^2)$ and $\phi_2 = \phi + \arctan(2 \omega \tau_{th})$. By using these expressions in equation~(\ref{S4}) and identifying it with equation~(\ref{S5}), we obtain the upper limit to the heating contribution to $\chi_3$ (dotted line in the Fig. 4 of the Letter).
We now move to the problem evoked above, namely the fact that the finite relaxation time $\tau$ of the dipoles
which contribute to the dielectric susceptibility should average out the modification of this susceptibility due to the oscillation $\delta T_{2}(t)$, specially in the case $\omega \tau \ge 1$. As a consequence, the heating contribution to $\chi_3$ should be reduced by a factor $R(\omega \tau) \le 1$. For a precise calculation of $R(\omega \tau)$, one should replace Eq. (\ref{S3}) by a microscopic equation accounting for the dynamics of the dipoles in the case of a thermal bath where the temperature has an oscillating component, which is of great complexity. For a first estimate, we make two very simplifying assumptions:
(i) We assume that the dipoles have a Debye dynamics with a given characteristic time $\tau(T)$. This is a simplifying assumption in the sense that when $T$ is close to $T_g$, it is well known that $\chi_1(\omega)$ is ``stretched'' with respect to a simple Debye law. In fact the Debye dynamics holds only at much higher temperatures, where the molecular motions are independent of each other, which allows to describe the non inertial rotational Brownian motion by the Smoluchowski equation for the probability distribution function of the orientations of the dipoles in configuration space \cite{SDejardin,SLangevin}. After an ensemble averaging of this equation, one gets the well known Debye equation for the dynamics of the average polarisation \cite{SDejardin} :
\begin{equation} \label{S8}
\tau \frac{\partial P}{\partial t} + P = \epsilon_{0} \Delta \chi_1 E \cos(\omega t)
\end{equation}
(ii) we assume that the main effect of $\delta T_2(t)$ is to modulate in time the value of $\tau$ while leaving unchanged the (Debye) dynamics. This can be justified by the fact that the temperature oscillation modulates the viscosity $\eta$, thus the relaxation time $\tau$ which is proportional to $\eta$. Considering the heatings of eq. (\ref{S7}), $\tau(t)$ is now given by
\begin{equation} \label{S9}
\tau(t) = \tau_{lin} + \left(\frac{\partial \tau_{lin}}{\partial T} \right) \delta T_0 + \left(\frac{\partial \tau_{lin}}{\partial T} \right) \delta T_2 (t)\; ,
\end{equation}
where $\tau_{lin}$ is the value of $\tau$ at zero field. In the following, $\delta \tau_2$ will denote the amplitude of the $2 \omega$ modulation of $\tau$ due to $\delta T_2(t)$ and corresponding to the last term of Eq. (\ref{S9}). Of course, using Eq. (\ref{S9}) for $\tau(t)$ assumes that $\delta T_2(t)$ instantaneously fully affects $\tau$. A thorough modelization of this problem could lead to a more involved expression where $\delta \tau_2$ should be weaker than in the above expression. As we shall find that the third harmonics is proportionnal to $\delta \tau_2 /\tau_{lin}$, see below Eq. (\ref{S11}), we are lead to the conclusion that our new estimate should, again, be slightly overstimated.
We now insert $\tau(t)$ in Eq. (\ref{S8}) and set :
\begin{equation} \label{S10}
P(t) = P_{lin} \cos(\omega t - \Psi_{lin}) + \delta P_1 \cos(\omega t - \Psi_1) + P_3 \cos(3\omega t - \Psi_3) + ...
\end{equation}
where $P_{lin}, \Psi_{lin}, \delta P_1, \Psi_1, P_3, \Psi_3$ are to be determined. As we are only interested in the onset of nonlinear effects, $P_{lin}\propto E$ is much larger than $\delta P_1 \propto E^3$ and than $P_3 \propto E^3$. This allows to neglect higher order harmonics (denoted by the dots in Eq. (\ref{S10})) and to resolve Eq. (\ref{S8}) by identification of the terms which have the same frequency and the same power of $E$. This yields:
\begin{equation}\label{S11}
P_3 = \frac{\epsilon_{0} \Delta \chi_1 E}{2}\frac{\delta \tau_2}{\tau_{lin}} \frac{\omega \tau_{lin}}{\sqrt{1+(\omega \tau_{lin})^2}\sqrt{1+(3\omega \tau_{lin})^2}}
\end{equation}
where $\delta \tau_2$ denotes the amplitude of the $2 \omega$ modulation of $\tau$ arising from $\delta T_2(t)$ in the last term of Eq. (\ref{S9}): thus $\delta \tau_2 \propto E^2$, see Eq. (\ref{S7}), which yields the expected $P_3 \propto E^3$.
We now have to compare this result to that obtained if we start from Eq. (\ref{S3}) and use a Debye linear susceptibility. A straightforward calculation shows that in that case the modulus of the third harmonics of the polarisation is given by the expression of Eq. (\ref{S11}) divided by a function $R(\omega \tau)$ given by :
\begin{equation}\label{S12}
R(\omega \tau) = \frac{\sqrt{1+(\omega \tau_{lin})^2}}{\sqrt{1+(3 \omega \tau_{lin})^2}}
\end{equation}
As expected $R(\omega = 0)=1$ and $R(\omega \tau \gg 1)$ is smaller than $1$. We note that $R(\omega \tau \to \infty) =1/3$, which comes from the fact that $\tau(t)$ enters in Eq. (\ref{S8}) as a factor of $\partial P/ \partial t$: this gives a weight $3 \omega \tau$ contrarily to the case where one starts from Eq. (\ref{S3}) where this weight is simply $\omega \tau$. This reduction of the effect of $\delta T_2(t)$ on the polarisation can be seen as a first estimate of the fact that the dipoles average out the temperature oscillations. Of course, one could build a much more thorough model of this effect, but the reduction given by Eq. (\ref{S12}) is enough to ensure that the power law regime reported in the main letter is not affected by the heating contribution calculated here. Finaly, the lengths of the arrows of the Fig. 4 of the main letter are those predicted by Eq. (\ref{S12}).
|
\section{Introduction}\label{s1}
This Festschrift contribution is devoted to a survey of Barry Simon's principal contributions to the area of inverse spectral theory for one-dimensional Schr\"odinger and Jacobi operators. We decided to put the emphasis on the following five groups of topics:
\smallskip
$\bullet$ The Dirichlet spectral deformation method
\smallskip
A general spectral deformation method applicable to Schr\"odinger, Jacobi, and Dirac-type operators in one dimension, which can be used to insert eigenvalues into spectral gaps of arbitrary background operators but is also an ideal technique to construct isospectral (in fact, unitarily equivalent) sets of operators starting from a given base operator.
\smallskip
$\bullet$ Renormalized oscillation theory
\smallskip
Renormalized oscillation theory formulated in terms of Wronskians of appropriate solutions rather than solutions themselves, applies, in particular, to energies above the essential spectrum where real-valued solutions exhibit infinitely many zeros and traditional eigenvalue counting methods would naively lead to $\infty - \infty$. While not directly related to inverse spectral methods, we chose to include this topic because of its fundamental importance to the Dirichlet spectral deformation method.
\smallskip
$\bullet$ The xi function and trace formulas for Schr\"odinger and Jacobi operators
\smallskip
The xi function, that is, essentially, the argument of the diagonal Green's function, which also takes on the role of a particular spectral shift function, is an ideal tool to derive a hierarchy of (higher-order) trace formulas for one-dimensional Schr\"odinger and Jacobi operators. The latter are the natural extensions of the well-known trace formulas for periodic and algebro-geometric finite-band potential coefficients to arbitrary coefficients. The xi function provides a tool for direct and inverse spectral theory.
\smallskip
$\bullet$ Uniqueness theorems in inverse spectral theorem
\smallskip
Starting from the Borg--Marchenko uniqueness theorem, the basic uniqueness result for Schr\"odinger and Jacobi operators in terms of the Weyl--Titchmarsh $m$-coefficient, a number of uniqueness results are discussed. The latter include the Borg-type two-spectra results as well as Hochstadt--Lieberman-type results with mixed prescribed data. In addition to these traditional inverse spectral problems, several new types of inverse spectral problems are addressed.
\smallskip
$\bullet$ Simon's new approach to inverse spectral theory
\smallskip
In some sense, Simon's new approach to inverse spectral theory for half-line problems, based on a particular representation of the Weyl--Titchmarsh $m$-function as a finite Laplace transform with control about the error term, can be viewed as a continuum analog of the continued fraction approach (based on the Riccati equation) to the inverse spectral problem for semi-infinite Jacobi matrices (the actual details, however, differ considerably). Among a variety of spectral-theoretic results, this leads to a formulation of the half-line inverse spectral problem alternative to that of Gel'fand and Levitan. In addition, it leads to a fundamental new result, the local Borg--Marchenko uniqueness theorem.
\medskip
Each individual section focuses on a particular paper or on a group of papers to be surveyed, representing the five items just discussed. Since this survey is fairly long, it was our intention to write each section in such a manner that it can be read independently.
Only self-adjoint Schr\"odinger and Jacobi operators are considered in this survey. In particular, all potential coefficients $V$ and Jacobi matrix coefficients $a$ and $b$ are assumed to be real-valued throughout this survey (although, we occasionally remind the reader of this assumption).
To be sure, this is not a survey of the state of the art of inverse spectral
theory for one-dimensional Schr\"odinger and Jacobi operators. Rather, it exclusively focuses on Barry Simon's contributions to and influence exerted on the field. Especially, the bibliography, although quite long, is far from complete and only reflects the particular purpose of this survey. The references included are typically of the following two kinds: First, background references that were used by Barry Simon and his coworkers in writing a particular paper. Such references are distributed throughout the particular survey of one of his papers. Second, at the end of each such survey we refer to more recent references which complement the results of the particular paper in question.
\bigskip
It was 23 years ago in April of 1983 that Barry and I first met in person at Caltech and started our collaboration. Barry became a mentor and then a friend, and it is fair to say he has had a profound influence on my career since that time. Working with Barry has been exciting and most rewarding for me. Happy Birthday, Barry, and many more such anniversaries!
\section{The Dirichlet Spectral Deformation Method}
\label{s2}
In this section we describe some of the principal results of the paper: \\
\cite{GST96} F.\ Gesztesy, B.\ Simon, and G.\ Teschl, {\it Spectral
deformations of one-dimension- al Schr\"odinger operators}, J. Analyse
Math. {\bf 70}, 267--324 (1996).
\medskip\smallskip
Spectral deformations of Schr\"odinger operators in $L^2 ({\mathbb{R}})$,
isospectral and certain
classes of non-isospectral ones, have attracted a lot of interest over the
past three decades
due to their prominent role in connection with a variety of topics,
including the
Korteweg-de Vries (KdV) hierarchy, inverse spectral problems,
supersymmetric quantum
mechanical models, level comparison theorems, etc. In fact, the
construction of $N$-soliton
solutions of the KdV hierarchy (and more generally, the construction of
solitons relative
to reflectionless backgrounds) is a typical example of a non-isospectral
deformation of
$H=-\frac{d^2}{dx^2}$ in $L^2 ({\mathbb{R}})$ since the resulting deformation
$\tilde H = -\frac{d^2}{dx^2} + \tilde V$ acquires an additional point spectrum
$\{\lambda_1, \dots,\lambda_N\}\subset (-\infty, 0)$, $N\in{\mathbb{N}}$, such that
$$
\sigma(\tilde H) =\sigma (H)\cup \{\lambda_1, \dots, \lambda_N\}
$$
($\sigma(\,\cdot\,)$ abbreviating the spectrum). In the $N$-soliton context
(ignoring the KdV time parameter for simplicity), $\widetilde V$ is of the explicit form
\begin{equation}
\widetilde V(x) = - 2 \frac{d^2}{dx^2} \ln[W(\Psi_1(x),\dots,\Psi_N(x))], \quad x\in{\mathbb{R}}, \label{2.2.2}
\end{equation}
where $W(f_1,\dots,f_N)$ denotes the Wronskian of $f_1,\dots,f_N$ and
the functions $\Psi_j$, $j=1,\dots,N$, are given by
\begin{align*}
& \Psi_j(x)=(-1)^{j+1}e^{-\kappa_j x} +\alpha_j e^{\kappa_j x}, \quad x\in{\mathbb{R}}, \\
& 0<\kappa_1<\cdots<\kappa_N, \quad \alpha_j >0, \; j=1,\dots,N.
\end{align*}
The Wronski-type formula in \eqref{2.2.2} is typical also for general background
potentials and typical for the Crum--Darboux-type commutation approach \cite{Cr55},
\cite{DT79} (cf.\ \cite{GSS91} and the references therein for general backgrounds) which underlies all standard spectral deformation methods for one-dimensional Schr\"odinger operators such as single and double commutation, and the Dirichlet deformation method presented in this section.
On the other hand, the solution of the inverse periodic problem and the corresponding solution of the algebro-geometric quasi-periodic finite-band inverse problem for
the KdV hierarchy (and certain almost-periodic limiting situations
thereof) are intimately connected with isospectral (in fact, unitary)
deformations of a given base (background) operator
$H=-\frac{d^2}{dx^2}+V$. Although not a complete bibliography on
applications of spectral deformations in mathematical physics,
the interested reader may consult \cite{Ba86}, \cite{BS89}, \cite{BGGSS87},
\cite{BF84}, \cite{Cr55}, \cite{De78}, \cite{DT79},
\cite[Sect.\ 4.3]{EK82a}, \cite{EK82},
\cite{EF85}, \cite{FIT87}, \cite{Fi89}, \cite{FM76},
\cite{GGKM74}, \cite{Ge91}, \cite{Ge92}, \cite{GS95},
\cite{Ge93}, \cite{GH00}, \cite[App.\ G]{GH03}, \cite{GSS91},
\cite{GST96}, \cite{GT96}, \cite{GW93}, \cite[Ch.\ 2, App.\ A]{GM97},
\cite{Iw87}, \cite{KM56}, \cite[Sect.\ 6.6]{Le87},
\cite{Mc85}--\cite{MM75}, \cite[Chs.\ 3, 4]{PT87}, \cite{RT88},
\cite{RS85}, \cite{Sc78}, \cite{Sc03}, \cite{Sh86}, \cite{Sh87}, and the
numerous references cited therein.
The main motivation for writing \cite{GST96} originated in our
interest in inverse spectral
problems. As pointed out later (see Remarks \ref{r2.4.5}--\ref{r2.4.8}),
spectral deformation methods can provide crucial insights into the
isospectral class of a given base potential $V$\!, and in some
cases can even determine the whole isospectral class
of such potentials. A particularly interesting open problem in inverse
spectral theory concerns
the characterization of the isospectral class of potentials $V$ with purely
discrete spectra
(e.g., the harmonic oscillator $V(x)=x^2$, cf.\ \cite{Ch03}--\cite{CK05},
\cite{GS04}, \cite{Le88}, \cite{MT82}, \cite{PS05}).
The principal result in \cite{GST96}, reviewed in this section (cf.\
Theorem \ref{t2.4.4}\,$(i)$), provides a complete spectral characterization
of a new method of constructing isospectral (in fact, unitary)
deformations of general Schr\"odinger operators $H=-\frac{d^2}{dx^2}+V$ in
$L^2 ({\mathbb{R}})$. The technique is connected to Dirichlet data, that is,to
the spectrum of the operator $H^D$ on $L^2 ((-\infty, x_0])\oplus L^2
([x_0, \infty))$ with a Dirichlet boundary condition at $x_0$. The
transformation moves a single eigenvalue of $H^D$ and perhaps flips the
half-line (i.e.,
$(-\infty,x_0)$ to $(x_0,\infty)$, or vice versa) to which the Dirichlet
eigenvalue belongs. On the remainder of the spectrum, the transformation
is realized by a unitary operator.
To describe the Dirichlet deformation method (DDM) as developed in
\cite{GST96} in some detail, we suppose that
$
V\in L^1_{\rm loc} ({\mathbb{R}})
$
is real valued and introduce the differential expression
$\tau=-\frac{d^2}{dx^2}+V(x)$, $x\in{\mathbb{R}}$. Assuming $\tau$ to be in the
limit point case at $\pm\infty$ (for the general case we refer to
\cite{GST96}) one defines the self-adjoint base
(i.e., background) operator $H$ in $L^2 ({\mathbb{R}})$ by
\begin{equation}
Hf =\tau f, \quad
f\in\text{\rm{dom}}(H) = \{g\in L^2 ({\mathbb{R}}) \,|\, g,g'\in
AC_{\rm loc}({\mathbb{R}});
\tau g\in L^2({\mathbb{R}})\}. \label{2.2.3}
\end{equation}
Here $W(f,g)(x)=f(x)g'(x)-f'(x)g(x)$ denotes the Wronskian of $f,g\in
AC_{\rm loc}
({\mathbb{R}})$ (the set of locally absolutely continuous functions on ${\mathbb{R}}$).
Given $H$ and a fixed reference point $x_0 \in{\mathbb{R}}$, we introduce
the associated Dirichlet operator $H^D_{x_0}$ in $L^2 ({\mathbb{R}})$ by
\begin{align*}
H^D_{x_0} f = \tau f, \quad
f\in \text{\rm{dom}}(H^D_{x_0}) &= \{g\in L^2 ({\mathbb{R}}) \,|\, g\in
AC_{\rm loc}({\mathbb{R}}), g'\in
AC_{\rm loc} ({\mathbb{R}}\backslash \{x_0\}); \\
& \quad \;\;\, \lim_{\epsilon\downarrow 0}
g(x_0 \pm\epsilon)=0; \, \tau g\in L^2 ({\mathbb{R}})\}.
\end{align*}
Clearly, $H^D_{x_0}$ decomposes into
$
H^D_{x_0} = H^D_{-,x_0} \oplus H^D_{+,x_0}
$
with respect to the orthogonal decomposition
$
L^2 ({\mathbb{R}}) =L^2 ((-\infty, x_0]) \oplus L^2 ([x_0, \infty)).
$
Moreover, for any
$\mu\in\sigma_d (H^D_{x_0})\backslash \sigma(H)$ ($\sigma_d
(\,\cdot\,)$, the discrete spectrum,
$\sigma(\,\cdot\,)$
and $\sigma_{\rm ess}(\,\cdot\,)$, the spectrum and essential
spectrum, respectively),
we introduce the Dirichlet datum
\begin{equation}
(\mu,\sigma) \in\{\sigma_d (H^D_{x_0})\backslash\sigma_d (H)\} \times
\{-,+\},
\label{2.2.7}
\end{equation}
which identifies $\mu$ as a discrete Dirichlet eigenvalue on the interval
$(x_0, \sigma
\infty)$, that is, $\mu\in\sigma_d (H^D_{\sigma,x_0})$, $\sigma\in
\{-,+\}$ (but excludes it
from being simultaneously a Dirichlet eigenvalue on $(x_0,
-\sigma\infty)$).
Next, we pick a fixed spectral gap $(E_0, E_1)$ of $H$, the endpoints of
which (without loss of generality) belong to the spectrum of $H$,
$$
(E_0, E_1) \subset {\mathbb{R}}\backslash\sigma(H), \quad
E_0, E_1 \in\sigma(H)
$$
and choose a discrete eigenvalue $\mu$ of $H^D_{x_0}$ in the closure of
that spectral gap,
\begin{equation}
\mu\in\sigma_d (H^D_{x_0})\cap [E_0, E_1] \label{2.2.9}
\end{equation}
(we note there is at most one such $\mu$ since $(H^D_{x_0} -z)^{-1}$ is a
rank-one perturbation
of $(H-z)^{-1}$). According to \eqref{2.2.7}, this either gives rise to a
Dirichlet datum
\begin{equation}
(\mu,\sigma)\in (E_0, E_1) \times \{-,+\}, \label{2.2.10}
\end{equation}
or else to a discrete eigenvalue of $H^D_{-,x_0}$ and $H^D_{+,x_0}$, that
is,
\begin{equation}
\mu\in\{E_0, E_1\}\cap\sigma_d (H)\cap\sigma_d(H^D_{-,x_0})\cap
\sigma_d (H^D_{+,x_0}) \label{2.2.11}
\end{equation}
since the eigenfunction of $H$ associated with $\mu$ has a zero at $x_0$.
In particular,
since $(H^D_{x_0} -z)^{-1}$ is a rank-one perturbation of $(H-z)^{-1}$,
one infers
$$
\sigma_{\rm ess} (H^D_{x_0}) = \sigma_{\rm ess}(H),
$$
and thus, $\mu\in \{E_0, E_1\}\cap\sigma_{\rm ess}(H)$ is
excluded by assumption \eqref{2.2.9}. Hence, the case distinctions
\eqref{2.2.10} and \eqref{2.2.11} are exhaustive.
In addition to $\mu$ as in \eqref{2.2.9}--\eqref{2.2.11}, we also need to
introduce
$\tilde\mu\in
[E_0, E_1]$ and $\tilde\sigma\in \{-,+\}$ as follows: Either
$$
(\tilde\mu, \sigma)\in (E_0, E_1) \times \{-,+\},
$$
or else
$$
\tilde\mu\in \{E_0, E_1\}\cap\sigma_d (H).
$$
Given $H$, one introduces Weyl--Titchmarsh-type solutions $\psi_\pm (z,x)$
of $(\tau-z)\psi(z)=0$ by
\begin{align*}
& \psi_\pm (z,\,\cdot\,)\in L^2 ((R,\pm\infty)), \quad R\in {\mathbb{R}}, \\
& \lim_{x\to\pm\infty} W(\psi_\pm (z), g)(x)=0 \quad \text{for all
$g\in\text{\rm{dom}}(H)$}.
\end{align*}
If $\psi_\pm (z,x)$ exist, they are unique up to constant multiples. In
particular, $\psi_\pm
(z,x)$ exist for $z\in{\mathbb{C}}\backslash\sigma_{\rm ess}(H)$ and we
can (and
will) assume them to be holomorphic with respect to
$z\in{\mathbb{C}}\backslash\sigma(H)$ and
real-valued for $z\in{\mathbb{R}}$ (cf.\ the discussion in connection with
\eqref{3.1.1}).
Given $\psi_\sigma (\mu, x)$ and $\psi_{-\tilde\sigma}(\tilde\mu, x)$, one
defines
$$
W_{(\tilde\mu, \tilde\sigma)} (x) =
\begin{cases} (\tilde\mu -\mu)^{-1} W(\psi_\sigma (\mu),
\psi_{-\tilde\sigma}(\tilde\mu))(x),
& \mu,\tilde\mu \in [E_0, E_1], \ \tilde\mu \neq \mu, \\
-\sigma \int^x_{\sigma\infty} dx' \psi_\sigma (\mu, x')^2, &
(\tilde\mu, \tilde\sigma)
=(\mu, -\sigma), \ \mu\in (E_0, E_1), \
\end{cases}
$$
and the associated Dirichlet deformation
\begin{align}
& \tilde\tau_{(\tilde\mu, \tilde\sigma)} = -\frac{d^2}{dx^2} +\tilde
V_{(\tilde\mu,
\tilde\sigma)}(x), \notag \\
& \tilde V_{(\tilde\mu, \tilde\sigma)}(x) = V(x)-2\{\ln [W_{(\tilde\mu,
\tilde\sigma)}
(x)]\}'', \quad x\in{\mathbb{R}}, \label{2.2.20} \\
& \mu, \tilde\mu \in [E_0, E_1], \ \mu\neq \tilde\mu
\text{ or } (\tilde\mu, \tilde\sigma) = (\mu, -\sigma), \
\mu\in (E_0, E_1). \notag
\end{align}
As discussed in Section \ref{s3}, $W_{(\tilde\mu, \tilde\sigma)}(x)\neq 0$,
$x\in{\mathbb{R}}$, and hence \eqref{2.2.20} yields a well-defined potential
$
\tilde V_{(\tilde\mu, \tilde\sigma)}\in L^1_{\rm loc} ({\mathbb{R}}).
$
In the remaining cases
$(\tilde\mu,
\tilde\sigma) =
(\mu, \sigma)$, $\mu\in [E_0, E_1]$, and $\mu=\tilde\mu \in\{E_0, E_1 \}
\cap \sigma_d (H)$, we define
$
\tilde V_{(\tilde\mu, \tilde\sigma)} =V
$
which represents the trivial deformation of $V$ (i.e., none at
all), and for notational simplicity these trivial cases are excluded in the
remainder of this section. For obvious
reasons we will allude to \eqref{2.2.20} as the Dirichlet
deformation method in the following.
If $\tilde\mu \in\sigma_d (H)$, then $\psi_- (\tilde\mu)=c\psi_+
(\tilde\mu)$ for some
$c\in{\mathbb{R}}\backslash \{0\}$, showing that $W_{(\tilde\mu,
\tilde\sigma)}(x)$ and hence,
$V_{(\tilde\mu, \tilde\sigma)}(x)$ in \eqref{2.2.20} becomes independent
of $\sigma$
or $\tilde\sigma$. In this case we shall occasionally use a more
appropriate notation
and write $\tilde V_{\tilde\mu}$ and $\tilde\tau_{\tilde\mu}$ (instead of
$\tilde V_{(\tilde\mu, \tilde\sigma)}$ and $\tilde\tau_{(\tilde\mu,
\tilde\sigma)}$).
For later reference, we now summarize our basic assumptions on $V$,
$\mu$, and
$\tilde\mu$ in the following hypothesis.
\begin{hypothesis} \label{h2.2.3} Suppose $V\in L^1_{\rm loc}({\mathbb{R}})$ to
be real-valued. In addition, we assume
\begin{align*}
&(E_0, E_1)\subset{\mathbb{R}}\backslash\sigma(H), \quad E_0, E_1 \in\sigma(H), \\
&\mu\in\sigma_d (H^D_{x_0}), \quad (\mu,\sigma)\in (E_0, E_1)\times
\{-,+\}
\text{ or } \mu\in\{E_0, E_1\}\cap \sigma_d (H), \\
&(\tilde\mu,\tilde\sigma)\in (E_0, E_1)\times\{-,+\} \text{ or }
\tilde\mu \in \{E_0, E_1\}\cap\sigma_d (H), \\
&\mu,\tilde\mu \in [E_0, E_1], \quad \mu\neq\tilde\mu \text{ or }
(\tilde\mu, \tilde\sigma) = (\mu,-\sigma), \quad \mu\in (E_0, E_1).
\end{align*}
\end{hypothesis}
Next, introducing the following solutions of
$(\tilde\tau_{(\tilde\mu,\tilde\sigma)}-z)
\tilde\psi (z)=0$,
\begin{align*}
&\tilde\psi_{-\sigma}(\mu,x) =\psi_{-\tilde\sigma}(\tilde\mu,x) \big/
W_{(\tilde\mu, \tilde\sigma)}(x), \\
&\tilde\psi_{\tilde\sigma}(\tilde\mu, x) = \psi_\sigma (\mu,x) \big/
W_{(\tilde\mu,\tilde\sigma)}(x), \quad \tilde\psi_{\tilde\sigma}(\tilde\mu,
x_0)=0,
\end{align*}
one
infers
$$
(\tilde\tau_{(\tilde\mu,\tilde\sigma)}\tilde\psi_{-\sigma}(\mu))(x)=\mu
\tilde\psi_{-\sigma}(\mu,x),
\quad
(\tilde\tau_{(\tilde\mu,\tilde\sigma)}
\tilde\psi_{\tilde\sigma}(\tilde\mu))(x)=\tilde\mu
\tilde\psi_{\tilde\sigma}
(\tilde\mu, x).
$$
The Dirichlet deformation operator $\tilde H_{(\tilde\mu,\tilde\sigma)}$
associated with
$\tilde\tau_{(\tilde\mu,\tilde\sigma)}$ in \eqref{2.2.20} is then defined
as follows:
\begin{align}
& \tilde H_{(\tilde\mu,\tilde\sigma)}f
=\tilde\tau_{(\tilde\mu,\tilde\sigma)} f, \quad
f\in\text{\rm{dom}}(\tilde H_{(\tilde\mu,\tilde\sigma)}) = \{g\in L^2({\mathbb{R}}) \,|\,
g,g'\in AC_{\rm loc}({\mathbb{R}}); \notag \\
& \hspace*{4.1cm} g \text{ satisfies the b.c. in \eqref{2.2.33};} \,
\tilde\tau_{(\tilde\mu,\tilde\sigma)} g\in L^2 ({\mathbb{R}})\}. \label{2.2.32}
\end{align}
The boundary conditions (b.c.'s) alluded to in \eqref{2.2.32} are chosen as
follows:
\begin{align}
\begin{split}
\lim_{x\to\tilde\sigma \infty}
W(\tilde\psi_{\tilde\sigma}(\tilde\mu), g)(x) &= 0
\text{ if $\tilde\tau_{(\tilde\mu,\tilde\sigma)}$ is l.c.~at $\tilde\sigma
\infty$}, \\
\lim_{x\to -\tilde\sigma\infty}
W(\tilde\psi_{-\sigma}(\mu), g)(x) &=0
\text{ if $\tilde\tau_{(\tilde\mu,\tilde\sigma)}$ is l.c.~at
$-\tilde\sigma\infty$}. \label{2.2.33}
\end{split}
\end{align}
Here we abbreviate the limit point and limit circle cases by l.p. and
l.c., respectively. As usual, the boundary condition at $\omega\infty$ in
\eqref{2.2.32} is omitted if
$\tilde\tau_{(\tilde\mu,\tilde\sigma)}$ is l.p.~at $\omega\infty$,
$\omega\in\{-,+\}$.
For future reference we note that in analogy to the Dirichlet operators
$H^D_{x_0}$, $H^D_{\pm,x_0}$ introduced in connection with the operator
$H$, one can also introduce the corresponding Dirichlet operators
$\tilde H^D_{(\tilde\mu,\tilde\sigma),x_0}$,
$\tilde H^D_{(\tilde\mu,\tilde\sigma),\pm,x_0}$ associated
with $\tilde H_{(\tilde\mu,\tilde\sigma)}$.
Next, we turn to the Weyl--Titchmarsh $m$-functions for the
Dirichlet deformation
operator $\tilde H_{(\tilde\mu,\tilde\sigma)}$ and relate them to those of
$H$.
Let $\phi(z,x),\theta(z,x)$ be the standard fundamental system of
solutions of
$(\tau -z)\psi(z)=0$, $z\in{\mathbb{C}}$ defined by
$\phi(z,x_0)=\theta'(z,x_0)=0$, $\phi'(z,x_0)=\theta(z,x_0)=1$,
$z\in{\mathbb{C}}$, and denote by
$\tilde\theta_{(\tilde\mu,\tilde\sigma)}(z,x),
\tilde\phi_{(\tilde\mu,\tilde\sigma)}(z,x)$
the analogously normalized fundamental system of solutions of
$(\tilde\tau_{(\tilde\mu,\tilde\sigma)}-z)
\tilde\psi(z)=0$, $z\in{\mathbb{C}}$, at $x_0$. One then has
$$
m_\sigma (z,x_0)=\psi'_\sigma (z,x_0)/\psi_\sigma (z,x_0) , \quad
\sigma\in\{-,+\}, \
z\in{\mathbb{C}}\backslash{\mathbb{R}},
$$
where $m_\sigma(\cdot,x_0)$ denotes the Weyl--Titchmarsh $m$-function of
$H$ with respect to the half-line $(x_0, \sigma\infty)$,
$\sigma\in\{-,+\}$. For the corresponding half-line Weyl--Titchmarsh
$m$-functions of $\tilde H_{(\tilde\mu, \tilde\sigma)}$ in terms of those
of $H$ one then obtains the following result.
\begin{theorem} \label{t2.3.2} Assume Hypothesis \ref{h2.2.3} and
$z\in{\mathbb{C}}\backslash{\mathbb{R}}$. Let $H$ and $\tilde H_{(\tilde\mu,
\tilde\sigma)}$ be given by \eqref{2.2.3} and \eqref{2.2.32},
respectively, and denote by $m_\pm$
and $\tilde m_{(\tilde\mu, \tilde\sigma), \pm}$ the corresponding
$m$-functions
associated with the half-lines $(x_0, \pm\infty)$. Then,
\begin{align*}
& \tilde m_{(\tilde\mu, \tilde\sigma), \pm}(z,x_0) =
\frac{z-\mu}{z-\tilde\mu}\, m_\pm (z,x_0)
-\frac{\tilde\mu-\mu}{z-\tilde\mu}\, m_{-\tilde\sigma}(\tilde\mu,x_0),
\quad
\tilde\mu\neq\mu, \\
& \tilde m_{(\tilde\mu, \tilde\sigma), \pm}(z,x_0) = m_\pm (z,x_0) -
\biggl(\,\int^{x_0}_{\sigma
\infty}dx\, \phi(\mu,x)^2\biggr)^{-1} \frac{1}{z-\mu}, \quad (\tilde\mu,
\tilde\sigma) =
(\mu, -\sigma).
\end{align*}
\end{theorem}
Given the fundamental relation between $\tilde m_{(\tilde\mu,
\tilde\sigma),\pm}$
and $m_\pm$ in Theorem \ref{t2.3.2}, one can now readily derive the
ensuing relation between
the corresponding spectral functions $\tilde\rho_{(\tilde\mu,
\tilde\sigma),\pm}$
and $\rho_\pm$ associated with the half-line Dirichlet operators
$\tilde H^D_{(\tilde\mu, \tilde\sigma),\pm,x_0}$ and $H^D_{\pm,x_0}$. For
this and a complete spectral characterization of
$\tilde H^D_{(\tilde\mu,\tilde\sigma), \pm,x_0}$ in terms of
$H^D_{\pm,x_0}$ we refer to \cite{GST96}.
Next we turn to the principal results of \cite{GST96} including explicit
computations of the
Weyl--Titchmarsh and spectral matrices of $\tilde H_{(\tilde\mu,
\tilde\sigma)}$ in terms
of those of $H$ and a complete spectral
characterization of
$\tilde H_{(\tilde\mu, \tilde\sigma)}$ and $\tilde H^D_{(\tilde\mu,
\tilde\sigma),x_0}$
in terms of $H$ and $H^D_{x_0}$.
We start with the Weyl--Titchmarsh matrices for $H$ and
$\tilde H_{(\tilde\mu,
\tilde\sigma)}$. To fix notation, we introduce the Weyl--Titchmarsh
$M$-matrix in
${\mathbb{C}}^2$ associated with $H$ by
\begin{align*}
&M(z,x_0) = (M_{p, q}(z,x_0))_{1\leq p, q\leq 2}
= [m_- (z,x_0) -m_+ (z,x_0)]^{-1} \\
& \quad \times \begin{pmatrix} m_- (z,x_0) m_+(z,x_0)
& [m_- (z,x_0)+m_+ (z,x_0)]/2 \\
[m_- (z,x_0)+m_+ (z,x_0)]/2 & 1 \end{pmatrix} , \quad
z\in{\mathbb{C}}\backslash{\mathbb{R}},
\end{align*}
and similarly $\tilde M_{(\tilde\mu, \tilde\sigma)}$ in connection with
$\tilde H_{(\tilde\mu,\tilde\sigma)}$.
An application of Theorem \ref{t2.3.2} then yields
\begin{theorem} \label{t2.4.1} Assume Hypothesis \ref{h2.2.3} and
$z\in{\mathbb{C}}\backslash{\mathbb{R}}$.Let $H$ and $\tilde H_{(\tilde\mu,
\tilde\sigma)}$ be given by {\rm{(2.3)}} and {\rm{(2.32)}},
respectively. Then the corresponding Weyl--Titchmarsh-matrices $M$ and
$\tilde M_{(\tilde\mu, \tilde\sigma)}$ are related by
\begin{align*}
\tilde M_{(\tilde\mu, \tilde\sigma), 1,1}(z,x_0) &=
\frac{z-\mu}{z-\tilde\mu}\, M_{1,1}(z,x_0)
- 2\frac{\tilde\mu-\mu}{z-\tilde\mu}\,
m_{-\tilde\sigma}(\tilde\mu,x_0)M_{1,2}(z,x_0) \\
&\quad + \frac {(\tilde\mu-\mu)^2}{(z-\mu)(z-\tilde\mu)}\,
m_{-\tilde\sigma} (\tilde\mu,x_0)^2 M_{2,2}(z,x_0), \\
\tilde M_{(\tilde\mu, \tilde\sigma), 1,2}(z,x_0) &= M_{1,2}(z,x_0))
-\frac{\tilde\mu-\mu}{z-\mu}\,
m_{-\tilde\sigma}(\tilde\mu,x_0) M_{2,2}(z,x_0), \\
\tilde M_{(\tilde\mu, \tilde\sigma),2,2}(z,x_0) &=
\frac{z-\tilde\mu}{z-\mu}\, M_{2,2}(z,x_0), \quad \tilde\mu\neq\mu.
\end{align*}
\end{theorem}
Given the basic connection between $\tilde M_{(\tilde\mu,
\tilde\sigma)}$ and $M$
in Theorem \ref{t2.4.1}, one can now proceed to derive the analogous
relations between the spectral
matrices $\tilde\rho_{(\tilde\mu, \tilde\sigma)}$ and
$\rho$ associated with
$\tilde H_{(\tilde\mu, \tilde\sigma)}$ and $H$, respectively (cf.\
\cite{GST96} for details).
The principal spectral deformation result of \cite{GST96} then reads as
follows.
\begin{theorem} \label{t2.4.4} Assume Hypothesis \ref{h2.2.3}. Then, \\
$(i)$ Suppose $\mu,\tilde\mu\in (E_0, E_1)$. Then $\tilde
H_{(\tilde\mu, \tilde\sigma)}$
and $H$ are unitarily equivalent. Moreover, $\tilde H^D_{(\tilde\mu,
\tilde\sigma),x_0}$ and
$H^D_{x_0}$, restricted to the orthogonal complements of the
one-dimensional eigenspaces
corresponding to $\tilde\mu$ and $\mu$, are unitarily equivalent. \\
$(ii)$ Assume $\mu\in\{E_0, E_1\}\cap\sigma_d (H)$, $\tilde\mu\in
(E_0, E_1)$. Then,
$$
\sigma_{(p)}(\tilde H_{(\tilde\mu, \tilde\sigma)})
=\sigma_{(p)}(H)\backslash
\{\mu\}, \quad
\sigma_{(p)}( \tilde H^D_{(\tilde\mu, \tilde\sigma),x_0})
= \{\sigma_{(p)}(H^D_{x_0})
\backslash \{\mu\}\}\cup\{\tilde\mu\}.
$$
$(iii)$ Suppose $\mu\in
(E_0, E_1)$, $\tilde\mu\in \{E_0, E_1\}\cap
\sigma_d (H)$.
Then,
$$
\sigma_{(p)}(\tilde H_{\tilde\mu}) =
\sigma_{(p)}(H)\backslash
\{\tilde\mu\}, \quad
\sigma_{(p)}(\tilde
H^D_{\tilde\mu,x_0}) =\sigma_{(p)}(H^D_{x_0})\backslash\{\mu\},
\quad
\tilde\mu\notin\sigma(\tilde H^D_{\tilde\mu,x_0}).
$$
$(iv)$ Assume
$\mu, \tilde\mu\in \{E_0, E_1\}\cap
\sigma_d (H)$, $\mu
\neq\tilde\mu$. Then,
$$
\sigma_{(p)}(\tilde H_{\tilde\mu}) =
\sigma_{(p)}(H)\backslash\{E_0, E_1\},
\quad
\sigma_{(p)}(\tilde
H^D_{\tilde\mu,x_0}) = \sigma_{(p)}(H^D_{x_0})\backslash \{\mu\},
\quad
\tilde\mu\notin\sigma (\tilde H^D_{\tilde\mu,x_0}).
$$
In cases
$(ii)$--$(iv)$, the corresponding pairs of operators,
restricted to
the obvious orthogonal complements of the eigenspaces
corresponding
to $\mu$ and/or $\tilde\mu$, are unitarily
equivalent. In
particular,
$$
\sigma_{\rm ess, ac, sc}(\tilde H_{(\tilde\mu,
\tilde\sigma)}) = \sigma_{\rm ess, ac, sc}(\tilde
H^D_{(\tilde\mu,\tilde\sigma),x_0}) = \sigma_{\rm ess, ac,
sc}(H^D_{x_0}) = \sigma_{\rm ess, ac, sc}(H).
$$
\end{theorem}
\begin{remark}
\label{r2.4.5} $(i)$ Perhaps the most important consequence of
Theorem
\ref{t2.4.4} (i),
from an inverse spectral point of view, is the fact
that {\it{any}}
finite
number of
deformations of Dirichlet data
within spectral gaps of $V$ yields a
potential $\tilde V$
in the
isospectral class of $V$\!. No further constraints on
$(\mu_j,
\sigma_j),
(\tilde\mu_j, \tilde\sigma_j)$, other than
$(\mu_j, \sigma_j),
(\tilde\mu_j, \tilde\sigma_j)
\in (E_{j-1}, E_j)
\times \{-,+\}$, $(E_{j-1}, E_j)\in{\mathbb{R}}\backslash
\sigma (H)$,
$j=1,
\dots, N$, $N\in{\mathbb{N}}$, are involved. \\
$(ii)$ The isospectral
property $(i)$ in Theorem
\ref{t2.4.4}, in the special case
of
periodic potentials, was first proved by Finkel, Isaacson,
and
Trubowitz \cite{FIT87}. Further results can be found in Buys
and
Finkel \cite{BF84} and Iwasaki \cite{Iw87} (see also
\cite{De78},
\cite{DT79}, \cite{Mc85},
\cite{Mc86}, \cite{Mc87},
\cite{Mc92}). Similar constructions for
Schr\"odinger operators on a
compact interval can be found in P\"oschel and
Trubowitz \cite[Chs.\
3, 4]{PT87} and Ralston and Trubowitz \cite{RT88}. \\
$(iii)$ Let
$\mu\in (E_0, E_1)$. Then the (isospectral) Dirichlet
deformation
$(\mu,\sigma)
\to (\mu, -\sigma)$ is precisely the isospectral case
of the double commutation method (cf.\ \cite{Ge93}, \cite{GH00},
\cite[App.\ B]{GST96},
\cite{GT96}). It simply flips the Dirichlet
eigenvalue
$\mu$ on the half-line $(x_0, \sigma\infty)$ to the other
half-line $(x_0,
-\sigma\infty)$.
In the special case where $V(x)$ is periodic, this procedure was first
used by McKean and van Moerbeke \cite{MM75}. \\
$(iv)$ The topology of these Dirichlet data strongly depends on the nature of the
endpoints $E_0, E_1$ of a particular spectral gap. For instance, in cases like the periodic one, different spectral gaps are separated by intervals of absolutely continuous spectrum and the two intervals $[E_0,E_1]$ together with $\sigma\in \{-,+\}$ can be identified with a circle (upon identifying the two intervals as two rims of a cut). Globally this then leads to a product of circles, that is, a torus. In particular, the Dirichlet eigenvalues in different spectral gaps can be prescribed independently of each other. The situation is entirely different if an endpoint, say $E_1$, belongs to the discrete spectrum of $H$. In this case there are two neighboring spectral gaps $(E_0,E_1)$ and
$(E_1,E_2)$ and the two Dirichlet eigenvalues $\mu_j \in (E_{j-1},E_j)$, $j=1,2$, are not independent of each other. In fact, if one of $\mu_1$ or $\mu_2$ approaches $E_1$, then necessarily so does the other. The topology is then not a product of circles. For instance, a closer analysis of the case of $N$-soliton potentials in \eqref{2.2.2} then illustrates that the appropriate coordinates parametrizing the $N$-soliton isospectral class are $\alpha_j \in (0,\infty)$ (compare also with positive norming constants), which results globally in a product of open half-lines.
\end{remark}
\begin{remark} \label{r2.4.7} In certain cases where the base
(background) potential $V$ is reflectionless (see, e.g., \cite{GY06})
and $H$ is bounded from below and has no singularly continuous
spectrum, the isospectral class $\text{Iso}(V)$ of $V$ (the set of all reflectionless
$\tilde V$'s such that $\sigma(\tilde H)=\sigma(H)$) is completely
characterized by the distribution of Dirichlet
(initial) data $(\mu_{j+1}(x_0), \sigma_{j+1}(x_0))\in [E_j, E_{j+1}]
\times \{-,+\}$, $j\in J$, in nontrivial spectral gaps of $H$. Here
$x_0\in{\mathbb{R}}$ is a fixed reference point and $J=\{0,1,\dots, N-1\}$,
$N\in{\mathbb{N}}$, or $j\in J={\mathbb{N}}_0$ ($={\mathbb{N}}\cup\{0\}$) parametrizes these
nontrivial spectral gaps $(E_j, E_{j+1})$ of $H$ (the trivial one being
$(-\infty, \inf \sigma(H))$). Prime examples of this type are periodic
potentials, algebro-geometric quasi-periodic finite-band potentials (cf.\
\cite[Ch.\ 3]{BBEIM94}, \cite[Ch.\ 1]{GH03}, \cite[Chs.\ 8--12]{Le87},
\cite[Ch.\ 4]{Ma86}, \cite[Ch.\ II]{NMPZ84}), and certain limiting cases
thereof (e.g., soliton potentials). In these cases, an iteration of the
Dirichlet deformation method, in the sense that
$(\mu_{j+1}(x_0), \sigma_{j+1} (x_0))\to (\tilde\mu_{j+1}(x_0),
\tilde\sigma_{j+1}(x_0))$ within $[E_j, E_{j+1}]\times \{-, +\}$ for each
$j\in J$, independently of each other, yields an explicit realization of
the underlying isospectral class $\text{Iso}(V)$ of reflectionless potentials with base
$V$\!. In the periodic case, this was first proved by Finkel,
Isaacson, and Trubowitz \cite{FIT87} (see also \cite{BF84}, \cite{Iw87}).
More precisely, the inclusion of limiting cases
$\mu_{j+1}(x_0)\in \{E _j, E_{j+1}\} \cap \sigma_{\rm ess}(H)$ requires a
special argument (since it is excluded by Hypothesis \ref{h2.2.3})
but this can be provided in the special cases at hand.
\end{remark}
\begin{remark} \label{r2.4.8} Another case of primary interest
concerns
potentials
$V$ with
purely discrete spectra bounded from
below, that is,
$$
\sigma(H)=\sigma_d(H) = \{E_j\}_{j\in{\mathbb{N}}_0},
\quad -\infty< E_0, \
E_j <E _{j+1}, \ j\in{\mathbb{N}}_0, \quad
\sigma_{\rm ess}(H)=\emptyset.
$$
(For simplicity, one may think in terms of the harmonic oscillator
$V(x)=x^2$, \cite{Ch03}--\cite{CK05}, \cite{GS04}, \cite{Le88}, \cite{MT82},
\cite{PS05}.) In this case, either
$$
(\mu_{j+1}(x_0), \sigma_{j+1}(x_0)) \in (E_j, E_{j+1}) \times \{-, +\}
\, \text{ or } \,
\mu_{j+1}(x_0) =E _{j+1} =\mu_{j+2} (x_0),
$$
that is, Dirichlet eigenvalues necessarily meet in pairs whenever they hit
an eigenvalue
of $H$. The following trace formula for $V$ in terms of $\sigma(H)=
\{E _j\}_{j\in{\mathbb{N}}_0}$ and $\sigma (H^D_x)=\{\mu_j (x)\}_{j\in{\mathbb{N}}}$
(with $H^D_y$ the Dirichlet operator associated with
$\tau=-\frac{d^2}{dx^2} +V(x)$ and a Dirichlet boundary condition at
$x=y$), proved in \cite{GS96} (cf.\ Section \ref{s4}),
\begin{equation}
V(x)=E _0 +\lim_{\alpha\downarrow 0} \alpha^{-1} \sum^{\infty}_{j=1}
\Big(2e^{-\alpha\mu_j (x)} - e^{-\alpha E _j} - e^{-\alpha E _{j+1}}\Big),
\label{2.4.23}
\end{equation}
then shows one crucial difference to the periodic-type cases mentioned
previously. Unlike
in the periodic case, though, the initial Dirichlet eigenvalues
$\mu_{j+1}(x_0)$ {\it{cannot}}
be prescribed arbitrarily in the spectral gaps $(E _j, E_{j+1})$ of $H$.
Indeed, the fact
that the Abelian regularization in the trace formula \eqref{2.4.23} for
$V(x)$ converges to a limit
restricts the asymptotic distribution of $\mu_{j+1}(x)\in [E_j, E_{j+1}]$
as $j\to\infty$. For instance, consider $V(x)=x^2-1$, then $E_j=2j$, $j\in{\mathbb{N}}_0$.
The choice $\mu_j(x_0)=2j-\gamma$, $\gamma\in (0,1)$, then yields for the
Abelian regularization on the right-hand side of \eqref{2.4.23},
$$
\lim_{\alpha\downarrow 0} \alpha^{-1} \sum_{j=1}^\infty
\Big(2 e^{-\alpha(2j-\gamma)-e^{-\alpha 2j}-e^{-\alpha(2j-2)}} \Big)
= \lim_{\alpha\downarrow 0} 2[(\gamma -1)+O(\alpha)] \frac{1}{1-e^{-2\alpha}} =\infty.
$$
Put differently, our choice of $\mu_j(x_0)=2j-\gamma$, $\gamma\in (0,1)$,
was not an admissible choice of Dirichlet eigenvalues for the (shifted) harmonic oscillator potential $V(x)=x^2-1$.
However, as stressed in Remark \ref{r2.4.5}\,$(i)$, one of the fundamental
consequences of \cite{GST96}
concerns the fact that there is no such restriction for any finite number
of spectral gaps of $H$. In other words, only the tail end of the Dirichlet
eigenvalues
$\mu_{j+1}(x_0)$ as $j\to\infty$ is restricted (the precise nature of this
restriction being
unknown at this point), any finite number of them can be placed arbitrarily
in the spectral
gaps $(E _j, E_{j+1})$ (with the obvious ``crossing'' restrictions at the
common boundary
$E _{j+1}$ of $(E _j, E_{j+1})$ and $(E _{j+1}, E_{j+2})$). The only other
known restriction to date on Dirichlet initial data $(\mu_j
(x_0), \sigma_j (x_0))$ is that $\sigma_j (x_0) =-$ and $\sigma_j (x_0)=+$
infinitely often, that is,
both half-lines $(-\infty, x_0)$ and $(x_0, \infty)$ support (naturally)
infinitely many Dirichlet eigenvalues.
\end{remark}
For various extensions of the results presented, including a careful
discussion of limit point/limit circle properties of the Dirichlet
deformation operators, iterations of DDM to insert finitely many
eigenvalues in spectral gaps, applications to reflectionless
Schr\"odinger operators, general
Sturm--Liouville operators in a weighted $L^2$-space, applications to
short-range scattering theory, and a concise summary of single and double
commutation methods, we refer to \cite{GST96}.
\medskip
{\it More recent references:} An interesting refinement of Theorem \ref{t2.4.4}\,$(i)$, in which a unitary operator relating $\tilde H_{(\tilde\mu, \tilde\sigma)}$ and $H$ is
explicitly characterized, is due to Schmincke \cite{Sc03}. DDM
for one-dimensional Jacobi and Dirac-type operators has been worked out by Teschl \cite{Te97}, \cite[Ch.\ 11]{Te00} (see also \cite{GT96a}, \cite{Te99}), \cite{Te98a}.
\section{Renormalized Oscillation Theory} \label{s3}
In this section we summarize some of the principal results of
the paper: \\
\cite{GST96a} F.\ Gesztesy, B.\ Simon, and G.\ Teschl, {\it Zeros of the
Wronskian and renormalized oscillation theory}, Amer. J. Math. {\bf 118},
571--594 (1996).
\medskip\smallskip
For over a hundred and fifty years, oscillation theorems for
second-order differential equations have fascinated mathematicians.
Originating with Sturm's celebrated memoir \cite{St36}, extended in a
variety of ways by B\^ocher \cite{Bo17} and others, a large body of
material has been accumulated since then (thorough treatments can be found,
e.g., in \cite{Co71}, \cite{Kr73}, \cite{Re80}, \cite{Sw68}, and the
references therein). In \cite{GST96a} a new wrinkle to oscillation theory
was added by showing that zeros of Wronskians can be used to count
eigenvalues in situations where a naive use of oscillation theory would
give $\infty -\infty$ (i.e., Wronskians lead to renormalized oscillation
theory). In a nutshell, we will show the following result for general
Sturm--Liouville operators $H$ in $L^2((a,b);rdx)$ with separated boundary
conditions at $a$ and $b$ in this section: If $E_{1,2}\in{\mathbb{R}}$ and if
$u_{1,2}$ solve the differential equation $Hu_j=E_j u_j$, $j=1,2$ and
respectively satisfy the boundary condition on the left/right, then the
dimension of the spectral projection $P_{(E_1, E_2)}(H)$ of $H$ equals the
number of zeros of the Wronskian of $u_1$ and $u_2$.
The main motivation in writing \cite{GST96a} had its origins in attempts
to provide a general construction of isospectral potentials for
one-dimensional Schr\"odinger operators following previous work
by Finkel, Isaacson, and Trubowitz \cite{FIT87} (see also \cite{BF84}) in
the special case of periodic potentials. In fact, in the case of periodic
Schr\"odinger operators $H$, the nonvanishing of $W(u_1, u_2)(x)$ for
Floquet solutions $u_1 = \psi_{\varepsilon_1}(E_1)$,
$u_2 = \psi_{\varepsilon_2}(E_2)$, \;
$\varepsilon_{1,2} \in \{+,-\}$ of $H_p$, for $E_1$ and $E_2$ in the
same spectral gap of $H$, is proved in \cite{FIT87}. The extension of these
ideas to general one-dimensional Schr\"odinger operators was done in
\cite{GST96} and is reviewed in Section \ref{s2} of this survey. So while
\cite{GST96a} is not directly related to the overarching inverse spectral theory
topic of this survey, we decided to include it because of its relevance in
connection with Section \ref{s2}.
To set the stage, we consider Sturm--Liouville differential expressions
of the form
$$
(\tau u)(x)=r(x)^{-1}[-(p(x)u'(x))' + q(x)u(x)], \quad x\in(a,b), \quad
-\infty\leq a<b\leq
\infty
$$
where
\begin{equation*}
r, p^{-1}, q \in L^1_{\text{\rm{loc}}}((a,b)) \text{ are real-valued
and $r,p >0$ a.e.~on $(a,b)$}.
\end{equation*}
We shall use $\tau$ to describe the formal differentiation expression and
$H$ for the operator in $L^{2}((a,b); r\,dx)$ given by $\tau$ with separated
boundary conditions at $a$ and/or $b$.
If $a$ (resp., $b$) is finite and $q,p^{-1},r$ are in addition integrable
near $a$ (resp., $b$), $a$ (resp., $b$) is called a {\it{regular}}
end point. $\tau$, respectively $H$, is called {\it{regular}} if
both $a$ and $b$ are regular. As is usual (\cite[Sect.\ XIII.2]{DS88},
\cite[Sect.\ 17]{Na68}, \cite[Ch.\ 3]{We87}), we consider the local
domain
\begin{equation*}
D_{\text{\rm{loc}}}=\{u \in AC_{\text{\rm{loc}}}((a,b)) \,|\, pu'\in
AC_{\text{\rm{loc}}}((a,b)),\ \tau u \in L^{2}_{\text{\rm{loc}}}((a,b); r\,dx)\},
\end{equation*}
where $AC_{\text{\rm{loc}}}((a,b))$ is the set of locally
absolutely continuous functions on $(a,b)$. General ODE
theory shows that for any $E\in{\mathbb{C}}$, $x_0\in (a,b)$, and $(\alpha,
\beta)\in{\mathbb{C}}^2$, there is a unique $u\in D_{\text{\rm{loc}}}$
such that $-(pu')'+qu - Eru=0$ for a.e.~$x \in (a,b)$ and $(u(x_0),
(pu')(x_0))=(\alpha,\beta)$.
The maximal and minimal operators are defined by taking
$$
\text{\rm{dom}}(T_{\rm max})=\{u\in L^{2}((a,b); r\,dx)\cap
D_{\text{\rm{loc}}}\,|\, \tau u\in L^{2}((a,b); r\,dx)\},
$$
with
\begin{equation*}
T_{\rm max}u=\tau u.
\end{equation*}
$T_{\rm min}$ is the operator closure of
$T_{\rm max}\restriction D_{\text{\rm{loc}}}\cap\{u
\text{ has compact support in $(a,b)$}\}$. Then $T_{\rm min}$
is symmetric and $T^{*}_{\rm min}=T_{\rm max}$.
According to Weyl's theory of self-adjoint extensions (\cite[Sect.\
XIII.6]{DS88}, \cite[Sect.\ 18]{Na68}, \cite[App.\ to X.1]{RS75},
\cite[Section 8.4]{We80}, \cite[Chs.\ 4, 5]{We87}), the deficiency indices
of $T_{\rm min}$ are $(0,0)$ or $(1,1)$ or $(2,2)$ depending
on whether it is limit point at both, one, or neither endpoint.
Moreover, the self-adjoint extensions can be described in terms of
Wronskians (\cite[Sect.\ XIII.2]{DS88}, \cite[Sects.\ 17, 18]{Na68},
\cite[Sect.\ 8.4]{We80}, \cite[Ch.\ 3]{We87}). Define
\begin{equation*}
W(u_1,u_2)(x)=u_1(x)(pu_2')(x)-(pu_1')(x)u_2(x).
\end{equation*}
Then if $T_{\rm min}$ is limit point at both ends,
$T_{\rm min}=T_{\rm max}=H$. If $T_{\rm min}$
is limit point at $b$ but not at $a$, for $H$ any self-adjoint extension
of $T_{\rm min}$, if $\varphi_-$ is any function in $\text{\rm{dom}}(H)
\backslash \text{\rm{dom}}(T_{\rm min})$, then
$$
\text{\rm{dom}}(H)=\{u\in \text{\rm{dom}}(T_{\rm max})\,|\, W(u,\varphi_-)(x)\to 0
\text{ as $x\downarrow a$}\}.
$$
Finally, if $u_1$ is limit circle at both ends, the operators $H$ with
separated boundary conditions are those for which we can find
$\varphi_\pm \in \text{\rm{dom}}(H)$, $\varphi_+= 0$ near $a$, $\varphi_-
= 0$ near $b$, and $\varphi_\pm \in \text{\rm{dom}}(H)\backslash
\text{\rm{dom}}(T_{\rm min})$. In that case,
$$
\text{\rm{dom}}(H)=\{u\in D(T_{\rm max}) \,|\, W(u,\varphi_-)(x)\to 0
\text{ as $x\downarrow a$}, \,
W(u,\varphi_+)(x)\to 0 \text{ as $x\uparrow b$}\}.
$$
Of course, if $H$ is regular, we can just specify the boundary
conditions by taking values at $a,b$ since by regularity any $u\in
\text{\rm{dom}}(T_{\rm max})$ has $u,pu'$ continuous on $[a,b]$.
It follows from this analysis that
$$
\text{if $u_{1,2}\in \text{\rm{dom}}(H)$, then $W(u_1,u_2)(x)
\to 0$ as $x\to a$ or $b$.}
$$
Such operators will be called SL operators (for Sturm--Liouville, but SL
includes separated boundary conditions (if necessary)) and denoted by $H$.
It will be convenient to write $\ell_-=a$, $\ell_+=b$.
Throughout this section we will denote by $\psi_{\pm}(z,x) \in
D_{\text{\rm{loc}}}$ solutions of $\tau \psi = z \psi$ so that
$\psi_{\pm}(z,\,.\,)$ is $L^2$ at $\ell_{\pm}$ and $\psi_{\pm}
(z,\,.\,)$ satisfies the appropriate boundary condition at $\ell_\pm$
in the sense that for any $u \in \text{\rm{dom}}(H)$, $\lim_{x\to\ell_{\pm}}
W(\psi_{\pm}(z),u)(x)=0$. If $\psi_{\pm}(z,\,.\,)$ exist, they are
unique up to constant multiples. In particular, $\psi_{\pm}(z,\,.\,)$
exist for $z$ not in the essential spectrum of $H$ and we can assume
them to be holomorphic with respect to $z$ in ${{\mathbb{C}}} \backslash
\sigma (H)$ and real for $z\in{\mathbb{R}}$. One can choose
\begin{equation}
\psi_\pm(z,x) = ((H-z)^{-1} \chi_{(c,d)})(x) \,
\text{ for } \, x<c \, \text{ and } x> d, \quad a<c<d<b \label{3.1.1}
\end{equation}
and uniquely continue $\psi_\pm(z,x)$ for $x > c$ and $x<d$.
Here $(H-z)^{-1}$ denotes the resolvent of $H$ and $\chi_\Omega$ the
characteristic function of the set $\Omega \subseteq {\mathbb{R}}$. Clearly
we can include a finite number of isolated eigenvalues in the domain
of holomorphy of $\psi_\pm$ by removing the corresponding poles. Moreover,
to simplify notations, all solutions $u$ of $\tau u = Eu$ are understood
to be not identically vanishing and solutions associated with real values
of the spectral parameter $E$ are assumed to be real-valued in this paper.
Thus if $E$ is real and in the resolvent set for $H$ or an isolated
eigenvalue, we are guaranteed there are solutions that satisfy the boundary
conditions at $a$ or $b$. If $E$ is in the essential
spectrum, it can happen that such solutions do not exist or it may happen that they do.
In Theorems \ref{t3.1.3}, \ref{t3.1.4} below, we shall explicitly assume
such solutions exist for the energies of interest. If these energies are
not in the essential spectrum, that is automatically fulfilled.
The key idea in \cite{GST96a} is to look at zeros of the Wronskian. That
zeros of the Wronskian are related to oscillation theory is indicated by
an old paper of Leighton \cite{Le52}, who noted that if $u_j,pu_j' \in
AC_{\text{\rm{loc}}}((a,b))$, $j=1,2$ and $u_1$ and $u_2$ have a
nonvanishing Wronskian $W(u_1,u_2)$ in $(a,b)$, then their zeros must
intertwine each other. (In fact, $pu_1'$ must have opposite signs at
consecutive zeros of $u_1$, so by nonvanishing of $W$, $u_2$ must have
opposite signs at consecutive zeros of $u_1$ as well. Interchanging the
role of $u_1$ and $u_2$ yields strict interlacing of their zeros.)
Moreover, let $E_1<E_2$ and $\tau u_j = E_j u_j$, $j=1,2$. If $x_0,x_1$
are two consecutive zeros of $u_1$, then the number of zeros of $u_2$
inside $(x_0,x_1)$ is equal to the number of zeros of the Wronskian
$W(u_1,u_2)$ plus one (cf.\ Theorem \ref{t3.7.4}). Hence the Wronskian
comes with a built-in renormalization counting the additional zeros of
$u_2$ in comparison to $u_1$.
We let $W_0(u_1,u_2)$ be the number of zeros of the Wronskian in the
open interval $(a,b)$ not counting multiplicities of zeros. Given
$E_1<E_2$, we let $N_0(E_1, E_2)=\text{\rm{dim}}(
\text{\rm{ran}}( P_{(E_1, E_2)}(H)))$ be the dimension of the spectral
projection $P_{(E_1, E_2)}(H)$ of $H$.
We begin by presenting two aspects of zeros of the Wronskian which are
critical for the two halves of our proofs (i.e., for showing $N_0\geq
W_0$ and that $N_0\leq W_0$). First, the vanishing of the Wronskian lets
us patch solutions together:
\begin{lemma} \label{l3.3.1} Suppose that $\psi_{+,j}, \psi_-\in
D_{\text{\rm{loc}}}$ and that $\psi_{+,j}$ and $\tau\psi_{+,j}$,
$j=1,2$ are in $L^{2}((c,b))$ and that $\psi_-$ and $\tau\psi_-$ are
in $L^{2}((a,c))$ for all $c\in (a,b)$. Suppose, in addition, that
$\psi_{+,j}$, $j=1,2$ satisfy the boundary condition defining $H$ at
$b$ $($i.e., $W(u,\psi_{+,j})(c)\to 0$ as $c\uparrow b$ for all $u\in
\text{\rm{dom}}(H)$$)$ and similarly, that $\psi_-$ satisfies the boundary condition
at $a$. Then, \\
$(i)$ If $W(\psi_{+,1}, \psi_{+,2})(c)=0$ and $(\psi_{+,2}(c),
(p\psi'_{+,2})(c)) \neq (0,0)$, then there exists a $\gamma$ such that
$$
\eta=\chi_{[c,b)}(\psi_{+,1}-\gamma\psi_{+,2})\in \text{\rm{dom}}(H)
$$
and
\begin{equation*}
H\eta=\chi_{[c,b)}(\tau\psi_{+,1}-\gamma\tau\psi_{+,2}).
\end{equation*}
$(ii)$ If $W(\psi_{+,1}, \psi_-)(c)=0$ and $(\psi_-(c),
(p\psi_-')(c))\neq (0,0)$, then there is a $\gamma$ such that
$$
\eta=\gamma\chi_{(a,c]}\psi_- + \chi_{(c,b)}\psi_{+,1}\in \text{\rm{dom}}(H)
$$
and
\begin{equation*}
H\eta =\gamma\chi_{(a,c]}\tau\psi_- + \chi_{(c,b)}\tau\psi_{+,1}.
\end{equation*}
\end{lemma}
The second aspect connects zeros of the Wronskian to Pr\"ufer
variables $\rho_u, \theta_u$ (for $u,pu'$ continuous) defined by
$$
u(x)=\rho_u(x)\sin(\theta_u(x)), \quad (pu')(x)=\rho_u(x)
\cos(\theta_u(x)).
$$
If $(u(x), (pu')(x))$ is never $(0,0)$, then $\rho_u$ can be chosen
positive and $\theta_u$ is uniquely determined once a value of $\theta_u
(x_0)$ is chosen subject to the requirement that $\theta_u$ be continuous in $x$.
Notice that
$$
W(u_1,u_2)(x)= \rho_{u_1}(x)\rho_{u_2}(x)\sin(\theta_{u_1}(x)
- \theta_{u_2}(x)).
$$
Thus, one obtains the following results.
\begin{lemma} \label{l3.3.2} Suppose $(u_j,pu_j')$, $j=1,2$ are never
$(0,0)$. Then $W(u_1,u_2)(x_0)$ is zero if and only if $\theta_{u_1}
(x_0)= \theta_{u_2}(x_0) \, (\text{mod} \; \pi)$.
\end{lemma}
In linking Pr\"ufer variables to rotation numbers, an important role
is played by the observation that because of
$$
u(x) = \int_{x_0}^x dt \, \frac{\rho_u(t) \cos(\theta_u(t))}{p(t)},
$$
$\theta_u(x_0)= 0 \, (\text{mod} \; \pi)$ implies $[\theta_u(x)-
\theta_u(x_0)]\big/ (x-x_0) >0$ for $0 <|x - x_0|$ sufficiently small
and hence for all $0 <|x - x_0|$ if $(u,pu') \neq (0,0)$. (In fact,
suppose
$x_1 \ne x_0$ is the closest $x$ such that $\theta_u(x_1)=\theta_u
(x_0)$ then apply the local result at $x_1$ to obtain a contradiction.)
We summarize:
\begin{lemma} \label{l3.3.3} $(i)$ If $(u,pu') \neq (0,0)$ then
$\theta_u(x_0)= 0 \, (\text{mod} \; \pi)$ implies
$$
[\theta_u(x)-\theta_u(x_0)]\big/ (x-x_0) >0
$$
for $x \ne x_0$. In particular, if $\theta_u(c)\in [0,\pi)$ and $u$ has
$n$ zeros in $(c,d)$, then $\theta_u(d-\epsilon)\in (n\pi, (n+1)\pi)$
for sufficiently small $\epsilon > 0$. \\
$(ii)$ Let $E_1<E_2$ and assume that $u_{1,2}$
solve $\tau u_j = E_j u_j$, $j=1,2$. Let $\Delta(x)=\theta_{u_2}(x)-
\theta_{u_1}(x)$. Then $\Delta(x_0)= 0 \, (\text{mod} \; \pi)$ implies
$(\Delta(x)-\Delta(x_0))/ (x-x_0)>0$ for $0<|x -x_0|$.
\end{lemma}
\begin{remark} \label{r3.3.4} $(i)$ Suppose $r,p$ are continuous on $(a,b)$.
If $\theta_{u_1}(x_0)= 0 \, (\text{mod} \; \pi)$ then $\theta_{u_1}(x) -
\theta_{u_1}(x_0) = c_0(x-x_0) + o(x-x_0)$ with $c_0>0$. If $\Delta
(x_0)= 0 \, (\text{mod} \; \pi)$ and $\theta_{u_1}(x_0)\neq 0 \, (\text{mod} \; \pi)$, then $\Delta(x)-\Delta (x_0)=c_1(x-x_0)+o(x-x_0)$ with
$c_{1}>0$. If $\theta_{u_1} (x_0)= 0=\Delta(x_0) \, (\text{mod} \; \pi)$,
then $\Delta(x) - \Delta(x_0)=c_2(x-x_0)^{3}+o(x-x_0)^{3})$ with $c_2>0$.
Either way, $\Delta$ increases through $x_0$. (In fact, $c_0=p(x_0)^{-1}$,
$c_{1}= (E_2-E_1)r(x_0)\sin^2 (\theta_{u_1}(x_0))$ and $c_{2}=\frac{1}{3}
r(x_0)p(x_0)^{-2}(E_2-E_1))$. \\
$(ii)$ In other words, Lemma \ref{l3.3.3} implies that the integer parts
of $\theta_u/ \pi$ and $\Delta_{u,v}/ \pi$ are increasing with respect
to $x\in(a,b)$ (even though $\theta_u$ and $\Delta_{u,v}$ themselves
are not necessarily monotone in $x$). \\
$(iii)$ Let $E \in [E_1,E_2]$ and assume $[E_1,E_2]$ to be outside the
essential spectrum of $H$. Then, for $x \in (a,b)$ fixed,
\begin{equation*}
\frac{d\theta_{\psi_\pm}}{dE}\,(E,x) =
-\frac{\int^{\ell_\pm}_x dt \, \psi_\pm(E,t)^2}
{\rho_{\psi_\pm}(E,x)}
\end{equation*}
proves that $\mp\theta_{\psi_\pm}(E,x)$ is strictly increasing with
respect to $E$.
\end{remark}
We continue with some preparatory results in the regular case.
\begin{lemma} \label{l3.4.2} Assume $H$ to be a regular SL operator. \\
$(i)$ Let $u_{1,2}$ be eigenfunctions of $H$ with eigenvalues $E_1<E_2$
and let $\ell$ be the number of eigenvalues of $H$ in $(E_1, E_2)$. Then
$W(u_1,u_2)(x)$ has exactly $\ell$ zeros in $(a,b)$. \\
$(ii)$ Let $E_1\leq E_2$ be eigenvalues of $H$ and suppose $[E_1, E_2]$ has
$\ell$ eigenvalues. Then for $\epsilon \geq 0$ sufficiently small,
$W_0(\psi_-(E_1 - \epsilon), \psi_+(E_2+\epsilon)) =\ell$. \\
$(iii)$ Let $E_3<E_4<E$ and $u$ be any solution of $\tau u=E u$. Then,
\begin{equation}
W_0(\psi_-(E_3), u)\geq W_0(\psi_-(E_4), u). \label{3.4.3}
\end{equation}
Similarly, if $E_3>E_4>E$ and $u$ is any solution of $\tau u=E u$,
then \eqref{3.4.3} holds. \\
$(iv)$ Item $(iii)$ remains true if every $\psi_-$ is
replaced by a $\psi_+$.
\end{lemma}
\begin{remark} \label{r3.4.3a} $(i)$ Since $(E_1, E_2)$ has $\ell-2$
eigenvalues, Lemma \ref{l3.4.2}\,$(i)$ implies that the Wronskian
$W(\psi_-(E_1), \psi_+(E_2))(x)$ has $\ell-2$ zeros in
$(a,b)$ and clearly it has zeros at $a$ and $b$. Essentially, Lemma
\ref{l3.4.2}\,$(ii)$ implies that replacing $E_1$ by $E_1-\epsilon$ and
$E_2$ by $E_2+\epsilon$ moves the zeros at $a,b$ inside $(a,b)$ to give
$\ell-2+2=\ell$ zeros. \\
$(ii)$ Lemma \ref{l3.4.2}\,$(iv)$ follows from Lemma \ref{l3.4.2}\,$(iii)$
upon reflecting at some point $c \in (a,b)$, implying an interchange of
$\psi_+$ and $\psi_-$.
\end{remark}
Lemma \ref{l3.4.2} then implies the following result.
\begin{lemma} \label{l3.4.1} Let $H$ be a regular SL operator and suppose
$E_1<E_2$. Then,
\begin{equation*}
W_0(\psi_-(E_1), \psi_+(E_2))\geq N_0(E_1, E_2).
\end{equation*}
\end{lemma}
Using the approach of Weidmann (\cite[Ch.\ 14]{We87}) to control some
limits, one can remove the assumption that $H$ is regular in Lemma
\ref{l3.4.1} as follows.
Fix functions $u_1,u_2 \in D_{\text{\rm{loc}}}$. Pick $c_{n}\downarrow a$,
$d_{n}\uparrow b$. Define $\tilde{H}_n$ on $L^{2}((c_{n}, d_{n}); r\,dx)$
by imposing the following boundary conditions on
$\eta\in \text{\rm{dom}}(\tilde{H}_{n})$
\begin{equation*}
W(u_1,\eta)(c_{n})=0=W(u_2,\eta)(d_{n}).
\end{equation*}
On $L^{2}((a,b);r\,dx)= L^{2}((a,c_{n});r\,dx)\oplus L^{2}
((c_{n}, d_{n});r\,dx)\oplus L^{2}((d_{n}, b);r\,dx)$ take $H_{n}
=\alpha I\oplus\tilde{H}_{n}\oplus\alpha I$ with $\alpha$ a
fixed real constant. Then Weidmann proves:
\begin{lemma} \label{l3.5.1} Suppose that either $H$ is limit point at
$a$ or that $u_1$ is a $\psi_-(E,x)$ for some $E$ and similarly, that
either $H$ is limit point at $b$ or $u_2$ is a $\psi_+(E',x)$ for
some $E'$. Then $H_n$ converges to $H$ in strong resolvent sense as
$n\to\infty$.
\end{lemma}
The idea of Weidmann's proof is that it suffices to find a core $D_0$ of
$H$ such that for every $\eta \in D_0$ there exists an $n_0 \in {\mathbb{N}}$
with $\eta\in D_0$ for $n \geq n_0$ and $H_{n}\eta\to H\eta$ as $n$
tends to infinity (see \cite[Theorem 9.16\,(i)]{We80}). If $H$ is limit
point at both ends, take $\eta\in D_0=\{u\in
D_{\text{\rm{loc}}}\,|\,
\text{supp}(u)\text{ compact in }(a,b)\}$. Otherwise,
pick
$\tilde{u}_1,\tilde{u}_2\in \text{\rm{dom}}(H)$ with $\tilde{u}_2 =u_2$
near $b$ and
$\tilde{u}_2=0$ near $a$ and with $\tilde{u}_1=u_1$ near
$a$ and
$\tilde{u}_1=0$ near $b$. Then pick $\eta\in D_0+\text{span}[\tilde{u}_1,
\tilde{u}_2]$ which one can show is a core for $H$ (\cite[Ch.\ 14]{We87}).
Secondly one uses the following fact:
\begin{lemma} \label{l3.5.2} Let $A_{n}\to A$ in strong resolvent sense as $n
\to \infty$. Then,
\begin{equation*}
\text{\rm{dim}}(\text{\rm{ran}}(P_{(E_1,E_2)}(A)))\leq
\varliminf_{n \to \infty} \text{\rm{dim}}(\text{\rm{ran}}(
P_{(E_1,E_2)}(A_{n}))).
\end{equation*}
\end{lemma}
Combining Lemmas \ref{l3.4.1}--\ref{l3.5.2} then yields the following
result.
\begin{theorem} \label{t3.1.5} Let $E_1 < E_2$. If $u_1=\psi_-(E_1)$ and
either $u_2=\psi_+(E_2)$ or $\tau u_2 = E_2 u_2$ and $H$ is limit point
at $b$. Then,
$$
W_0(u_1,u_2)\geq\text{\rm{dim}}(\text{\rm{ran}}((P_{(E_1, E_2)}(H))).
$$
\end{theorem}
Next, we indicate how the following result can be proved:
\begin{theorem} \label{t3.1.6} Let $E_1 < E_2$. Let either $u_1=\psi_+(E_1)$
or
$u_1=\psi_-(E_1)$ and either $u_2=\psi_+(E_2)$ or $u_2=\psi_-(E_2)$.
Then,
\begin{equation}
W_0(u_1,u_2)\leq\text{\rm{dim}}(\text{\rm{ran}}(P_{(E_1, E_2)}(H))). \label{3.1.8}
\end{equation}
\end{theorem}
Let $E_1<E_2$. Suppose
first that $u_1=\psi_-(E_1)$ and $u_2=\psi_+(E_2)$. Let $x_1,\dots,
x_{m}$ be zeros of $W(u_1,u_2)(x)$. Suppose we can prove that $\text{\rm{dim}}
P_{(E_1, E_2)}(H) \geq m$. If $W_0(u_1,u_2)=m$, this proves \eqref{3.1.8}.
If
$W_0=\infty$, we can take $m$ arbitrarily large, and again \eqref{3.1.8}
holds. Define
$$
\eta_j(x) = \begin{cases} u_1(x), & x\leq x_j, \\
\gamma_j u_2(x), & x\geq x_j, \end{cases} \quad 1 \le j \le m,
$$
where $\gamma_j$ is defined such that $\eta_j\in \text{\rm{dom}}(H)$ by Lemma
\ref{l3.3.1}. Let
$$
\tilde{\eta}_j(x) = \begin{cases} u_1(x), & x\leq x_j, \\
-\gamma_j u_2(x), & x>x_j, \end{cases} \quad 1 \le j \le m.
$$
If $E_2$ is an eigenvalue of $H$, we define in addition $\eta_0 = u_2
= -\tilde{\eta}_0$, $x_0=a$ and if $E_1$ is an eigenvalue of $H$,
$\eta_{m+1} = u_1 = \tilde{\eta}_{m+1}$, $x_{m+1}=b$.
\begin{lemma} \label{l3.6.1} $\langle\eta_j, \eta_{k}\rangle =\langle
\tilde{\eta}_j, \tilde{\eta}_{k}\rangle$ for all $j,k$, where
$\langle \,\cdot\, ,\,\cdot\,\rangle$ is the $L^{2}((a,b);r\,dx)$
inner product.
\end{lemma}
Notice that by (3.2),
\begin{equation*}
\biggl(H-\frac{E_2+E_1}{2}\biggr)\, \eta_j=
\biggl(\frac{E_1-E_2}{2}\biggr)\, \tilde{\eta}_j.
\end{equation*}
This result and Lemma \ref{l3.6.1} imply the following lemma.
\begin{lemma} \label{l3.6.2} If $\eta$ is in the span of the $\eta_j$, then
$$
\left\|\biggl(H-\frac{E_2+E_1}{2}\biggr)\, \eta\right\| =
\frac{|E_2-E_1|}{2}\, \|\eta\|.
$$
\end{lemma}
Thus, $\text{\rm{dim}}(\text{\rm{ran}}( P_{[E_1, E_2]}(H)))\geq
\text{\rm{dim}}(\text{span}(\{\eta_j\}))$. But $u_1$ and $u_2$ are independent
on each interval (since their Wronskian is nonconstant) and so the
$\eta_j$ are linearly independent. This proves Theorem \ref{t3.1.6} in the
$\psi_-(E_1), \psi_+(E_2)$ case. The case $u_1=\psi_-(E_1)$,
$u_2=\psi_-(E_2)$ is proved similarly. The cases $u_1=\psi_+(E_1)$,
$u_2=\psi_\pm(E_2)$ can be obtained by reflection.
Next one proves the following result.
\begin{theorem} \label{t3.7.1} Let $E_1\neq E_2$. Let $\tau u_j=E_j u_j$,
$j=1,2$, $\tau v_2=E_2 v_2$ with $u_2$ linearly independent of $v_2$.
Then the zeros of $W(u_1,u_2)$ interlace the zeros of $W(u_1,v_2)$ and
vice versa $($in the sense that there is exactly one zero of one
function in between two zeros of the other$)$. In particular,
$|W_0(u_1,u_2)-W_0(u_1,v_2)| \leq 1$.
\end{theorem}
Theorems \ref{t3.1.5}, \ref{t3.1.6}, and \ref{t3.7.1} then yield
the following two theorems, the principal results of \cite{GST96a}:
\begin{theorem} \label{t3.1.3} Suppose $E_1 < E_2$. Let $u_1=\psi_-(E_1)$ and
$u_2=\psi_+(E_2)$. Then,
$$
W_0(u_1,u_2)=N_0(E_1, E_2).
$$
\end{theorem}
\begin{theorem} \label{t3.1.4} Suppose $E_1 < E_2$. Let $u_1=\psi_-(E_1)$ and
$u_2=\psi_-(E_2)$. Then either,
\begin{equation}
W_0(u_1,u_2)=N_0(E_1, E_2), \label{3.1.6}
\end{equation}
or,
\begin{equation}
W_0(u_1,u_2)=N_0(E_1, E_2) -1. \label{3.1.7}
\end{equation}
If either $N_0=0$ or $H$ is limit point at $b$, then \eqref{3.1.6} holds.
\end{theorem}
One infers that if $b$ is a regular point and $E_2 > e > E_1$ with $e$
an eigenvalue and $|E_2-E_1|$ is small, then \eqref{3.1.7} holds rather
than \eqref{3.1.6}. One also sees that if $u_{1,2}$ are arbitrary
solutions of $\tau u_j=E_j u_j$, $j=1,2$, then, in general, $|W_0-N_0|\leq
2$ (this means that if one of the quantities is infinite, the other is as
well) and we note that any of $0, \pm 1,\pm 2$ can occur for $W_0-N_0$.
Especially, if either $E_1$ or $E_2$ is in the interior of the essential
spectrum of $H$ (or $\text{\rm{dim}}(\text{\rm{ran}}(P_{(E_1, E_2)}(H)))=\infty$), then $W_0
(u_1,u_2)=\infty$ for any $u_1$ and $u_2$ satisfying $\tau u_j=E_j
u_j$, $j=1,2$ (cf.\ Theorem \ref{t3.7.3}).
\begin{remark} \label{r3.1.4a} Of course, by reflecting about a point
$c \in (a,b)$, Theorems \ref{t3.1.5}, \ref{t3.1.3}, and \ref{t3.1.4} hold
for $u_1= \psi_+(E_1)$ and $u_2 =
\psi_-(E_2)$ (and either $N_0 = 0$ or $H$ is limit point at $a$ in the
corresponding analog of Theorem \ref{t3.1.4} yields \eqref{3.1.6}) and
similarly,
$\tau u_2 = E_2 u_2$ and $H$ is limit point at $a$ yields the conclusion
in the corresponding analog of Theorem \ref{t3.1.5}.
\end{remark}
We add a few more results proved in \cite{GST96a}.
By applying Theorem \ref{t3.7.1} twice, one concludes
\begin{theorem} \label{t3.7.2} Let $E_1\neq E_2$. Let $u_1, u_2, v_1, v_2$ be
the linearly independent functions with $\tau u_j=E_j u_j$ and $\tau v_j
=E_j v_j$. Then,
$$
|W_0(u_1, u_2)-W_0(v_1, v_2)|\leq 2.
$$
\end{theorem}
\begin{theorem} \label{t3.7.3} If $\text{\rm{dim}}(\text{\rm{ran}}(P_{(E_1,E_2)}(H)))
=\infty$, then $W_0(u_1,u_2)=\infty$ for any $u_1$ and $u_2$ satisfying
$\tau u_j = E_j u_j$, $j=1,2$.
\end{theorem}
\begin{theorem} \label{t3.7.4} Let $E_1<E_2$. Let $\tau u_j = E_j u_j$,
$j=1,2$. If $a < x_0 < x_1 < b$ are zeros of $u_1$ or of $W(u_1, u_2)
(\,.\,)$, then the number of zeros of $u_2$ inside $(x_0,x_1)$ equals
the number of zeros of $W(u_1,u_2)(\,.\,)$ inside $(x_0,x_1)$ plus the
number of zeros of $u_1$ inside $(x_0,x_1)$ plus one.
\end{theorem}
The following result is of special interest in connection with
the problem of whether the total number of eigenvalues of $H$ in
one of its essential spectral gaps is finite or infinite. In
particular, the energies $E_1, E_2$ in Theorem 7.5 below may lie
in the essential spectrum of $H$. For this purpose we consider an
auxiliary Dirichlet operator $H_{x_0}^D$, $x_0\in(a,b)$ associated
with $H$. $H_{x_0}^D$ is obtained by taking the direct sum of the
restrictions $H_{x_0,\pm}^D$ of $H$ to $(a,x_0)$, respectively
$(x_0,b)$, with a Dirichlet boundary condition at $x_0$. We
emphasize that the Dirichlet boundary conditions can be replaced
by boundary conditions of the type $\lim_{\epsilon\downarrow 0}
[u'(x_0\pm\epsilon)+ \beta u(x_0\pm\epsilon)] =0$, $\beta\in{\mathbb{R}}$.
\begin{theorem} \label{t3.7.5}
Let $E_1<E_2$. Let $\tau u_j = E_j u_j$, $\tau s_j = E_j s_j$,
and $s_j(E_j,x_0)=0$, $j=1,2$. Then,
\begin{align*}
& \text{\rm{dim}}(\text{\rm{ran}}(P_{(E_1,E_2)}(H)))<\infty
\, \text { if and only if } \, W_0(u_1,u_2)<\infty, \\
& \text{\rm{dim}}(\text{\rm{ran}}(P_{(E_1,E_2)}(H)))-1
\le \text{\rm{dim}}(\text{\rm{ran}}(P_{(E_1,E_2)}(H^D_{x_0}))) \\
& \hspace*{4.13cm} \le
\text{\rm{dim}}(\text{\rm{ran}}(P_{(E_1,E_2)}(H)))+2, \\
& W_0(s_1,s_2)-1 \le \text{\rm{dim}}(\text{\rm{ran}}(
P_{(E_1,E_2)}(H^D_{x_0}))) \le W_0(s_1,s_2)+1.
\end{align*}
\end{theorem}
For an application of this circle of ideas to the notion of the density of
states we refer to \cite{GST96a}.
\medskip
{\it More recent references:} Oscillation and renormalized oscillation theory was also put in perspective by Simon's contribution \cite{Si05} to the the Festschrift \cite{AHP05} in honor of Sturm and Liouville.
Renormalized oscillation theory for one-dimensional Jacobi and Dirac-type operators was
developed by Teschl \cite{Te96} (see also \cite[Sect.\ 4.3]{Te00}) and \cite{Te98b}. For an
interesting application of some of the results in \cite{GST96a} to the stability theory of complete minimal surfaces we refer to a paper by Schmidt \cite{Sc01}. For additional results on oscillation theory, critical coupling constants, and eigenvalue asymptotics we refer to Schmidt \cite{Sc00}.
\section{Trace Formulas for Schr\"odinger and Jacobi Operators: The xi
Function} \label{s4}
In this section we summarize some of the principal results of
the following papers: \\
\cite{GHS95} F.~Gesztesy, H.~Holden, and B.~Simon, {\it Absolute
summability of the trace relation for certain Schr\"odinger operators},
Commun. Math. Phys. {\bf 168}, 137--161 (1995). \\
\cite{GHSZ93} F.~Gesztesy, H.~Holden, B.~Simon, and Z.~Zhao, {\it Trace
formulae and inverse spectral theory for Schr\"odinger operators},
Bull. Amer. Math. Soc. {\bf 29}, 250--255 (1993). \\
\cite{GHSZ95} F.~Gesztesy, H.~Holden, B.~Simon, and Z.~Zhao, {\it Higher
order trace relations for Schr\"odinger operators}, Rev. Math.
Phys. {\bf 7}, 893--922 (1995).\\
\cite{GHSZ96} F.~Gesztesy, H.~Holden, B.~Simon, and Z.~Zhao, {\it A trace
formula for multidimensional Schr\"odinger operators}, J. Funct. Anal.
{\bf 141}, 449--465 (1996). \\
\cite{GS96} F.~Gesztesy and B.~Simon, {\it The xi function}, Acta Math.
{\bf 176}, 49--71 (1996).
\medskip\smallskip
We start with \cite{GS96}. One of the principal goals in \cite{GS96} was
to introduce a special function
$\xi(\cdot,\cdot)$ on ${\mathbb{R}}\times{\mathbb{R}}$ associated with one-dimensional
Schr\"odinger operators $H$ (and Jacobi operators $h$) which led to a
generalization of the known trace formula for periodic Schr\"odinger
operators for general potentials $V$ and established
$\xi$ as a new tool in the spectral theory of one-dimensional
Schr\"odinger operators and (multi-dimensional) Jacobi operators.
To illustrate this point we recall the well-known trace formula for
periodic potentials $V$ of period $a>0$: Then, by Floquet theory (see,
e.g., \cite{Ea73}, \cite{MW79}, \cite{RS78})
$$
\sigma (H)=[E_{0}, E_{1}]\cup [E_{2}, E_{3}]\cup\dots
$$
a set of bands. If $V$ is $C^1({\mathbb{R}})$, one can show that the sum of the
gap sizes is finite, that is,
\begin{equation}
\sum^{\infty}_{n=1}|E_{2n}-E_{2n-1}|<\infty. \label{4.1.1}
\end{equation}
For fixed $y$, let $H_y$ be the operator $-\frac{d^2}{dx^2}+V$
in $L^{2}([y, y+a])$ with Dirichlet boundary conditions $u(y)=u(y+a)=0$ .
Its spectrum is discrete, that is, there are eigenvalues
$\{\mu_{n}(y)\}^{\infty}_{n=1}$ with
\begin{equation}
E_{2n-1}\leq\mu_{n}(y)\leq E_{2n}. \label{4.1.2}
\end{equation}
The trace formula for $V$ then reads
\begin{equation}
V(y)=E_{0}+\sum^{\infty}_{n=1} [E_{2n}+E_{2n-1}-2\mu_{n}(y)]. \label{4.1.3}
\end{equation}
By \eqref{4.1.2},
$
|E_{2n}+E_{2n-1}-2\mu_{n}(y)|\leq |E_{2n}-E_{2n-1}|
$
so \eqref{4.1.1} implies the convergence of the sum in \eqref{4.1.3}. An
elegant direct proof of \eqref{4.1.3} can be found, for instance, in
\cite[Sect.\ 26]{Si79}.
The earliest trace formula for Schr\"odinger operators was found on a
finite interval in 1953 by Gel'fand and Levitan \cite{GL87} with later
contributions by Dikii \cite{Di61}, Gel'fand \cite{Ge87},
Halberg and Kramer \cite{HK60}, and Gilbert and Kramer
\cite{GK64}. The first trace formula for periodic $V$ was obtained in
1965 by Hochstadt \cite{Ho65}, who showed that for finite-band
potentials
$$
V(x)-V(0)=2\sum^{g}_{n=1}[\mu_{n}(0)-\mu_{n}(x)].
$$
Dubrovin [9] then proved \eqref{4.1.3} for finite-band potentials. The
general formula \eqref{4.1.3} under the hypothesis that $V$ is periodic
and in $C^{\infty}({\mathbb{R}})$ was proved in 1975 by Flaschka \cite{Fl75}
and McKean and van Moerbeke \cite{MM75}, and later for general $C^3({\mathbb{R}})$
potentials by Trubowitz \cite{Tr77}. Formula \eqref{4.1.3} is a key
element of the solution of inverse spectral problems for periodic
potentials \cite{Du75}, \cite{Fl75}, \cite{Ho65}, \cite[Ch.\ 11]{Le87},
\cite[Sect.\ 4.3]{Ma86}, \cite{MM75}, \cite{MT76},, \cite{Tr77}.
There have been two classes of potentials for which \eqref{4.1.3} has been
extended. Certain almost periodic potentials are studied in Craig
\cite{Cr89}, Levitan \cite{Le85}, \cite[Ch.\ 11]{Le87}, and Kotani-Krishna
\cite{KK88}.
In 1979, Deift and Trubowitz \cite{DT79} proved that if $V(x)$ decays
sufficiently rapidly at infinity and $-\frac{d^2}{dx^2}+V$ has no negative
eigenvalues, then
\begin{equation}
V(x)=\frac{2i}{\pi}\int^{\infty}_{-\infty}\,dk\,k\,
\ln\biggl[1+R(k)\frac{f_{+}(x, k)}{f_{-}(x, k)}\biggr] \label{4.1.4}
\end{equation}
(where $f_{\pm}(x, k)$ are the Jost functions at energy $E=k^{2}$ and
$R(k)$ is a reflection coefficient) which can be shown to be an analog
of \eqref{4.1.3}. Previously, Venakides \cite{Ve88} studied a trace
formula for $V$, a positive smooth potential of compact support, by
writing \eqref{4.1.3} for the periodic potential
$
V_{L}(x)=\sum^{\infty}_{n=-\infty}V(x+nL)
$
and then taking $L$ to $\infty$. He found an integral formula which
is precisely \eqref{4.1.4} (although, this was not identified as such in
\cite{Ve88}).
The basic definition of $\xi$ depends on the theory of the Lifshits--Krein
spectral shift function \cite{Kr62}. If $A$ and $B$ are self-adjoint
operators bounded from below, that is, $A\geq\eta$, $B\geq\eta$ for some
real
$\eta$, and so that
$[(A+i)^{-1}- (B+i)^{-1}]$ is trace class, then there exists a measurable
function
$\xi(\lambda)$ associated with the pair $(B, A)$ so that
\begin{equation}
\Tr[f(A)-f(B)]=-\int_{{\mathbb{R}}} d\lambda \, f'(\lambda)\xi(\lambda)
\label{4.1.5}
\end{equation}
for a class of functions $f$ which are sufficiently smooth and which
decay sufficiently rapidly at infinity, and, in particular, for
$f(\lambda)=e^{-t\lambda}$ for any $t>0$; and so that
\begin{equation}
\xi(\lambda)=0 \, \text{ for } \, \lambda<\eta. \label{4.1.6}
\end{equation}
Moreover, \eqref{4.1.5}, \eqref{4.1.6} uniquely determine $\xi(\lambda)$
for a.e.~$\lambda$. Moreover, if $[(A+i)^{-1}-(B+i)^{-1}]$ is rank $n$,
then
$
|\xi(\lambda)|\leq n
$
and if $B\geq A$, then $\xi(\lambda)\geq 0$.
For the rank-one case of importance in this paper, an extensive study
of $\xi$ can be found in \cite{Si95}.
Let $V$ be a continuous function on ${\mathbb{R}}$ which is bounded from below.
Let $H=-\frac{d^2}{dx^2}+V$ in $L^2({\mathbb{R}})$ which is essentially self-adjoint
on $C^{\infty}_{0} ({\mathbb{R}})$ and let $H_{x}^D$ be the operator on
$L^{2}((-\infty, x))\oplus L^{2}((x, \infty))$ with $u(x)=0$ Dirichlet
boundary conditions. Then $[(H_{x}^D+i)^{-1}-(H+i)^{-1}]$ is rank one,
so there results a spectral shift function $\xi(\lambda,x)$ for the pair
$(H_{x}^D, H)$ which, in particular, satisfies,
\begin{equation}
\Tr[e^{-tH}-e^{-tH_{x}^D}]=t\int^{\infty}_{0} d\lambda \,
e^{-t\lambda}\xi(\lambda,x). \label{4.1.7}
\end{equation}
While $\xi$ is defined in terms of $H$ and $H_{x}^D$, there is a
formula that only involves $H$, or more precisely, the Green's
function $G(z,x,y)$ of $H$ defined by
$$
((H-z)^{-1}f)(x)=\int_{{\mathbb{R}}} dy\,G(z,x, y)f(y), \quad \Im(z)\neq 0.
$$
Then by general principles, $\lim_
{\epsilon\downarrow 0}\, G(\lambda+i\epsilon,x, y)$ exists for
a.e.~$\lambda\in{\mathbb{R}}$, and
\begin{equation}
\xi(\lambda,x)=\frac{1}{\pi}\,\text{\rm{Arg}} \bigl(\lim_
{\epsilon\downarrow 0}\, G(\lambda+i\epsilon, x, x)\bigr). \label{4.1.8a}
\end{equation}
This is {\it formally} equivalent to formulas that Krein \cite{Kr62}
has for
$\xi$ but in a singular setting (i.e., corresponding to an infinite
coupling constant). With
this definition out of the way, we can state the general trace formula
derived in \cite{GS96}:
\begin{theorem} \label{t4.3.1} Suppose $V$ is a continuous function bounded
from below on ${\mathbb{R}}$. Let $\xi(\lambda,x)$ be the spectral
shift function for the pair $(H_{x}^D, H)$ with $H_{x}^D$ the operator on
$L^{2}((-\infty, x))\oplus L^{2}((x, \infty))$ obtained from
$H=-\frac{d^2} {dx^2}+V$ in $L^2({\mathbb{R}})$ with a Dirichlet boundary
condition at $x$. Let $E_{0}\leq \inf\,\sigma(H)$. Then
\begin{equation}
V(x)=\lim_{\alpha\downarrow 0} \biggl[E_{0}+\int^{\infty}
_{E_0} d\lambda \, e^{-\alpha\lambda} [1-2\xi(\lambda,x)]\,
\biggr].
\label{4.3.1}
\end{equation}
\end{theorem}
In particular, if
$\int^{\infty}_{E_0} d\lambda \, |1-2\xi(\lambda,x)| <\infty$,
then
$$
V(x)=E_{0}+\int^{\infty}_{E_0} d\lambda \, [1-2\xi(\lambda,x)].
$$
We note that the trace formula extends to real-valued potentials $V\in
L^1_{\text{\rm{loc}}}({\mathbb{R}})$ as long as $H$ stays bounded from below (it then is in
the limit point case at $\pm\infty$). Equation \eqref{4.3.1} then holds at
all Lebesgue points of $V$ and hence for a.e.\ $x\in{\mathbb{R}}$.
For certain almost periodic potentials, Craig \cite{Cr89} used a
regularization similar to the $\alpha$-regularization in \eqref{4.3.1}.
Basically, \eqref{4.3.1} follows from \eqref{4.1.7} and an asymptotic
formula,
\begin{equation}
\Tr[e^{-tH}-e^{-tH_{x}^D}]=\frac{1}{2}[1-tV(x)+o(t)]. \label{4.3.2}
\end{equation}
We present a few examples next:
\begin{example} \label{e4.3.2} Pick a constant $C\in{\mathbb{R}}$ such that $V(x)=C$
for all $x\in{\mathbb{R}}$. Then
$G(\lambda,x,x)=(C-\lambda)^{-1/2}/2$ and hence one infers that
$\text{\rm{Arg}}(G(\lambda,x,x))=0$ (resp., $\pi/2$) if
$\lambda<C$ (resp., $\lambda>C$). Thus, by \eqref{4.1.8a},
$\xi(\lambda,x)=1/2$ on $(C, \infty)$ and $\xi(\lambda,x)=0$ on $(-\infty,C)$.
When $\xi=1/2$ on a subset
of $\sigma (H)$, that set does not contribute to the integral in
\eqref{4.3.1} and one verifies for $E_0 \leq C$,
$$
V(x)=E_0+\int_{E_0}^C d\lambda = E_0 + (C-E_0)=C, \quad x\in{\mathbb{R}}.
$$
\end{example}
\begin{example} \label{e4.3.3} Suppose that $V(x)\to\infty$ as $|x|\to\infty$.
Then $H$ has eigenvalues $E_{0}<E_{1}<E_{2}<\cdots$ and $H^{D}_{x}$ has
eigenvalues $\{\mu_{j}(x)\}^{\infty}_{j=1}$ with
$E_{j-1}\leq\mu_{j}(x)\leq E_{j}$. We have
$$
\xi(\lambda,x) = \begin{cases} 1, & E_{j-1}<\lambda <\mu_{j}(x), \\
0, & \lambda<E_{0} \text{ or } \mu_{j}(x)<\lambda<E_{j}.
\end{cases}
$$
Thus \eqref{4.3.1} becomes:
\begin{equation}
V(x)=E_{0}+\lim_{\alpha\downarrow 0}\, \biggl[\sum^{\infty}_{j=1}
\bigl(\left.2e^{-\alpha\mu_{j}(x)}-e^{-\alpha E_{j}}-e^{-\alpha E_{j-
1}}\bigr)\right/\alpha\biggr]. \label{4.3.3}
\end{equation}
If we could take $\alpha$ to zero inside the sum, we would get
\begin{equation}
V(x)=E_{0}+\sum^{\infty}_{j=1}\,[E_{j}+E_{j-1}-2\mu_{j}(x)]
\;\; \text{ (formal!)} \label{4.3.4}
\end{equation}
which is just a limit of the periodic formula \eqref{4.1.3} in the limit
of vanishing band widths. \eqref{4.3.3} is just a kind of Abelianized
summation procedure applied to \eqref{4.3.4}.
As a special case of this example, consider $V(x)=x^{2}-1$. Then
$E_{j}=2j$ and $\{\mu_{j}(0)\}$ is the set $\{2, 2, 6, 6, 10, 10, 14,
14, \dots \}$ of $j$ odd eigenvalues, each doubled. Thus \eqref{4.3.4} is the
formal sum
$$
-1=-2+2-2+2\dots \, \text{ (formal!)}
$$
with \eqref{4.3.3}
\begin{align*}
-1 &=\lim_{\alpha\downarrow 0}\,[(e^{-2\alpha}-1)/\alpha]\,[1-e^{-2\alpha}
+e^{-4\alpha}\dots ] \\
&=\lim_{\alpha\downarrow 0}\,[(e^{-2\alpha}-1)/\alpha]\,
[1/(1+e^{-2\alpha})]
\end{align*}
with a true abelian summation.
\end{example}
\begin{example} \label{e4.3.4} Suppose $V$ is periodic. Let
$E_{j}, \mu_{j}(x)$ be the band edges and Dirichlet eigenvalues as in
\eqref{4.1.2}, \eqref{4.1.3}. Then it follows from the fact that the two
Floquet solutions are complex conjugates of each other on the spectrum
of $H$, and the Wronskian is antisymmetric in its argument ($W(f,g)=-W(g,f)$),
that $g(\lambda,x)$ is purely imaginary on
$\sigma (H)$; that is,
$\xi(\lambda,x)=1/2$ there, so
$$
\xi(\lambda,x)= \begin{cases} 1/2, & E_{2n}<\lambda<E_{2n+1},
\\ 1, & E_{2n-1}<\lambda<\mu_{n}(x), \\
0, & \mu_{n}(x)<\lambda<E_{2n}.
\end{cases}
$$
(This fact has also been used by Deift and Simon
\cite{DS83} and Kotani
\cite{Ko84} and also follows from the fact that $g(\lambda,x)=G(\lambda+i0,x,x)
=-[m_{+}(\lambda,x)+m_{-} (\lambda,x)]^{-1}$ in terms of the Weyl--Titchmarsh
$m$-functions.) Thus,
$$
\int^{\infty}_{E_0} d\lambda \, |1-2\xi(\lambda,x)|=\sum^{\infty}_{n=1}
|E_{2n}-E_{2n-1}|
$$
is finite if \eqref{4.1.1} holds. In that case one can take the limit
inside the integral in \eqref{4.3.1} and so recover \eqref{4.1.3}.
\end{example}
\begin{example} \label{e4.3.5} In \cite{GHS95} it is proved that if $V$ is
short-range, that is, $V\in H^{2, 1}({\mathbb{R}})$, then, $\int^{\infty}_{E_0}
d\lambda \, |1-2\xi(\lambda,x)| <\infty$ and one can take the
limit in \eqref{4.3.1} inside the integral. This recovers Venakides'
result \cite{Ve88} with an explicit form for $\xi$ in terms of the Green's
function (see Theorem \ref{t4.3.1}). Similarly, one can treat short-range
perturbations $W$ of periodic background potentials $V$ (modeling
scattering off defects or impurities, described by $W$, in
one-dimensional solids) and ``cascading'' potentials, that is, potentials
approaching different spatial asymptotes sufficiently fast (cf.\
\cite{GHS95} for details).
\end{example}
Next we mention a striking inverse spectral application of the trace
formula \eqref{4.3.1} to a celebrated two-spectra inverse spectral
theorem due to Borg \cite{Bo46}:
\begin{theorem} \label{t4.3.2}
Let $V\in L^1_{\text{\rm{loc}}} ({\mathbb{R}})$ be real-valued and periodic. Let
$H=-\frac{d^2}{dx^2}+V$ be the associated self-adjoint
Schr\"odinger operator in $L^2({\mathbb{R}})$ and suppose that
$\sigma(H)=[E_0,\infty)$ for some $E_0\in{\mathbb{R}}$. Then
$$
V(x)=E_0 \, \text{ for a.e.\ $x\in{\mathbb{R}}$}.
$$
\end{theorem}
Given the trace formula \eqref{4.3.1} (observing the a.e.\ extension
noted after Theorem \ref{t4.3.1}) and using the fact that for all
$x\in{\mathbb{R}}$ and a.e.\ $\lambda > E_0$, $\xi(\lambda,x)=1/2$ (cf.\ Example
\ref{4.3.4}), the proof of Borg's Theorem \ref{t4.3.2} is effectively
reduced to just a one-line argument (as was observed in \cite{CGHL00}).
In addition, the new proof permits one to replace periodic by
reflectionless potentials and hence applies to algebro-geometric
quasiperiodic (KdV) potentials and certain classes of almost periodic
potentials.
Now we turn to an analog for Theorem \ref{t4.3.1} for
Jacobi operators. This turns out to be a special case of the following
result.
\begin{theorem} \label{t4.4.1} Let $A$ be a bounded self-adjoint operator
in some complex separable Hilbert space ${\mathcal H}$ with $\alpha=\inf\,
\sigma(A)$, $\beta=\sup\,\sigma(A)$. Let $\varphi\in{\mathcal H}$,
$\|\varphi\|_{{\mathcal H}}=1$ be an arbitrary unit vector in ${\mathcal H}$ and let
$\xi(\lambda)$ be the spectral shift function for the pair $(A_{\infty},A)$,
where
$A_\infty$ is defined by
$$
(A_\infty-z)^{-1}=(A-z)^{-1}-(\varphi,(A-z)^{-1}\varphi)^{-1}
((A-\overline z)^{-1}\varphi,\cdot)(A-z)^{-1}\varphi.
$$
Then for any $E_{-}\leq\alpha$ and $E_{+}\geq\beta$:
\begin{align*}
(\varphi, A\varphi) &=E_{-}+\int^{E_{+}}_{E_{-}} d\lambda \,
[1-\xi(\lambda)]
= E_{+}-\int^{E_{+}}_{E_{-}} d\lambda \, \xi(\lambda) \\
&=\frac12\, (E_{+}+E_{-})+\frac12 \int^{E_{+}}_{E_{-}} d\lambda \,
[1-2\xi(\lambda)].
\end{align*}
\end{theorem}
\begin{corollary} \label{c4.4.2} Let $H$ be a Jacobi matrix on
$\ell^{2}({\mathbb{Z}}^{\nu})$, that is, for a bounded function $V$ on
${\mathbb{Z}}^{\nu}$,
\begin{equation}
(Hu)(n)=\sum_{|n-m|=1} u(m)+V(n)u(n), \quad n\in{\mathbb{Z}}^{\nu}. \label{4.4.4}
\end{equation}
For $r\in{\mathbb{Z}}^{\nu}$, let $H^{D}_{r}$ be the operator on
$L^{2}({\mathbb{Z}}^{\nu}\backslash\{r\})$ given by \eqref{4.4.4} with $u(r)=0$
boundary conditions. Let $\xi(\lambda,r)$ be the spectral shift function for
the pair
$(H^{D}_{r}, H)$. Then
\begin{align}
\begin{split}
V(r) &= E_{-}+\int^{E_{+}}_{E_{-}} d\lambda \, [1-\xi(\lambda,r)]
= E_{+}-\int^{E_{+}}_{E_{-}} d\lambda \, \xi(\lambda,r) \\
&= \frac12\, (E_{+}+E_{-})+\frac12
\int^{E_{+}}_{E_{-}} d\lambda \, [1-2\xi(\lambda,r)] \label{4.4.5}
\end{split}
\end{align}
for any $E_{-}\leq\inf\,\sigma(H)$,
$E_{+}\geq\sup\,\sigma(H)$.
\end{corollary}
Only when $\nu=1$ does this have an interpretation as a formula using
Dirichlet problems on the half-line.
Next, we look at some applications to absolutely continuous spectra. In
particular, we will point out that $\xi(\lambda,x)$ for a single fixed
$x\in{\mathbb{R}}$ determines the absolutely continuous spectrum of a
one-dimensional Schr\"odinger or Jacobi operator. We begin with a
result that holds for higher-dimensional Jacobi operators as well:
\begin{lemma} \label{l4.5.1} $(i)$ For an arbitrary Jacobi matrix,
$H$, on ${\mathbb{Z}}^{\nu}$, $\operatornamewithlimits{\cup}_
{j\in{\mathbb{Z}}^{\nu}}\{\lambda\in{\mathbb{R}} \,|\, 0<\xi(\lambda,j)<1\}$ is an
essential support for the absolutely continuous spectrum of $H$. \\
$(ii)$ For a one-dimensional Schr\"odinger operator,
$H=-\frac{d^2}{dx^2}+V$ bounded from below,
$\operatornamewithlimits{\cup}_{x\in{\mathbb{Q}}}\{\lambda
\in{\mathbb{R}}\,|\, 0<\xi(\lambda,x)<1\}$ is an essential support for the
absolutely continuous spectrum of $H$.
\end{lemma}
\begin{remark} \label{r4.5.2} We recall that every absolutely continuous
measure, $d\mu$, has the form $f(E)\,dE$. $S=\{E\in{\mathbb{R}}\,|\, f(E)\neq
0\}$ is called an essential support. Any Borel set which differs from
$S$ by sets of zero Lebesgue measure is also called an essential support.
If $A$ is a self-adjoint operator on ${\mathcal H}$ and $\varphi_{n}$, an
orthonormal basis for ${\mathcal H}$, and $d\mu_{n}$, the spectral measure for
the pair, $A, \varphi_{n}$ (i.e., $(\varphi_{n}, e^{isA}\varphi_{n})=
\int_{{\mathbb{R}}} e^{isE}\,d\mu_{n}(E)$) and if $d\mu^{\rm ac}_{n}$
is the absolutely continuous component of $d\mu_{n}$ with $S_n$ its
essential support, then $\operatornamewithlimits{\cup}_{n}S_n$ is the
essential support of the absolutely continuous spectrum for $A$.
\end{remark}
In one dimension though, a single $x$ suffices:
\begin{theorem} \label{t4.5.2} For one-dimensional Schr\"odinger
$($resp., Jacobi$)$ operators, $\{\lambda\in{\mathbb{R}}\,|\,
0<\xi(\lambda,x)<1\}$ is an essential support for the absolutely
continuous measure for \rm{any } fixed $x\in{\mathbb{R}}$ $($resp., ${\mathbb{Z}}$$)$.
\end{theorem}
These results are of particular interest because of their implications
for a special kind of semi-continuity of the spectrum.
\begin{definition} \label{d4.5.2} Let $\{V_{n}\},V$ be continuous potentials
on ${\mathbb{R}}$ (resp., on ${\mathbb{Z}}$). We say that $V_n$ converges to $V$ locally
as $n\to\infty$ if and only if \\
$(i)$ $\inf_{(n, x)\in{\mathbb{N}}\times{\mathbb{R}}}V_{n}(x)>-\infty$
(${\mathbb{R}}$ case) or $\sup_{(n, j)\in{\mathbb{N}}\times{\mathbb{Z}}}
|V_{n}(j)|<\infty$ (${\mathbb{Z}}$ case). \\
$(ii)$ For each $R<\infty$, $\sup_{|x|\leq R} |V_{n}(x)
-V(x)|\to 0$ as $n\to\infty$.
\end{definition}
\begin{lemma} \label{l4.5.3} If $V_{n}\to V$ locally as $n\to\infty$ and
$H_n, H$ are the corresponding Schr\"odinger operators $($resp., Jacobi
matrices$)$, then $(H_{n}-z)^{-1}\to (H-z)^{-1}$ strongly for
$\text{Im}\,z\neq 0$ as $n\to\infty$.
\end{lemma}
\begin{theorem} \label{t4.5.4} If $V_{n}\to V$ locally as $n\to\infty$ and
$\xi_{n}(\lambda,x)$, $\xi(\lambda,x)$ are the corresponding xi
functions for fixed $x$, then $\xi_{n}(\lambda,x)\, d\lambda$ converges
to
$\xi(\lambda,x)\, d\lambda$ weakly in the sense that for any $f\in
L^{1}({\mathbb{R}}; d\lambda)$,
$$
\int_{{\mathbb{R}}} d\lambda \, f(\lambda) \, \xi_{n}(\lambda,x)\, \to
\int_{{\mathbb{R}}} d\lambda \, f(\lambda) \, \xi(\lambda,x)\, \,
\text{ as } \, n\to\infty.
$$
\end{theorem}
\begin{definition} \label{d4.5.4} For any $H$, let $|S_{\rm ac}(H)|$
denote the Lebesgue measure of the essential support of the absolutely
continuous spectrum of $H$.
\end{definition}
\begin{theorem} \label{t4.5.5} $($For one-dimensional Schr\"odinger
or Jacobi operators$)$ Suppose $V_{n}\to V$ locally as
$n\to\infty$ and each $V_n$ is periodic. Then for any interval
$(a, b)\subset{\mathbb{R}}$,
$$
|(a, b)\cap S_{\rm ac}|\geq \varlimsup_{n\to\infty}
|(a, b)\cap S_{\rm ac} (H_{n})|.
$$
\end{theorem}
We note that the periods of $V_{n}$ need {\it not} be fixed;
indeed, almost periodic potentials are allowed.
\begin{example} \label{e4.5.6} Let $\alpha_n$ be a sequence of rationals and
$\alpha=\lim_{n\to\infty}\,\alpha_n$. Let $H_n$ be the Jacobi
matrix with potential $\lambda\cos (2\pi\alpha_{n}+\theta)$ for $\lambda,
\theta$ fixed. Then [2] have shown for $|\lambda|\leq 2$, $|S_{n}|
\geq 4-2|\lambda|$. It follows from Theorem \ref{t4.5.5} that $|S|\geq
4-2|\lambda|$, providing a new proof (and a strengthening) of a result of
Last \cite{La93}. At present much more is known about this example and the interested reader may want to consult the survey by Last \cite{La05} for additional results.
\end{example}
\begin{example} \label{e4.5.7} Let $\{a_{m}\}_{m\in{\mathbb{N}}}$ be a sequence with
$s=\sum^{\infty}_{m=1}2^{m}|a_{m}|<2$. Let
$V(n)=\sum^{\infty}_{m=1}a_{m}\cos(2\pi n/2^{m})$, a limit periodic
potential on ${\mathbb{Z}}$. Let $h$ be the corresponding Jacobi matrix, then
one can show that
$
|\sigma_{\rm ac}(h)|\geq 2(2-s).
$
\end{example}
\smallskip
\begin{center}
$\bigstar$ \qquad $\bigstar$ \qquad $\bigstar$
\end{center}
\smallskip
Next we very briefly turn to higher-order trace formulas derived in \cite{GHSZ95} obtained by
higher-order expansions in \eqref{4.3.2} as $t\downarrow 0$. For
simplicity we now assume that $V\in C^\infty({\mathbb{R}})$ is bounded from
below. Then \eqref{4.3.2} can be extended to
$$
\Tr[e^{tH_{x}^D}-e^{-tH}]
\operatornamewithlimits{\sim}\limits_{t\downarrow 0}
\sum^{\infty}_{j=0} s_{j}(x)t^{j}, \quad x\in{\mathbb{R}}.
$$
Similarly, one has,
\begin{align*}
& \Tr[(H_{x}^D-z)^{-1}-(H-z)^{-1}]\operatornamewithlimits{\sim}
\limits_{z\downarrow -\infty} \sum^{\infty}_{j=0} r_{j}(x)z^{-j-1}, \\
& r_{0}(x)=1/2, \quad r_{1}(x)=V(x)/2, \quad x\in{\mathbb{R}}
\end{align*}
and one can show that
$$
s_{j}(x)=(-1)^{j+1}(j!)^{-1}r_{j}(x), \quad j\in{\mathbb{N}}\cup \{0\}.
$$
In particular, $r_j(x)$ and $s_{j}(x)$ are the celebrated KdV invariants
(up to inessential numerical factors). They can be computed recursively
(see, e.g., \cite{GHSZ95}). The higher-order analogs of \eqref{4.3.1}
then read
\begin{align*}
s_{0}(x) &=-\frac12 , \\
s_{j}(x) &=\frac{(-1)^{j+1}}{j!} \left\{\frac{E^{j}_{0}}{2}\right. + j \,
\lim_{t\downarrow 0} \int^{\infty}_{E_0}d\lambda\,
e^{-t\lambda}\lambda^{j-1} \biggl[\frac12 -\xi(\lambda, x)\biggr]
\quad j\in{\mathbb{N}},\, x\in{\mathbb{R}}.
\end{align*}
and similarly using a resolvent rather than a heat kernel regularization,
\begin{align*}
&r_{1}(x) =\frac12 \,V(x)
=\frac{E_0}{2}+\lim_{z\to i\infty}\int^{\infty}_{E_0}
d\lambda\, \frac{z^2}{(\lambda -z)^{2}}\biggl[\frac12 -\xi(\lambda,
x)\biggr], \\
&r_{j}(x)=\frac{E^{j}_{0}}{2}+\lim_{z\to
i\infty}\int^{\infty}_{E_0}
d\lambda\,\frac{z^{j+1}}{(\lambda-z)^{j+1}}\,j(-\lambda)^{j-1}
\biggl[\frac12 -\xi(\lambda, x)\biggr], \quad j\in{\mathbb{N}},\,
x\in{\mathbb{R}}.
\end{align*}
In the special periodic case, the corresponding extension of
\eqref{4.1.3} then reads
$$
2(-1)^{j+1}j!\, s_{j}(x)=2r_{j}(x)=E^{j}_{0}+\sum^{\infty}_{n=1}
[E^{j}_{2n-1}+E^{j}_{2n}-2\mu_{n}(x)^{j}], \quad j\in{\mathbb{N}},\, x\in{\mathbb{R}}.
$$
The latter formulas were originally found in \cite{Fl75} and \cite{MM75}.
We also note that the use of the Dirichlet boundary boundary
condition $u(x)=0$ and hence the choice of the Dirichlet operator
$H_x^D$ in connection with \eqref{4.1.7} can be replaced by any
self-adjoint boundary condition of the type
$u'(x_{\pm})+\beta u(x_{\pm})=0$, $\beta\in{\mathbb{R}}$, and the corresponding
Schr\"odinger operator $H_x^\beta$ in $L^{2}((-\infty, x))\oplus
L^{2}((x,\infty))$. This is worked out in detail in \cite{GHSZ95}.
Additional results on trace formulas for Schr\"odinger operators were presented in
\cite{Ge95}, \cite{GH95}, \cite{GH97}, \cite{GM00}, \cite{GRT96}.
\medskip
{\it More recent references:} Important extensions of the trace formula \eqref{4.3.1},
including the case of Schr\"odinger operators unbounded from below, were
discussed by Rybkin \cite{Ry01}, \cite{Ry01a}. Further discussions of the
trace formula \eqref{4.4.5} for Jacobi operators can be found in
\cite{GS97a} and Teschl \cite{Te98}, \cite[Ch.\ 6]{Te00}. An extension of
Corollary \ref{c4.4.2} to Schr\"odinger operators on a countable set was
discussed by Shirai \cite{Sh98}.
Removal of the resolvent regularization limit in the above trace formula for
$r_1$ (resp., $V$) under optimal conditions on $V$ has been studied by
Rybkin \cite{Ry01a}, \cite{Ry03} (the latter reference offers necessary
and sufficient conditions for absolute summability of the trace formula).
A certain multi-dimensional variant of these trace formulas, inspired by
work of Lax \cite{La94}, was established in \cite{GHSZ96} (see also
\cite{GH95}).
Matrix-valued extensions of the trace formula for Schr\"odinger,
Dirac-type, and Jacobi matrices, as well as Borg and Hochstadt-type
theorems were studied in \cite{BGMS03}, \cite{CG01}, \cite{CG02},
\cite{CGHL00}, \cite{CGR05}, \cite{GH97}, and \cite{GS03}.
Trace formulas and an ensuing general Borg-type theorem for CMV operators
(i.e., in connection with orthogonal polynomials on the unit circle, cf.\
\cite{Si05}) appeared in \cite{GZ05}.
An application of $\xi$-function ideas to obtain Weyl-type
asymptotics using $\zeta$-function regularizations of determinants of
certain operators on complete Riemannian manifolds can be found in
Carron \cite{Ca02}.
Theorem \ref{t4.5.2} was used in \cite{GKT96} to solve an inverse
spectral problem for Jacobi matrices and most recently in \cite{GY06} in
connection with proving purely absolutely continuous spectrum of a class
of reflectionless Schr\"odinger operators with homogeneous spectrum. It
has also recently been discussed in \cite[Sect.\ 1.5]{DK05}.
\section{Various Uniqueness Theorems in Inverse Spectral Theory}
\label{s5}
In this section we summarize some of the principal results of
the following papers: \\
\cite{DGS97} R.~del Rio, F.~Gesztesy, and B.~Simon, {\it Inverse
spectral analysis with partial information on the potential, III.
Updating boundary conditions}, Intl. Math. Research Notices {\bf
1997, No.\ 15}, 751--758. \\
\cite{DGS99} R.~del Rio, F.~Gesztesy, and B.~Simon, {\it Corrections and
Addendum to ``Inverse spectral analysis with partial information on the
potential, III. Updating boundary conditions''}, Intl. Math. Research
Notices {\bf 1999, No.\ 11}, 623--625. \\
\cite{GS96a} F.~Gesztesy and B.~Simon, {\it Uniqueness theorems in inverse
spectral theory for one-dimensional Schr\"odinger operators},
Trans. Amer. Math. Soc. {\bf 348}, 349--373 (1996). \\
\cite{GS97} F.~Gesztesy and B.~Simon, {\it Inverse spectral
analysis with partial information on the potential, I. The
case of an a.c.\ component in the spectrum}, Helv. Phys. Acta
{\bf 70}, 66--71, 1997. \\
\cite{GS97a} F.~Gesztesy and B.~Simon, {\it $m$-functions and inverse
spectral analysis for finite and semi-infinite Jacobi matrices},
J. Analyse Math. {\bf 73}, 267--297 (1997). \\
\cite{GS99} F.~Gesztesy and B.~Simon, {\it On the determination of a
potential from three spectra}, in {\it Differential Operators and Spectral Theory},
V.\ Buslaev, M.\ Solomyak, and D.\ Yafaev (eds.), Amer. Math. Soc.
Transl. Ser.\ 2, {\bf 189}, 85--92 (1999). \\
\cite{GS00} F.~Gesztesy and B.~Simon, {\it Inverse spectral analysis with
partial information on the potential, II. The case of discrete spectrum},
Trans. Amer. Math. Soc. {\bf 352}, 2765--2787 (2000).
\medskip\smallskip
One can argue that inverse spectral theory, especially, the case of uniqueness theorems in inverse spectral theory, started with the paper by Ambarzumian
\cite{Am29} in 1929 and was turned into a full fledged discipline by the seminal 1946 paper by Borg \cite{Bo46}. Ambarzumian proved the special uniqueness theorem that if the eigenvalues of a Schr\"odinger operator $-d^2/dx^2 + V$ in $L^2([0,\pi])$ with Neumann boundary conditions at the endpoints $x=0$ and $x=\pi$ coincide with the sequence of numbers $n^2$, $n=0,1,2,\dots$, then $V=0$ a.e.\ on $[0,\pi]$. This result is very special. Indeed, Borg showed that for more general boundary conditions, one set of eigenvalues, in general (i.e., in the absence of symmetries of $V$), is insufficient to determine $V$ uniquely. Moreover, he described in great detail when two spectra guarentee unique determination of the potential $V$. In this section we will discuss Borg's celebrated two-spectra uniqueness result and many of its extension due to Gasymov, Hald, Hochstadt, Levitan, Lieberman, Marchenko, and Simon and collaborators.
We start with paper \cite{GS96a}. It contains a variety of
new uniqueness theorems for potentials $V$ in one-dimensional
Schr\"odinger operators $-\frac{d^2}{dx^2}+V$ on ${\mathbb{R}}$ and on
the half-line ${\mathbb{R}}_{+}=[0,\infty)$ in terms of appropriate
spectral shift functions introduced in a series of papers
describing new trace formulas for $V$ on ${\mathbb{R}}$
\cite{GHS95}, \cite{GHSZ93}, \cite{GHSZ95}, \cite{GS96} and on
${\mathbb{R}}_{+}$ \cite{GH95}. In particular, it contains a generalization
of a well-known uniqueness theorem of Borg and Marchenko for Schr\"odinger
operators on the half-line with purely discrete spectra to arbitrary
spectral types and a new uniqueness result for Schr\"odinger operators
with confining potentials on the entire real line.
Turning to the half-line case first, we recall one of the principal
uniqueness results proved in \cite{GS96a}, which extends a well-known
theorem of Borg \cite{Bo52} and Marchenko \cite{Ma73} in the special
case of purely discrete spectra to arbitrary spectral types. We suppose
\begin{equation}
V\in L^{1}([0,R]) \text{ for all $R>0$}, \quad V\text{ real-valued,}
\label{5.2.1}
\end{equation}
and introduce the differential expression
$$
\tau_{+}=-\frac{d^2}{dx^2}+V(x), \quad x\geq 0,
$$
for simplicity assuming that $\tau$ is in the limit point case at
$\infty$. (We refer to \cite{GS96a} for a general treatment that includes
the limit circle case.) Associated with
$\tau_+$ one introduces the following self-adjoint operator $H_{+,\alpha}$
in $L^{2}({\mathbb{R}}_{+})$.
\begin{align}
&H_{+,\alpha}f=\tau_{+}f, \quad \alpha\in [0,\pi), \notag \\
&f\in\text{\rm{dom}}(H_{+,\alpha})=\{g\in L^{2}({\mathbb{R}}_{+}) \,|\, g,g'\in
AC([0,R])\text{ for all } R>0; \label{5.2.2} \\
& \hspace*{3.02cm}
\sin(\alpha)g'(0_{+})+\cos(\alpha)g(0_{+})=0; \, \tau_{+}g\in
L^{2}({\mathbb{R}}_{+})\}. \notag
\end{align}
Then $H_{+,\alpha}$ has uniform spectral multiplicity one.
Next we introduce the fundamental system $\phi_{\alpha}(z,x)$,
$\theta_{\alpha}(z,x)$, $z\in{\mathbb{C}}$, of solutions of
$\tau_{+}\psi(z,x)=z\psi(z,x)$, $x\geq 0$, satisfying
$$
\phi_{\alpha}(z,0)=-\theta'_{\alpha}(z,0)=-\sin(\alpha), \quad
\phi'_{\alpha}(z,0)=\theta_{\alpha}(z,0)=\cos(\alpha)
$$
such that $W(\theta_{\alpha}(z), \phi_{\alpha}(z))=1$.
Furthermore, let
$\psi_{+,\alpha}(z,x)$, $z\in{\mathbb{C}}\backslash{\mathbb{R}}$ be the unique
solution of $\tau\psi(z)=\psi(z)$ which satisfies
$$
\psi_{+,\alpha}(z,\,\cdot\,)\in L^{2}({\mathbb{R}}_{+}), \quad
\sin(\alpha)\psi'_{+,\alpha}(z,0_{+}) + \cos(\alpha) \psi_{+,\alpha}
(z,0_{+})=1.
$$
$\psi_{+,\alpha}$ is of the form
$$
\psi_{+,\alpha}(z,x)=\theta_{\alpha}(z,x)+m_{+,\alpha}(z)
\phi_{\alpha}(z,x)
$$
with $m_{+,\alpha}(z)$ the half-line Weyl--Titchmarsh $m$-function. Being
a Herglotz function (i.e., an analytic function in the open upper-half
plane that maps the latter to itself)),
$m_{+,\alpha}(z)$ has the following representation in terms of a positive
measure $d\rho_{+,\alpha}$,
$$
m_{+,\alpha} = \begin{cases} a_{+,\alpha}
+\int_{{\mathbb{R}}} \biggl[\frac{1}{\lambda-z}-\frac{\lambda}
{1+\lambda^2}\biggr]\,d\rho_{+,\alpha}
(\lambda), \quad \alpha\in [0,\pi), \\
\cot(\alpha)+\int_{{\mathbb{R}}}(\lambda-z)^{-1} d\rho_{+,\alpha}
(\lambda), \quad \alpha\in (0,\pi). \end{cases}
$$
The basic uniqueness criterion for Schr\"odinger operators on the
half-line $[0,\infty)$, due to Marchenko \cite{Ma73}, that we shall rely
on repeatedly in the following, can be stated as follows.
\begin{theorem} \label{t5.2.1} Suppose
$\alpha_{1}, \alpha_{2}\in [0,\pi)$, $\alpha_{1}\neq\alpha_{2}$ and
define $H_{+, j, \alpha_{j}}$, $m_{+, j, \alpha_{j}}$,
$\rho_{+,j,\alpha_{j}}$ associated with the differential expressions
$\tau_{j}=-\frac{d^2}{dx^2}+V_{j}(x)$, $x\geq 0$, where $V_{j}, j=1,2$
satisfy assumption \eqref{5.2.1}. Then the following three assertions are
equivalent:
\\
$(i)$ $m_{+,1,\alpha_{1}}(z)=m_{+,2,\alpha_{2}}(z)$, $z\in{\mathbb{C}}_{+}$. \\
$(ii)$ $\rho_{+,1,\alpha_{1}}((-\infty,\lambda])=
\rho_{+,2,\alpha_{2}} ((-\infty,\lambda])$, $\lambda\in{\mathbb{R}}$. \\
$(iii)$ $\alpha_{1}=\alpha_{2}$ and $V_{1}(x)=V_{2}(x)$ for
a.e.~$x\geq 0$.
\end{theorem}
Next we relate Green's functions for different boundary conditions
at $x=0$.
\begin{lemma} \label{l5.2.2} Let $\alpha_{j}\in [0,\pi)$, $j=1,2$, $x,x'\in
{\mathbb{R}}_{+}$, and $z\in{\mathbb{C}}\backslash\{\sigma(H_{+,\alpha_{1}})\cup
\sigma(H_{+,\alpha_{2}})\}$. Then,
\begin{align*}
&G_{+,\alpha_{2}}(z,x,x')-G_{+,\alpha_{1}}(z,x,x')=-
\frac{\psi_{+,\alpha_{1}}(z,x)\psi_{+,\alpha_{1}}(z,x')}
{\cot(\alpha_{2}-\alpha_{1})+m_{+, \alpha_{1}}(z)}, \\
&\frac{G_{+,\alpha_{2}}(z,0,0)}{G_{+,\alpha_{1}}(z,0,0)} =\frac{1}
{(\beta_{1}-\beta_{2})\sin^{2}(\alpha_{1})[\cot(\alpha_{2}-\alpha_{1})
+m_{+,\alpha_{1}}(z)]} \\
&\quad = (\beta_{1}-
\beta_{2})\sin^{2}(\alpha_{2})[\cot(\alpha_{2}-\alpha_{1})-
m_{+,\alpha_{2}}(z)],\quad \beta_{j}=\cot(\alpha_{j}), j=1,2 \\
& \Tr [(H_{+,\alpha_{2}}-z)^{-1}
-(H_{+,\alpha_{1}}-z)^{-1}] =
-\frac{d}{dz}\,\ln[\cot(\alpha_{2}-\alpha_{1})+m_{+,\alpha_{1}}(z)] \\
&\hspace*{5.42cm} =\frac{d}{dz}\,
\ln [\cot(\alpha_{2}-\alpha_{1})-m_{+,\alpha_{2}}(z)].
\end{align*}
\end{lemma}
Since $m_{+,\alpha}(z)$ is a Herglotz function, we may now introduce
spectral shift function [27] $\xi_{\alpha_{1},
\alpha_{2}}(\lambda)$ for the pair $(H_{+,\alpha_{2}}, H_{+,
\alpha_{1}})$ via the exponential Herglotz representation of
$m_{+,\alpha}(z)$ (cf.\ \cite{AD56})
\begin{align*}
& \cot(\alpha_{2}-\alpha_{1})+m_{+,\alpha_{1}}(z)
=\exp\biggl\{\text{Re}[\ln
(\cot(\alpha_{2}-\alpha_{1})+m_{+,\alpha_{1}}(i))] \\
&\quad + \int_{{\mathbb{R}}} \biggl[\frac{1}{\lambda-z}-
\frac{\lambda}{1+\lambda^2}\biggr]\xi_{\alpha_{1}, \alpha_{2}}
(\lambda)\,d\lambda\biggr\}, \quad 0\leq\alpha_{1} <\alpha_{2} <\pi,
z\in{\mathbb{C}}\backslash{\mathbb{R}}.
\end{align*}
This is extended to all $\alpha_{1},\alpha_{2}\in [0,\pi)$ by
$$
\xi_{\alpha,\alpha} (\lambda)=0, \quad \xi_{\alpha_{2}, \alpha_{1}}
(\lambda)=-\xi_{\alpha_{1}, \alpha_{2}}(\lambda)
\text{ for a.e.~$\lambda\in{\mathbb{R}}$}.
$$
Next we summarize a few properties of $\xi_{\alpha_{1},\alpha_{2}}
(\lambda)$.
\begin{lemma} \label{l5.2.3} $(i)$ Suppose $0\leq\alpha_{1} <\alpha_{2}
<\pi$. Then for a.e.~$\lambda\in{\mathbb{R}}$,
$$
\xi_{\alpha_{1},\alpha_{2}}(\lambda) = \begin{cases}
\lim\limits_{\epsilon\downarrow 0}\,\pi^{-1} \Im
\{\ln[\cot(\alpha_{2}-\alpha_{1})+m_{+,\alpha_{1}}(\lambda+
i\epsilon)]\}, \\
-\lim\limits_{\epsilon\downarrow 0}\,\pi^{-1}\Im \{\ln
[\cot(\alpha_{2}-\alpha_{1})-m_{+,\alpha_{2}}(\lambda+
i\epsilon)]\}, \\
\lim\limits_{\epsilon\downarrow 0}\,\pi^{-1}\Im \bigl\{\ln
\bigl[\tfrac{1}
{\sin(\alpha_{1})}\,\tfrac{G_{+,\alpha_{1}}(\lambda +
i\epsilon, 0, 0)}{G_{+,\alpha_{2}}(\lambda + i\epsilon, 0,0)}\bigr]
\bigr\}.
\end{cases}
$$
$($For $\alpha_{1}=0$,
$G_{+,\alpha_{1}}(\lambda+i\epsilon,0,0)/\sin
(\alpha_{1})$ has to be replaced by $-1$ in the last expression.$)$
Moreover,
$$
0\leq \xi_{\alpha_{1},\alpha_{2}}(\lambda)\leq 1 \, \text{ a.e.}
$$
$(ii)$ Let $\alpha_{j}\in [0,\pi)$, $1\leq j\leq 3$. Then the
``chain rule''
$$
\xi_{\alpha_{1},\alpha_{3}}(\lambda)=\xi_{\alpha_{1},
\alpha_{2}}(\lambda)
+\xi_{\alpha_{2},\alpha_{3}}(\lambda)
$$
holds for a.e.~$\lambda\in{\mathbb{R}}$. \\
$(iii)$ For all $\alpha_{1},\alpha_{2}\in [0,\pi)$,
$$
\xi_{\alpha_{1},\alpha_{2}}\in L^{1}({\mathbb{R}}; (1+\lambda^{2})^{-1}\,
d\lambda).
$$
$(iv)$ Assume $\alpha_{1}, \alpha_{2}\in [0,\pi)$,
$\alpha_{1}\neq \alpha_{2}$. Then,
$$
\xi_{\alpha_{1},\alpha_{2}}\in L^{1}({\mathbb{R}}; (1+|\lambda|)^{-1}
d\lambda) \, \text{ if and only if } \, \alpha_{1},\alpha_{2}\in
(0,\pi).
$$
$(v)$ For all $\alpha_{1},\alpha_{2}\in [0,\pi)$,
$$
\Tr [(H_{+,\alpha_{2}}-z)^{-1} -(H_{+,\alpha_{1}}-z)^{-1}]
=-\int_{{\mathbb{R}}} \frac{d\lambda \,
\xi_{\alpha_{1},\alpha_{2}} (\lambda)}{(\lambda-z)^{-2}}.
$$
\end{lemma}
We note that $\xi_{\alpha_{1}, \alpha_{2}}(\lambda)$ (for $\alpha_{1},
\alpha_{2}\in (0,\pi)$) has been introduced by Javrjan \cite{Ja66},
\cite{Ja71}. In particular, he proved Lemma \ref{l5.2.2}\,$(iii)$ and
Lemma \ref{l5.2.3}\,$(v)$ in the non-Dirichlet cases where
$0<\alpha_{1},\alpha_{2}<\pi$.
Given these preliminaries, we are now able to state the main
uniqueness result for half-line Schr\"odinger operators of \cite{GS96a}.
\begin{theorem} \label{t5.2.4} Suppose $V_j$ satisfy assumption
\eqref{5.2.1} and define $H_{+,j,\alpha_{j,\ell}}$,
$j,\ell=1,2$, associated with the differential expressions
$\tau_{j}=-\frac{d^2} {dx^2}+V_{j}(x)$, $x\geq 0$, $j=1,2$, where
$\alpha_{j,\ell}\in [0,\pi)$,
$\ell=1,2$, and we suppose $0\leq \alpha_{1,1} <\alpha_{1,2} <\pi$, $0\leq
\alpha_{2,1} <\alpha_{2,2} <\pi$. In
addition, let $\xi_{j,\alpha_{j,1},\alpha_{j,2}}$, $j=1,2$ be the
spectral shift function for the pair $(H_{+,j,\alpha_{j,1}},
H_{+,j,\alpha_{j,2}})$. Then the following two assertions are equivalent:
\\
$(i)$ $\xi_{1,\alpha_{1,1},\alpha_{1,2}}(\lambda) =
\xi_{2,\alpha_{2,1},\alpha_{2,2}}(\lambda)$ for a.e.~$\lambda\in{\mathbb{R}}$. \\
$(ii)$ $\alpha_{1,1}=\alpha_{2,1}$,
$\alpha_{1,2}=\alpha_{2,2}$, and $V_{1}(x)=V_{2}(x)$ for a.e.~$x\geq
0$.
\end{theorem}
As a corollary, one obtains a well-known uniqueness result
originally due to Borg \cite[Theorem\ 1]{Bo52} and Marchenko
\cite[Theorem\ 2.3.2]{Ma73} (see also \cite{LG64}).
\begin{corollary} \label{c5.2.5} Define
$\tau_j$, $H_{+,j,\alpha}$, $\alpha\in [0,\pi)$ as in Theorem
\ref{t5.2.4}. Assume
in addition that $H_{+,1,\alpha_{1}}$ and $H_{+,2,\alpha_{2}}$
have purely discrete spectra for some $($and hence for all$)$
$\alpha_{j}\in [0,\pi)$, that is,
$$
\sigma_{\rm ess} (H_{+,j,\alpha_{j}})=\emptyset
\, \text{ for some } \, \alpha_{j}\in [0,\pi), j=1,2.
$$
Then the following two assertions are equivalent: \\
$(i)$ $\sigma(H_{+,1,\alpha_{1,1}})=\sigma
(H_{+,2,\alpha_{2,1}})$, $\sigma(H_{+,1,\alpha_{1,2}})=
\sigma (H_{+,2,\alpha_{2,2}})$,
$\alpha_{j,\ell}\in [0,\pi)$, $j,\ell=1,2$, $\sin(\alpha_{1,1}-
\alpha_{1,2})\neq 0$. \\
$(ii)$ $\alpha_{1,1}=\alpha_{2,1}$,
$\alpha_{1,2}=\alpha_{2,2}$, and $V_{1}(x)=V_{2}(x)$ for
a.e.~$x\geq 0$.
\end{corollary}
Roughly speaking, Corollary \ref{c5.2.5} implies that two sets of purely
discrete spectra $\sigma(H_{+,\alpha_{1}}), \sigma(H_{+,\alpha_{2}})$
associated with distinct boundary conditions at $x=0$ (but a fixed
boundary condition (if any) at $+\infty$), that is,
$\sin(\alpha_{2}-\alpha_{1})\neq 0$, uniquely determine $V$ a.e. The
first main result in \cite{GS96a}, Theorem \ref{t5.2.4}, removes all a
priori spectral hypotheses and shows that the spectral shift function
$\xi_{\alpha_{1}, \alpha_{2}}(\lambda)$ for the pair $(H_{+,\alpha_{2}},
H_{+,\alpha_{1}})$ with distinct boundary conditions at $x=0$,
$\sin(\alpha_{2}-\alpha_{1})\neq 0$, uniquely determines $V$ a.e. This
illustrates that Theorem \ref{t5.2.4} is the natural generalization of
Borg's and Marchenko's theorems from the discrete spectrum case to
arbitrary spectral types.
Now we turn to uniqueness results for Schr\"odinger operators
on the whole real line. We shall rely on the notation $\tau$,
$\phi_{\alpha}$, $\theta_\alpha$, $\psi_{\pm,\alpha}$, $m_{\pm,\alpha}$,
$d\rho_{\pm,\alpha}$, which are defined in complete analogy to the
half-line case (with $x\in{\mathbb{R}}$), and we shall assume
\begin{equation}
V\in L^{1}_{\text{loc}}({\mathbb{R}}),
\quad V\text{ real-valued}. \label{5.3.1}
\end{equation}
Following \cite{GHSZ95}, we introduce, in
addition, the following family of self-adjoint operators $H^{\beta}_{y}$
in
$L^{2}({\mathbb{R}})$,
\begin{align*}
& H^{\beta}_{y}f=\tau f, \quad \beta\in{\mathbb{R}}\cup\{\infty\}, \quad
y\in{\mathbb{R}}, \\
& \text{\rm{dom}} (H^{\beta}_{y})
=\{g\in L^{2}({\mathbb{R}}\,|\, g,g'\in AC([y,\pm
R])\text{ for all }R>0;\, g'(y_{\pm})+\beta g(y_{\pm})=0; \\
& \hspace*{3.7cm} \lim\limits_{R\to\pm\infty}\,W(f_{\pm}(z_{\pm}),g)(R)=0;
\, \tau g\in L^{2}({\mathbb{R}})\}.
\end{align*}
Thus $H^{D}_{y}:=H^{\infty}_{y}$ (resp., $(H^{N}_{y}:=H^{0}_{y})$
corresponds to the Schr\"odinger operator with an additional Dirichlet
(resp., Neumann) boundary condition at $y$. In obvious notation,
$H^{\beta}_{y}$ decomposes into the direct sum of half-line operators
$$
H^{\beta}_{y}=H^{\beta}_{-,y}\oplus H^{\beta}_{+,y}
$$
with respect to
$
L^{2}({\mathbb{R}})=L^{2}((-\infty, y]) \oplus L^{2}([y,\infty))
$.
In particular, $H^{\beta}_{+,y}$ equals $H_{+,\alpha}$ for $\beta
=\cot(\alpha)$ (and $y=0$)) in our notation \eqref{5.2.2} and, as
done in our previous Sections \ref{s2} and \ref{s4}, the reference
point $y$ will be added as a subscript to obtain $\theta_{\alpha,
y}(z,x)$, $\phi_{\alpha, y}(z,x)$, $\psi_{\pm,\alpha,y}(z,x)$,
$m_{\pm,\alpha,y}(z)$, $M_{\alpha, y}(z)$, etc.
Next, we recall a few results from \cite{GHSZ95}. With $G(z,x,x')$ and
$G^{\beta}_{y}(z,x,x')$ the Green's functions of $H$ and
$H^{\beta}_{y}$, one obtains (for $z\in{\mathbb{C}}\backslash
\{\sigma(H^{\infty}_{y})\cup\sigma(H)\}$)
\begin{align*}
G^{\beta}_{y}(z,x,x') &=G(z,x,x')
-\frac{(\beta+\partial_{2})G(z,x,y)
(\beta+\partial_{1}) G(z,y,x')}{(\beta+\partial_{1})
(\beta+\partial_{2})G(z,y,y)}, \\
&\hspace*{3.4cm} \beta\in{\mathbb{R}}, z\in{\mathbb{C}}\backslash
\{\sigma(H^{\beta}_{y})\cup\sigma(H)\}, \\
G^{\infty}_{y}(z,x,x') &= G(z,x,x')-G(z,y,y)^{-1} G(z,x,y)
G(z,y,x').
\end{align*}
Here we abbreviated
\begin{align*}
& \partial_{1}G(z,y,x')
=\partial_{x}G(z,x,x') |_{x=y}, \quad \partial_{2}(G,z,x,y)
=\partial_{x'}G(z,x,x') |_{x'=y}, \\
& \partial_{1}\partial_{2} G(z,y,y)
=\partial_{x}\partial_{x'}
G(z,x,x') |_{x=y=x'}, \, \text{ etc.} \\
& \partial_{1}G(z,y,x)=\partial_{2}G(z,x,y), \quad x\neq y.
\end{align*}
As a consequence,
\begin{align*}
&\Tr [(H^{\beta}_{y}-z)^{-1}-(H-z)^{-1}]
=-\frac{d}{dz}\,\ln [(\beta+\partial_{1})(\beta+\partial_{2})G(z,y,y)],
\quad \beta\in{\mathbb{R}}, \\
&\Tr [(H^{\infty}_{y}-z)^{-1}-(H-z)^{-1}] =-\frac{d}{dz}\,\ln
[G(z,y,y)].
\end{align*}
In analogy to the Herglotz property of $G(z,y,y)$,
$(\beta+\partial_{1})(\beta+\partial_{2}) G(z,y,y)$ is also Herglotz
for each $y\in{\mathbb{R}}$. Hence, both admit exponential representations of the
form
\begin{align*}
& G(z,y,y)=\exp\biggr\{c_{\infty}+\int_{{\mathbb{R}}} d\lambda\,
\biggl[\frac
{1}{\lambda-z}-\frac{\lambda}{1+\lambda^2}\biggl] \xi^{\infty}
(\lambda, y)\biggr\}, \\
& c_{\infty}\in{\mathbb{R}}, \quad 0\leq\xi^{\infty}(\lambda, y)\leq 1
\text{ a.e.}, \\
& \xi^{\infty}(\lambda, y)
=\lim\limits_{\epsilon\downarrow 0}\, \pi^{-1}
\Im \{\ln [G(\lambda+i\epsilon, y, y)]\}
\text{ for a.e.~$\lambda\in{\mathbb{R}}$}, \\
& (\beta+\partial_{1})(\beta+\partial_{2})G(z,y,y)
=\exp\biggl\{c_{\beta}
+\int_{{\mathbb{R}}} d\lambda \,
\biggl[\frac{1}{\lambda-z}-\frac{\lambda}
{1+\lambda^2}\biggr] [\xi^{\beta}(\lambda, y)+1]\biggr\}, \\
& c_{\beta}\in{\mathbb{R}}, \quad -1\leq\xi^{\beta}(\lambda, y)\leq 0
\text{ a.e.}, \quad \beta\in{\mathbb{R}}, \\
& \xi^{\beta}(\lambda, y)=
\lim\limits_{\epsilon\downarrow 0}\, \pi^{-1}
\Im \{\ln [(\beta+\partial_{1})(\beta+\partial_{2})
G(\lambda+i\epsilon, y, y)]\}-1, \quad \beta\in{\mathbb{R}}
\end{align*}
for each $y\in{\mathbb{R}}$. Moreover,
$$
\Tr [(H^{\beta}_{y}-z)^{-1}-(H-z)^{-1}]=-\int_{{\mathbb{R}}} d\lambda \,
(\lambda-z)^{-2}\xi^{\beta}(\lambda, y), \quad
\beta\in{\mathbb{R}}\cup\{\infty\}.
$$
Applying the basic uniqueness criterion for Schr\"odinger operators to
both half-lines $(-\infty,y]$ and $[y,\infty)$ then yields the following
principal characterization result for Schr\"odinger operators on
${\mathbb{R}}$ first proved in \cite{GS96a}.
\begin{theorem} \label{t5.3.2} Let $\beta_{1},\beta_{2}\in{\mathbb{R}}\cup
\{\infty\}$, $\beta_{1}\neq\beta_{2}$, and $x_{0}\in{\mathbb{R}}$. Then the following
assertions hold: \\
$(i)$ $\xi^{\beta_1}(\lambda, x_{0})$ and $\xi^{\beta_2}
(\lambda, x_{0})$ for a.e.~$\lambda\in{\mathbb{R}}$ uniquely determine
$V(x)$ for a.e.~$x\in{\mathbb{R}}$ if the pair $(\beta_{1},\beta_{2})$
differs from $(0,\infty)$, $(\infty,0)$. \\
$(ii)$ If $(\beta_{1},\beta_{2})=(0,\infty)$ or $(\infty,
0)$, assume in addition that $\tau$ is in the limit point case at
$+\infty$ and $-\infty$. Then $\xi^{\infty}(\lambda, x_{0})$ and
$\xi^{0}(\lambda, x_{0})$ for a.e.~$\lambda\in{\mathbb{R}}$ uniquely
determine $V$ a.e.~up to reflection symmetry with respect to $x_0$;
that is, both $V(x)$, $V(2x_{0}-x)$ for a.e.~$x\in{\mathbb{R}}$ correspond
to $\xi^{\infty}(\lambda, x_{0})$ and $\xi^{0}(\lambda, x_{0})$ for
a.e.~$\lambda\in{\mathbb{R}}$.
\end{theorem}
\begin{corollary} \label{c5.3.3} Suppose $\tau$ is in the limit point case at
$+\infty$ and $-\infty$ and let $\beta\in{\mathbb{R}}\cup\{\infty\}$ and
$x_{0}\in{\mathbb{R}}$. Then $\xi^{\beta}(\lambda, x_{0})$ for
a.e.~$\lambda\in{\mathbb{R}}$ uniquely determines $V(x)$ for a.e.~$x\in{\mathbb{R}}$
if and only if $V$ is reflection symmetric with respect to $x_0$,
that is, $V(2x_{0}-x)=V(x)$ a.e.
\end{corollary}
In view of Corollary \ref{c5.2.5}, it seems appropriate to formulate
Theorem \ref{t5.3.2} in the special case of purely discrete spectra.
\begin{corollary} \label{c5.3.4} Suppose $H$ $($and hence $H^{\beta}_{y}$
for all $y\in{\mathbb{R}}$, $\beta\in{\mathbb{R}}\cup\{\infty\}$$)$ has purely
discrete spectrum, that is, $\sigma_{\text{\rm{ess}}}(H)=\emptyset$
and let $\beta_{1},\beta_{2}\in{\mathbb{R}}\cup\{\infty\}$, $\beta_{1}\neq
\beta_{2}$, and $x_{0}\in{\mathbb{R}}$. \\
$(i)$ $\sigma(H)$, $\sigma(H^{\beta_j}_{x_0})$, $j=1,2$
uniquely determine $V$ a.e.~if the pair $(\beta_{1},\beta_{2})$
differs from $(0,\infty)$ and $(\infty, 0)$. \\
$(ii)$ If $(\beta_{1},\beta_{2})=(0,\infty)$ or $(\infty,
0)$, assume in addition that $\tau$ is in the limit point case at
$+\infty$ and $-\infty$. Then $\sigma(H)$, $\sigma(H^{\infty}_{x_0})$,
and $\sigma(H^{0}_{x_0})$ uniquely determine $V$ a.e.~up to reflection
symmetry with respect to $x_0$, that is, both $V(x)$ and $\widehat{V}
(x)=V(2x_{0}-x)$ for a.e.~$x\in{\mathbb{R}}$ correspond to $\sigma(H)=\sigma
(\widehat{H})$, $\sigma(H^{\infty}_{x_0})=\sigma(\widehat{H}^{\infty}
_{x_0})$, and $\sigma(H^{0}_{x_0})=\sigma(\widehat{H}^{0}_{x_0})$.
Here, in obvious notation, $\widehat H$, $\widehat{H}^{\infty}_{x_0}$,
$\widehat{H}^{0}_{x_0}$ correspond to $\widehat{\tau}=-\frac{d^2}
{dx^2}+\widehat{V}(x)$, $x\in{\mathbb{R}}$. \\
$(iii)$ Suppose $\tau$ is in the limit point case at
$+\infty$ and $-\infty$ and let $\beta\in{\mathbb{R}}\cup\{\infty\}$. Then
$\sigma(H)$ and $\sigma(H^{\beta}_{x_0})$ uniquely determine $V$
a.e.~if and only if $V$ is reflection symmetric with respect to $x_0$. \\
$(iv)$ Suppose that $V$ is reflection symmetric with
respect to $x_0$ and that $\tau$ is nonoscillatory at $+\infty$ and $-
\infty$. Then $V$ is uniquely determined a.e.~by $\sigma(H)$ in the
sense that $V$ is the only potential symmetric with respect to $x_0$
with spectrum $\sigma(H)$.
\end{corollary}
Of course, Corollary \ref{c5.3.4}\,$(iii)$ is implied
by the result of Borg [5] and Marchenko [32] (see Corollary \ref{c5.2.5}
with $\alpha_{1}=0$, $\alpha_{2}=\pi/2$).
Thus far, we exclusively dealt with $\xi$-functions and spectra in
connection with uniqueness theorems. A variety of further uniqueness
results can be obtained by invoking alternative information such as
the left/right distribution of $\lambda^{\beta}_{n}(x_{0})$ (i.e.,
whether $\lambda^{\beta}_{n}(x_{0})$ is an eigenvalue of $H^{\beta}_
{-,x_{0}}$ in $L^{2}((-\infty, x_{0}])$ or of $H^{\beta}_{+,x_{0}}$ in
$L^{2}([x_{0},\infty))$) and/or associated norming constants. For details
we refer to the discussion in \cite{GS96a}.
\medskip
{\it More recent references:} Uniqueness theorems related to Theorem \ref{t5.2.4} in the short-range case with spectral shift data replaced by scattering data were studied by Aktosun and Weder \cite{AW06}, \cite{AW07}. Analogs of Corollaries \ref{c5.2.5} and \ref{c5.3.4}\,$(i)$ for Jacobi operators were derived by Teschl \cite{Te98}.
\smallskip
\begin{center}
$\bigstar$ \qquad $\bigstar$ \qquad $\bigstar$
\end{center}
\smallskip
Next we focus on \cite{GS00}, which discussed results where the discrete
spectrum (or partial information on the discrete spectrum) and partial
information on the potential $V$ of a one-dimensional Schr\"odinger
operator $H=-\frac{d^2}{dx^2}+q$ determines the potential completely.
Included are theorems for finite intervals and for the whole line. In
particular, a new type of inverse spectral problem involving fractions
of the eigenvalues of $H$ on a finite interval and knowledge of $V$ over
a corresponding fraction of the interval was posed and solved in
\cite{GS00}. The methods employed in \cite{GS00} rest on Weyl--Titchmarsh
$m$-function techniques starting with the basic Borg--Marchenko uniqueness
result (cf.\ Theorems \ref{t6.1.5} and \ref{t6.1.6}) and densities of zeros of a class of entire functions since the $m$-functions are meromorhic functions in this context.
In 1978, Hochstadt and Lieberman \cite{HL78} proved the following
remarkable theorem:
\begin{theorem} \label{t6.1.1} Let $h_0 \in {\mathbb{R}}$, $h_1 \in {\mathbb{R}} \cup
\{\infty\}$ and assume $V_1, V_2 \in L^1 ((0,1))$ to be real-valued.
Consider the Schr\"odinger operators $H_1, H_2$ in $L^2 ([0,1])$ given by
$$
H_j = -\frac{d^2}{dx^2}+ V_j, \quad j=1,2,
$$
with the boundary conditions
\begin{align}
\begin{split}
u'(0) + h_0 u(0) & = 0, \label{6.1.1} \\
u'(1) + h_1 u(1) & = 0.
\end{split}
\end{align}
Let $\sigma(H_j)=\{\lambda_{j,n}\}$ be the $($necessarily simple$)$
spectra of $H_j, j=1,2$. Suppose that $V_1 = V_2$ a.e.\ on
$[0,1/2]$ and that $\lambda_{1,n} = \lambda_{2,n}$ for all $n$. Then
$V_1 = V_2$ a.e.\ on $[0,1]$.
\end{theorem}
Here, in obvious notation, $h_1 =\infty$ in \eqref{6.1.1} singles out
the Dirichlet boundary condition $u(1)=0$.
For each $\varepsilon >0$, there are simple examples where $V_1 = V_2$ on
$[0,(1/2) - \varepsilon]$ and $\sigma (H_1)$ = $\sigma (H_2)$ but $V_1
\neq V_2$. (Choose $h_0=-h_1$, $V_1 (x)=0$ for $x \in (0,(1/2) -
\varepsilon] \cup [1/2, 1]$ and nonzero on $((1/2) - \varepsilon,
1/2)$, and $V_2 (x) = V_1 (1-x)$. See also Theorem I$^\prime$ in the
appendix of \cite{Su86}.)
Later refinements of Theorem \ref{t6.1.1} in \cite{Ha80}, \cite{Su86}
(see also the summary in \cite{Su82}) showed that the boundary
condition for $H_1$ and
$H_2$ at $x=1$ need not be assumed a priori to be the same, and that if
$V$ is continuous, then one only needs $\lambda_{1,n} = \lambda_{2,m(n)}$
for all values of $n$ but one. The same boundary condition for $H_1$
and $H_2$ at $x=0$, however, is crucial for Theorem~1.1 to hold (see
\cite{Ha80}, \cite{De90}).
Moreover, analogs of Theorem \ref{t6.1.1} for certain Schr\"odinger
operators are considered in \cite{Kh84} (see also \cite[Ch.\ 4]{PT87}).
Reconstruction techniques for $V$ in this context are discussed in
\cite{RS92}.
Our purpose in \cite{GS00} was to provide a new approach to Theorem
\ref{t6.1.1} that we felt was more transparent and, moreover, capable of
vast generalizations. To state our generalizations, we will introduce a
shorthand notation to paraphrase Theorem \ref{t6.1.1} by saying ``$V$ on
$[0, 1/2]$ and the eigenvalues of $H$ uniquely determine $V$.'' This
is just a shorthand notation for saying $V_1 =V_2$ a.e.\ if the obvious
conditions hold.
Unless explicitly stated otherwise, all potentials $V, V_1$, and $V_2$
will be real-valued and in $L^1 ([0,1])$ for the remainder of this paper.
Moreover, to avoid too many case distinctions we shall assume $h_0 ,
h_1 \in {\mathbb{R}}$ in \eqref{6.1.1}. In particular, for
$h_0, h_1 \in {\mathbb{R}}$ we index the corresponding eigenvalues $\lambda_n$
of $H$ by $n \in {\mathbb{N}}_0 = {\mathbb{N}} \cup \{0\}$. The case of Dirichlet
boundary conditions, where $h_0 = \infty$ and/or $h_1 = \infty$ has been
dealt with in detail in \cite[Appendix\ A]{GS00}.
Here is a summary of the generalizations proved for
Schr\"odinger operators on $[0,1]$ in \cite{GS00}:
\begin{theorem} \label{t6.1.2} Let $H=-\frac{d^2}{dx^2}+ V$ in $L^2 ([0,1])$
with boundary conditions \eqref{6.1.1} and $h_0,h_1 \in {\mathbb{R}}$. Suppose
$V$ is $C^{2k}(((1/2) - \varepsilon, (1/2) + \varepsilon))$ for some
$k=0,1,\dots$\, and for some $\varepsilon > 0$. Then $V$ on
$[0,1/2]$,
$h_0$, and all the eigenvalues of $H$ except for $(k+1)$ uniquely
determine $h_1$ and $V$ on all of $[0,1]$.
\end{theorem}
\begin{remark} $(i)$. The case $k=0$ in Theorem \ref{t6.1.2} is due to
Hald \cite{Ha80}. \\
$(ii)$ In the non-shorthand form of this theorem (cf.\ the paragraphs preceding Theorem
\ref{t6.1.2}), we mean that both $V_1$
and $V_2$ are $C^{2k}$ near $x=1/2$. \\
$(iii)$ One need not know which eigenvalues are missing. Since the
eigenvalues asymptotically satisfy
$$
\lambda_n = (\pi n)^2 + 2(h_1 - h_0) + \int_0^1 dx\, V(x) +o(1) \,
\text{ as } \, n \to \infty,
$$
given a set of candidates for the spectrum, one can tell how many are
missing. \\
$(iv)$ For the sake of completeness we mention the precise
definition of $H$ in
$L^2 ([0,1])$ for real-valued $V \in L^1([0,1])$ and boundary condition
parameters
$h_0, h_1 \in {\mathbb{R}} \cup \{\infty \}$ in (1.1):
\begin{align}
H=&-\frac{d^2}{dx^2} + V, \notag \\
\text{\rm{dom}} (H)=&\{g \in L^2([0,1]) \, | \, g,g' \in AC([0,1]); \, (-g'' + Vg)
\in L^2([0,1]); \label{6.1.3} \\
& \hspace*{3.1cm} g'(0) + h_0g(0)=0, \, g'(1) + h_1g(1) = 0 \},
\notag
\end{align}
where $AC([0,1])$ denotes the set of absolutely continuous functions on
$[0,1]$
and $h_{x_0}=\infty$ represents the Dirichlet boundary condition
$g(x_0)=0$ for $x_0 \in \{0,1\}$ in \eqref{6.1.3}.
\end{remark}
By means of explicit examples, it has been shown in Section\ 3 of
\cite{GS00}, that Theorem \ref{t6.1.2} is optimal in the sense that if $V$
is only assumed to be $C^{2k-1}$ near
$x=1/2$ for some $k \geq 1$, then it is not uniquely determined by
$V\restriction [0,1/2]$ and all the
eigenvalues but $(k+1)$.
Theorem \ref{t6.1.2} works because the condition that $V$ is $C^{2k}$ near
$x=1/2$ gives us partial information about $V$ on $[1/2, 1]$; indeed, we
know the values of
$$
V(1/2), V'(1/2), \dots, V^{(2k)}(1/2)
$$
computed on $[1/2, 1]$ since one can compute them on $[0,1/2]$. This
suggests that knowing $V$ on more than $[0,1/2]$ should let one dispense
with a finite density of eigenvalues. That this is indeed the case is the
content of the following theorem. (We denote by $\#\{\cdots\}$ the cardinality of
the set $\{\cdots\}$.)
\begin{theorem} \label{t6.1.3} Let $H = -\frac{d^2}{dx^2}+V$ in $L^2 ([0,1])$
with boundary conditions \eqref{6.1.1} and $h_0,h_1 \in {\mathbb{R}}$. Then $V$
on $[0,(1+\alpha)/2]$ for some $\alpha \in (0,1)$, $h_0$, and a
subset $S \subseteq \sigma(H)$ of all the eigenvalues $\sigma(H)$ of $H$
satisfying
\begin{equation}
\#\{\lambda\in S \,|\, \lambda\leq \lambda_0 \} \geq (1-\alpha)
\# \{\lambda\in \sigma(H) \,|\, \lambda\leq \lambda_0 \} + (\alpha/2)
\label{6.1.4}
\end{equation}
for all sufficiently large $\lambda_0 \in {\mathbb{R}}$, uniquely determine
$h_1$ and $V$ on all of $[0,1]$.
\end{theorem}
\begin{remark} $(i)$ As a typical example, knowing slightly more than half
the eigenvalues and knowing $V$ on $[0,\frac34]$ determines $V$ uniquely
on all of $[0,1]$. To the best of our knowledge, Theorem \ref{t6.1.3}
introduced and solved a new type of inverse
spectral problem. \\
$(ii)$ As in the case $\alpha =0$, one has an extension of the same type
as Theorem \ref{t6.1.2}. Explicitly, if $V$ is assumed to be $C^{2k}$ near
$x=(1 +\alpha)/2$, we only need
$$
\# \{\lambda\in S \,|\, \lambda\leq \lambda_0 \} \geq (1-\alpha)
\# \{\lambda\in \sigma(H) \,|\, \lambda\leq \lambda_0 \} + (\alpha/2) -
(k+1)
$$
instead of \eqref{6.1.4}.
\end{remark}
One can also derive results about problems on all of ${\mathbb{R}}$.
\begin{theorem} \label{t6.1.4} Suppose that $V\in L^1_{\text{\rm{loc}}}
({\mathbb{R}})$ satisfies the following two conditions: \\
$(i)$ $V(x) \geq C|x|^{2+\varepsilon} -D$ for some
$C,\varepsilon, D>0$. \\
$(ii)$ $V(-x) \geq V(x) \quad x\geq 0$. \\
Then $V$ on $[0,\infty)$ and the spectrum of $H = -\frac{d^2}{dx^2} + V$
in $L^2 ({\mathbb{R}})$ uniquely determine $V$ on all of ${\mathbb{R}}$.
\end{theorem}
Hochstadt-Lieberman \cite{HL78} used the details of the inverse
spectral theory in their proof. In a sense, we only used in \cite{GS00}
the main uniqueness theorem of that theory due to Marchenko
\cite{Ma73}, which we now describe. For $V \in L^1([a,b])$
real-valued, $-\infty <a <b <\infty$, consider $-u'' + Vu = zu$
with the boundary condition
\begin{equation}
u'(b) + h_b u(b) = 0 \label{6.1.7}
\end{equation}
at $x=b$. Let $u_+ (z,x)$ denote the solution of this equation, normalized,
say, by
$u_+ (z,b) = 1$. The $m_+$-function is then defined by
\begin{equation}
m_+ (z,a) = \frac{u'_+ (z,a)}{u_+ (z,a)}\, . \label{6.1.8}
\end{equation}
Similarly, given a boundary condition at $x=a$,
\begin{equation}
u' (a) + h_a u(a) = 0, \label{6.1.9}
\end{equation}
we define the solution $u_- (z,x)$ of $-u'' + Vu = zu$ normalized by
$u_- (z,a)=1$ and then define
\begin{equation}
m_- (z,b) = \frac{u'_- (z,b)}{u_- (z,b)}\, . \label{6.1.10}
\end{equation}
In our present context where $-\infty < a < b < \infty$, $m_{\pm}$ are
even meromorphic on ${\mathbb{C}}$). Moreover,
$$
\text{Im}\, (z) > 0 \, \text{ implies } \, \Im (m_- (z,b)) < 0,
\, \Im (m_+ (z,a)) > 0.
$$
Marchenko's \cite{Ma73} fundamental uniqueness theorem of inverse
spectral theory then reads as follows:
\begin{theorem} \label{t6.1.5} $m_+ (z,a)$ uniquely determines $h_b$ as well
as $V$ a.e.\ on $[a,b]$.
\end{theorem}
If $V \in L^1_{\text{\rm{loc}}}([a,\infty))$ is real-valued (with
$|a|<\infty$) and
$-\frac{d^2}{dx^2} + V$ is in the limit point case at infinity, one can
still define a unique
$m_+ (z,a)$ function but now for $Im (z) \neq 0$ rather than all
$z \in {\mathbb{C}}$. For such $z$, there is a unique function $u_+(z,
\,\cdot\,)$ which is
$L^2$ at infinity (unique up to an overall scale factor which drops out of
$m_+ (z,a)$ defined by \eqref{6.1.8}). Again, one has the following
uniqueness result independently proved by Borg \cite{Bo52} and
Marchenko \cite{Ma73}.
\begin{theorem} \label{t6.1.6} $m_+ (z,a)$ uniquely determines $V$
a.e.\ on $[a,\infty)$.
\end{theorem}
It is useful to have $m_-(z,b)$ because of the following basic fact:
\begin{theorem} \label{t6.1.7} Let $H = -\frac{d^2}{dx^2} + V$ be a
Schr\"odinger operator in $L^2 ([a,b])$ with boundary conditions
\eqref{6.1.7}, \eqref{6.1.9} and let $G(z,x,y)$ be the integral kernel
of $(H-z)^{-1}$. Suppose $c \in (a,b)$ and let $m_+ (z,c)$ be the
corresponding $m_+$-function for $[c,b]$ and $m_-(z,c)$ the
$m_-$-function for $[a,c]$. Then
\begin{equation}
G(z,c,c) = \frac{1}{m_- (z,c) - m_+ (z,c)}\, . \label{6.1.12}
\end{equation}
\end{theorem}
Theorems \ref{t6.1.5} and \ref{t6.1.6} are deep facts; Theorem
\ref{t6.1.7} is an elementary calculation following from the explicit
formula for the integral kernel of $(H-z)^{-1}$,
$$
G(z,x,y) = \frac{u_- (z,\min (x,y)) u_+ (z, \max (x,y))}{W (u_- (z), u_+
(z))(x)},
$$
where as usual
$
W(f,g)(x) = f' (x) g(x) - f(x)g'(x)
$
denotes the Wronskian of $f$ and $g$. An analog of Theorem \ref{t6.1.7}
holds in case $[a,b]$ is replaced by $(-\infty,\infty)$.
We can now describe the strategy of our proofs of Theorems
\ref{t6.1.1}--\ref{t6.1.4}. $G(z,c,c)$ has poles at the eigenvalues of $H$
(this is not quite true; see below), so by \eqref{6.1.12}, at eigenvalues
$\lambda_n$ of $H$:
\begin{equation}
m_+ (\lambda_n,c) = m_- (\lambda_n,c). \label{6.1.13}
\end{equation}
If we know $V$ on a left partial interval $[a,c]$ and we know some
eigenvalue $\lambda_n$, then we know $m_- (z,c)$ exactly; so by
\eqref{6.1.13}, we know the value of $m_+ (\lambda_n,c)$ at the point
$\lambda_n$. Below we indicate when knowing the values of
$f(\lambda_n)$ of an analytic function of the type of the
$m$-functions uniquely determines $f(z)$. If $m_+(z,c)$ is determined,
then by Theorem \ref{t6.1.5}, $V$ is determined on $[a,b]$ and so is
$h_b$.
So the logic of the argument for a theorem like Theorem \ref{t6.1.1} is
the following: \\
$(i)$ $V$ on $[0,1/2]$ and $h_0$ determine $m_- (z,1/2)$
by direct spectral theory. \\
$(ii)$ The $\lambda_n$ and \eqref{6.1.13} determine $m_+ (\lambda_n,1/2)$, and
then by suitable theorems in complex analysis, $m_+ (z, 1/2)$ is uniquely
determined for all $z$. \\
$(iii)$ $m_+ (z,1/2)$ uniquely determines $V$ (a.e.) on
$[1/2,1]$ and $h_1$ by inverse spectral theory.
It is clear from this approach why $h_0$ is required and $h_1$ is free in
the context of Theorem \ref{t6.1.1} (see \cite{De90} for examples where
$h_1$ and $V \restriction [0,1/2]$ do not determine $V$); without
$h_0$ we cannot compute $m_-(z, 1/2)$ and so start the process.
As indicated before \eqref{6.1.13}, $G(z,c,c)$ may not have a pole at an
eigenvalue $\lambda_n$ of $H$. It will if $u_n (c) \neq 0$, but if $u_n
(c) =0$, then $G(z,c,c) = 0$ rather than $\infty$. Here $u_n$ denotes the
eigenfunction of $H$ associated with the (necessarily simple) eigenvalue
$\lambda_n$. Nevertheless, \eqref{6.1.13} holds at points where $u_n (c)
=0$ since then $u_- (c)=u_+ (c) = 0$, and so both sides of \eqref{6.1.13}
are infinite. (In spite of \eqref{6.1.13}, $m_- - m_+$ is also infinite at
$z = \lambda_n$ and so $G(\lambda_n, c, c) = 0$.) We summarize this
discussion next:
\begin{theorem} \label{t6.1.8} For any $c\in (a,b)$, \eqref{6.1.13} holds at
any eigenvalue $\lambda_n$ of $H_{[a,b]}$ $($with the possibility of
both sides of \eqref{6.1.13} being infinite$)$.
\end{theorem}
\medskip
{\it More recent references:} A new inverse nodal problem was reduced to Theorem \ref{t6.1.3} by Yang \cite{Ya01}. A substantial generalization of Theorem \ref{t6.1.4}, replacing condition $(i)$ by $H$ being bounded from below with purely discrete spectrum, was proved by Khodakovsky \cite{Kh99}, \cite{Kh00}. He also found other variants of Theorem \ref{t6.1.4}.
\smallskip
\begin{center}
$\bigstar$ \qquad $\bigstar$ \qquad $\bigstar$
\end{center}
\smallskip
We end our survey of \cite{GS00} by briefly indicating the
uniqueness theorems for entire functions needed in the proofs of Theorems
\ref{t6.1.1}--\ref{t6.1.4}. In discussing extensions of Hochstadt's
discrete (finite matrix) version \cite{Ho79} of the
Hochstadt--Lieberman theorem in \cite{GS97a}, we made use of the
following simple lemma which is an elementary consequence of the fact
that any polynomial of degree $d$ with $d+1$ zeros must be the zero
polynomial:
\begin{lemma} \label{t6.B.1} Suppose $f_1 = \frac{P_1}{Q_1}$ and $f_2 =
\frac{P_2}{Q_2}$ are two rational fractions where the polynomials satisfy
$\deg (P_1) = \deg (P_2)$ and $\deg (Q_1)=\deg (Q_2)$. Suppose that $d =
\deg (P_1) + \deg (Q_1)$ and that $f_1 (z_n) = f_2 (z_n)$ for $d+1$ distinct
points $\{z_n\}^{d+1}_{n=1}\in {\mathbb{C}}$. Then $f_1 = f_2$.
\end{lemma}
In the context of \cite{GS00}, one is interested in entire functions of the
form
\begin{equation}
f(z) = C \prod^\infty_{n=0} \biggl( 1- \frac{z}{x_n}\biggr), \label{6.B.1}
\end{equation}
where $0<x_0 < x_1 < \cdots$ is a suitable sequence of positive numbers which
are the zeros of $f$ and $C$ is some complex constant.
Given a sequence $\{x_n\}_{n=0}^{\infty}$ of positive reals, we define
$$
N(t) = \#\{n \in {\mathbb{N}} \cup \{0\}\,|\, x_n < t\}.
$$
We recall the following basic theorem (see, e.g.,
\cite[Ch.\ I]{Le80}, \cite[Sects.\ II.48, II.49]{Ma85}):
\begin{theorem} \label{t6.B.2} Fix $0 < \rho_0 < 1$. Then: \\
$(i)$ If $\{x_n\}^\infty_{n=0}$ is a sequence of positive reals with
\begin{equation}
\sum_{n=0}^{\infty} x^{-\rho}_n < \infty \, \text{ for all
$\rho > \rho_0$,}
\label{6.B.3}
\end{equation}
then the product in \eqref{6.B.1} defines an entire function $f$ with
\begin{equation}
|f(z)| \leq C_1 \exp (C_2 |z|^\rho ) \, \text{ for all $\rho > \rho_0$}.
\label{6.B.4}
\end{equation}
$(ii)$ Conversely, if $f$ is an entire function satisfying \eqref{6.B.4}
with all its $($complex$)$ zeros on $(0, \infty)$, then its zeros
$\{x_n \}^\infty_{n=0}$ satisfy \eqref{6.B.3}, and $f$ has the canonical
product expansion \eqref{6.B.1}. \\
Moreover, \eqref{6.B.3} holds if and only if
\begin{equation}
N(t) \leq C |t|^\rho \, \text{ for all $\rho > \rho_0$}. \label{6.B.5}
\end{equation}
\end{theorem}
Given this theorem, we single out the following definition.
\begin{definition} \label{d6.B.3} A function $f$ is called of $m$-type if and
only if $f$ is an entire function satisfying \eqref{6.B.4} (of order $0<\rho<1$
in the usual definition) with all the zeros of $f$ on $(0, \infty)$.
\end{definition}
Our choice of ``$m$-type'' in Definition \ref{d6.B.3} comes from the fact
that in many cases we discuss in this paper, the $m$-function is a ratio
of functions of $m$-type. By Theorem \ref{t6.B.2}, $f$ in Definition
\ref{d6.B.3} has the form \eqref{6.B.1} and $N(t)$, which we will denote
as $N_f (t)$, satisfies \eqref{6.B.5}.
\begin{lemma} \label{l6.B.5} Let $f$ be a function of $m$-type. Then there
exists
a $0 < \rho < 1$ and a sequence $\{R_k\}_{k=1}^{\infty}$, $R_k \to \infty$ as
$k \to \infty$, so that
$$
\inf \{ |f(z)| \,|\, |z| = R_k\} \geq C_1 \exp (-C_2 R^{\rho}_k).
$$
\end{lemma}
\begin{lemma} \label{l6.B.6} Let $F$ be an entire function that satisfies
the following two conditions: \\
$(i)$ $\sup_{|z|=R_k} |F(z)| \leq C_1 \exp (C_2 R^\rho_k)$ for
some $0 \leq \rho <1$, $C_1, C_2 >0$, and some sequence $R_k \to \infty$ as
$k \to \infty$. \\
$(ii)$ $\lim_{|x|\to\infty; x\in{\mathbb{R}}} |F(ix)| =0$. \\
Then $F \equiv 0$.
\end{lemma}
Lemmas \ref{l6.B.5} and \ref{l6.B.6} finally yield the following result.
\begin{theorem} \label{t6.B.4} Let $f_1 ,f_2 ,g$ be three functions of
$m$-type so that the following two conditions hold: \\
$(i)$ $f_1 (z) = f_2 (z)$ at any point $z$ with $g(z) =0$.
$(ii)$ For all sufficiently large $t$,
$$
\max (N_{f_1} (t), N_{f_2} (t)) \leq N_g (t) -1.
$$
Then, $f_1 =f_2$.
\end{theorem}
\smallskip
\begin{center}
$\bigstar$ \qquad $\bigstar$ \qquad $\bigstar$
\end{center}
\smallskip
Refinements of the results of \cite{GS00} can be found in
\cite{DGS97}, \cite{DGS99}. Here we just mention the following facts.
\begin{theorem} \label{6.3.1a} Let $H_1 (h_0), H_2(h_0)$ be associated with
two potentials $V_1, V_2$ on $[0,1]$ and two potentially distinct
boundary conditions $h^{(1)}_1, h^{(2)}_1 \in {\mathbb{R}}$ at $x=1$.
Suppose that $\{(\lambda_n, h^{(n)}_0)\}_{n\in{\mathbb{N}}_0}$ is a sequence
of pairs with $\lambda_0 <\lambda_1 <\cdots \to \infty$ and $h^{(n)}_0
\in{\mathbb{R}} \cup\{\infty\}$ so that both $H_1 (h^{(n)}_0)$ and
$H_2 (h^{(n)}_0)$ have eigenvalues at $\lambda_n$. Suppose that
$$
\sum^\infty_{n=0} \frac{(\lambda_n -
\frac{1}{4} \pi^2 n^2)_+}{n^2}
<\infty
$$
holds. Then $V_1 = V_2$ a.e.\ on $[0,1]$ and $h^{(1)}_1 = h^{(2)}_1$.
\end{theorem}
This implies Borg's celebrated two-spectra uniqueness result \cite{Bo46}
(see also, \cite{Le68}, \cite{LG64}, \cite[Ch.\ 3]{Le87}, \cite{Ma73}):
\begin{corollary} \label{c6.3.2a} Fix $h^{(1)}_0, h^{(2)}_0 \in {\mathbb{R}}$.
Then all the eigenvalues of $H(h^{(1)}_0)$ and all the
eigenvalues of $H(h^{(2)}_0)$, save one, uniquely determine
$V$ a.e.\ on $[0,1]$.
\end{corollary}
It also implies the following amusing result:
\begin{corollary} \label{c6.3.3a} Let $h^{(1)}_0, h^{(2)}_0, h^{(3)}_0
\in {\mathbb{R}}$ and denote by $\sigma_j=\sigma(H(h^{(j)}_0))$ the spectra
of $H(h^{(j)}_0)$, $j=1,2,3$. Assume $S_j \subseteq \sigma_j$, $j=1,2,3$
and suppose that for all sufficiently large $\lambda_0 >0$ one has
$$
\#\{\lambda \in \{S_1 \cup S_2 \cup S_3\} \, \text{ with } \,
\lambda\leq \lambda_0 \} \geq \tfrac23 \#\{\lambda \in
\{\sigma_1 \cup \sigma_2 \cup \sigma_3 \} \,
\text{ with } \, \lambda\leq \lambda_0 \} -1.
$$
Then $V$ is uniquely determined a.e.\ on $[0,1]$.
\end{corollary}
In particular, two-thirds of three spectra determine $V$.
\medskip
{\it More recent references:} Further refinements of Corollary \ref{c6.3.3a}, involving $N$ spectra, were proved by
Horv\'ath \cite{Ho01} (he also studies the corresponding analog for a Dirac-type operator). Optimal and nearly optimal conditions for a set of eigenvalues to determine the potential in terms of closedness properties of the exponential system corresponding to the known eigenvalues (implying Theorem \ref{6.3.1a} and a generalization thereof) were also derived by Horv\'ath \cite{Ho05}. For an interesting half-line problem related to this circle of ideas we also refer to Horv\'ath \cite{Ho06}. A variant of Theorem
\ref{6.3.1a} was discussed by Ramm \cite{Ra99}, \cite{Ra00}.
Hochstadt--Lieberman-type problems for Schr\"odinger operators including a reconstruction algorithm has been presented by L.\ Sakhnovich.
The analog of the two-spectra result, Corollary \ref{c6.3.2a}, including a reconstruction algorithm, for a class of singular potentials has been discussed by Hryniv and Mykytyuk \cite{HM03a}, \cite{HM04} (see also \cite{HM05}). They also studied Hochstadt--Lieberman-type results for such a class of singular potentials in \cite{HM04a}.
Hochstadt--Lieberman-type results for a class of Dirac-type operators relevant to the AKNS system were published by del Rio and Gr\'ebert \cite{DG01}. Borg- and Hochstadt--Lieberman-type inverse problems for systems including matrix-valued
Schr\"odinger and Dirac-type equations, were studied in depth by M.\ Malamud
\cite{Ma99}, \cite{Ma99a}, \cite{Ma05} \cite{Ma06}. He also studied Borg-type theorems for $n$th-order scalar equations \cite{Ma06a}. Borg- and Hochstadt--Lieberman-type inverse problems for matrix-valued Schr\"odinger operators were also studied by Shen \cite{Sh01}. He also considered Borg-type inverse problems for Schr\"odinger operators with weights \cite{Sh05}.
Additional results on determining the potential uniquely from spectra associated to
three intervals of the type $[0,1]$, $[0,a]$, and $[a,1]$ for some $a\in(0,1)$ (and similarly for whole-line problems with purely discrete spectra) can be found in \cite{GS99}. This has been inspired by work of Pivovrachik \cite{Pi99}, who also addressed the reconstruction algorithm from three spectra in the symmetric case $a=1/2$ (see also
\cite{Pi99b}, \cite{Pi01}, \cite{Pi05}). He also considered the analogous Sturm--Liouville problem applicable to a smooth inhomogeneous partially damped string in \cite{Pi99a} and extended some of these results to Sturm--Liouville equations on graphs in \cite{Pi00}, \cite{Pi06}. Uniqueness and characterization problems for a class of singular Sturm--Liouville problems associated with three spectra were studied by Hryniv and Mykytyuk \cite{HM03}. The reconstruction of a finite Jacobi matrix from three of its spectra was presented by Michor and Teschl \cite{MT04}.
\smallskip
\begin{center}
$\bigstar$ \qquad $\bigstar$ \qquad $\bigstar$
\end{center}
\smallskip
These results are related to two other papers: In \cite{GS97a}, we
considered, among other topics, analogs of Theorems \ref{t6.1.1} and
\ref{t6.1.3} for finite tri-diagonal (Jacobi) matrices, extending a result
in \cite{Ho79}. The approach there is very similar
to the current one except that the somewhat subtle theorems on zeros of
entire functions in this paper are replaced by the elementary fact that a
polynomial of degree at most $N$ with $N+1$ zeros must be identically
zero. In \cite{GS97}, we consider results related to Theorem \ref{t6.1.4}
in that for Schr\"odinger operators on
$(-\infty, \infty)$, ``spectral'' information plus the potential on one of
the half-lines determine the potential on all of $(-\infty,\infty)$. In
that paper, we considered situations where there are scattering states for
some set of energies and the ``spectral'' data are given by a reflection
coefficient on a set of positive Lebesgue
measure in the a.c.\ spectrum of $H$. The approach is not as close to this
paper as is \cite{GS97a}, but $m$-function techniques (see also
\cite{GS96a}) are critical in all three papers.
\medskip
{\it More recent references:} For additional results on inverse scattering with partial information on the potential we refer to Aktosun and Papanicolaou \cite{AP03}, Aktosun and Sacks \cite{AS98}, Aktosun and Weder \cite{AW02}, and the references therein.
\smallskip
\begin{center}
$\bigstar$ \qquad $\bigstar$ \qquad $\bigstar$
\end{center}
\smallskip
We conclude this section by briefly describing some of the results in
\cite{GS97a}, where inverse spectral analysis for finite and
semi-infinite Jacobi operators $H$ was studied. While discussing a variety of topics
(including trace formulas), we also provided a new proof of a result of
Hochstadt \cite{Ho79} and its extension, which can be viewed as the
discrete analog of the Hochstadt and Lieberman result in \cite{HL78}.
Moreover, we solved the inverse spectral problem for $(\delta_n,
(H-z)^{-1}\delta_n)$ in the case of finite Jacobi matrices. As mentioned
earlier, the tools we apply are grounded in $m$-function techniques.
Explicitly, \cite{GS97a} studied finite $N\times N$ matrices of the form:
$$
H=\begin{pmatrix}
b_1 & a_1 & 0 & 0 & \cdot &
\cdot & \cdot \\
a_1 & b_2 & a_2 & 0 & \cdot &
\cdot & \cdot \\
0 & a_2 & b_3 & a_3 & \cdot &
\cdot & \cdot \\
\cdot & \cdot & \cdot & \cdot &
\cdot & \cdot & \cdot \\
\cdot & \cdot & \cdot & \cdot &
\cdot & \cdot & \cdot \\
\cdot & \cdot & \cdot & \cdot &
0 & a_{N-1} & b_N
\end{pmatrix}
$$
and the semi-infinite analog $H$ defined on
$$
\ell^2 ({\mathbb{N}}) = \bigg\{ u = (u(1), u(2), \dots) \,\bigg|\,
\sum^\infty_{n=1} |u(n)|^2 < \infty \bigg\}
$$
given by
$$
(Hu)(n) = \begin{cases} a_n u(n+1) + b_n u(n) + a_{n-1}
u(n-1), & n \geq 2, \\
a_1 u(2) + b_1 u(1), & n=1. \end{cases}
$$
In both cases, we assume $a_n, b_n \in{\mathbb{R}}$
with $a_n >0$. To avoid inessential technical complications,
we will only consider the case where $\sup_n [|a_n| + |b_n|] <\infty$ in
which case $H$ is a map from $\ell^2$ to $\ell^2$ and defines a
bounded self-adjoint operator. In the semi-infinite case, we will set
$N=\infty$. It will also be useful to consider
the $b$'s and $a$'s as a single sequence
$b_1, a_1, b_2, a_2, \dots = c_1, c_2, \dots$, that is,
$$
c_{2n-1} = b_n, \quad c_{2n} = a_n, \quad n\in{\mathbb{N}}.
$$
Concerning the recovery of a finite Jacobi matrix from parts of the matrix
and additional spectral information (i.e., mixed data), Hochstadt
\cite{Ho79} proved the following remarkable theorem.
\begin{theorem} \label{t6.4.1a} Let $N \in {\mathbb{N}}$. Suppose that
$c_{N+1}, \dots, c_{2N-1}$ are known, as well as the eigenvalues
$\lambda_1, \dots,\lambda_N$ of $H$. Then $c_1, \dots, c_N$ are uniquely
determined.
\end{theorem}
The discrete Hochstadt--Lieberman-type theorem proved in \cite{GS97a}
reads as follows.
\begin{theorem} \label{t6.4.2a} Suppose that
$1 \leq j \leq N$ and $c_{j+1}, \dots, c_{2N-1}$ are known, as well as $j$
of the eigenvalues. Then $c_1, \dots, c_j$ are uniquely determined.
\end{theorem}
We emphasize that one need {\it{not}} know which of the $j$ eigenvalues
one has.
Borg \cite{Bo46} proved the celebrated theorem that the spectra for two
boundary conditions of a bounded interval regular Schr\"odinger operator
uniquely determine the potential. Later refinements (see, e.g.,
\cite{Bo52}, \cite{Ho67a}, \cite{Le49}, \cite{Le68}, \cite{LG64}, \cite{Ma73})
imply that they even determine the two boundary conditions.
Next, we consider analogs of this result for a finite Jacobi
matrix. Such analogs were first considered by Hochstadt
\cite{Ho67}, \cite{Ho74} (see also \cite{BG78}, \cite{FH74}, \cite{GH84},
\cite{GW76}, \cite{Ha76}, \cite{Ho79}). The results below are
adaptations of known results for the continuum or the semi-infinite case,
but the ability to determine parameters by counting sheds light on facts
like the one that the lowest eigenvalue in the Borg result is not needed
under certain circumstances.
Given $H$, an $N\times N$ Jacobi matrix, one defines $H(b)$ to be the
Jacobi matrix where all $a$'s and $b$'s are the same as $H$, except $b_1$
is replaced by $b_1 + b$, that is,
$$
H(b) = H + b(\delta_1, \, \cdot\,
)\delta_1.
$$
\begin{theorem} \label{t6.5.1a} The eigenvalues $\lambda_1, \dots,
\lambda_N$ of $H$, together with $b$ and $N-1$ eigenvalues $\lambda
(b)_1, \dots, \lambda (b)_{N-1}$ of $H(b)$, determine $H$ uniquely.
\end{theorem}
Again it is irrelevant which $N-1$ eigenvalues of the $N$ eigenvalues of
$H(b)$ are known.
\begin{theorem} \label{t6.5.2a} The eigenvalues
$\lambda_1, \dots, \lambda_N$ of $H$, together with the $N$ eigenvalues
$\lambda (b)_1, \dots, \lambda (b)_N$ of some $H(b)$
$($with $b$ unknown$)$, determine $H$ and $b$.
\end{theorem}
\begin{remark} Since
$$
b= \text{Tr}(H(b)-H) = \sum^N_{j=1}
(\lambda (b)_j - \lambda_j),
$$
we can a priori deduce $b$ from the $\lambda (b)$'s and $\lambda$'s and
so deduce Theorem \ref{t6.5.2a} from Theorem \ref{t6.5.1a}. We note that
the parameter counting works out. In Theorem \ref{t6.5.1a}, $2n-1$
eigenvalues determine $2n-1$ parameters; and in Theorem \ref{t6.5.2a},
$2n$ eigenvalues determine $2n$ parameters.
\end{remark}
The basic inverse spectral theorem for finite Jacobi matrices shows that
$(\delta_1, (H-z)^{-1} \delta_1)$ determines $H$ uniquely. In
\cite{GS97a} we considered $N \in {\mathbb{N}}$, $1\leq n \leq N$, and asked
whether $(\delta_n, (H-z)^{-1} \delta_n)$ determines $H$ uniquely. For
notational convenience, we occasionally allude to $G(z,n,n)$ as the $n,n$
Green's function in the remainder of this section. The $n=1$ result can
be summarized via:
\begin{theorem} \label{6.6.1a} $(\delta_1, (H-z)^{-1}\delta_1)$ has the form
$\sum^N_{j=1}\alpha_j (\lambda_j -z)^{-1}$ with $\lambda_1 <\cdots <
\lambda_N, \, \sum^N_{j=1} \alpha_j =1$ and each $\alpha_j >0$. Every
such sum arises as the $1,1$ Green's function of an $H$ and of exactly
one such $H$.
\end{theorem}
For general $n$, define $\tilde n = \min (n, N+1 -n)$. Then the following
theorems were proved in \cite{GS97a}:
\begin{theorem} \label{t6.6.2a} $(\delta_n, (H-z)^{-1}\delta_n)$ has the form
$\sum^k_{j=1} \alpha_j (\lambda_j -z)^{-1}$ with $k$ one of $N, N-1,
\dots, N-\tilde n + 1$ and $\lambda_1< \cdots < \lambda_k, \,\sum^k_{j=1}
\alpha_j =1$ and each $\alpha_j >0$. Every such sum arises
as the $n,n$ Green's function of at least one $H$.
\end{theorem}
\begin{theorem} \label{t6.6.3a} If $k=N$, then precisely $\binom{N-1}{n-1}$
operators $H$ yield the given $n,n$ Green's function.
\end{theorem}
\begin{theorem} \label{t6.6.4a} If $k<N$, then infinitely many Jacobi
matrices $H$ yield the given $n,n$ Green's function. Indeed, the inverse
spectral family is then a collection of
$\binom{k-1}{N-k} \binom{k-1-N+k}{n-1-N+k}$ disjoint manifolds, each of
dimension $N-k$ and diffeomorphic to an $(N-k)$-dimensional open ball.
\end{theorem}
\medskip
{\it More recent references:} Additional geometric information in connection with Theorem \ref{t6.6.4a} and a version for off-diagonal Green's functions were studied by Gibson \cite{Gi02}.
Borg- and discrete Hochstadt--Lieberman-type results for generalized (i.e., certain
tri-diagonal block) Jacobi matrices were studied by Derevyagin \cite{De06} (see also
Shieh \cite{Sh04}). The case of non-self-adjoint Jacobi matrices with a rank-one imaginary part, and an extension of Hochstadt--Lieberman-type results to this situation was recently discussed by Arlinski\u i and Tsekanovski\u i \cite{AT06}. An extension of results of Hochstadt \cite{Ho74} to the case of normal matrices was found by
S.\ Malamud \cite{Ma04}. A detailed treatment of two-spectra inverse problems of semi-infinite Jacobi operators, including reconstruction, has recently been presented by Silva and Weder \cite{SW06}.
\section{The Crown Jewel: Simon's New Approach to Inverse Spectral
Theory}
\label{s6}
In this section we summarize some of the principal results of
the following papers: \\
\cite{Si99} B.~Simon, {\it A new approach to inverse spectral theory, I.
Fundamental formalism}, Ann. Math. {\bf 150}, 1029--1057 (1999). \\
\cite{GS00a} F.\ Gesztesy and B.\ Simon, {\it A new approach to
inverse spectral theory, II. General real potentials and the connection to
the spectral measure}, Ann. Math. {\bf 152}, 593--643 (2000). \\
\cite{RS00} A.\ Ramm and B.\ Simon, {\it A new approach to inverse spectral
theory, III. Short range potentials}, J. Analyse Math. {\bf 80},
319--334 (2000). \\
\cite{GS00b} F.~Gesztesy and B.~Simon, {\it On local Borg--Marchenko
uniqueness results}, Commun. Math. Phys. {\bf 211}, 273--287 (2000).
\medskip\smallskip
As the heading of this section suggests, we are approaching the pinnacle of
Barry Simon's contributions to inverse scattering theory thus far: In his
spectacular paper \cite{Si99}, he single-handedly developed a new approach
to inverse spectral theory for Schr\"odinger operators on a half-line, by
starting from a particular representation of the Weyl--Titchmarsh
$m$-function as a finite Laplace-type transform with control over the error term. In addition to establishing this feat, it also led to a completely unexpected
uniqueness result for Weyl--Titchmarsh functions, what is now called the local
Borg--Marchenko uniqueness theorem, but which really should have been
named Simon's local uniqueness theorem. The inverse spectral approach for
Schr\"odinger operators on a half-line (including a reconstruction algorithm for the potential) originated with the celebrated
paper \cite{GL55} by Gelfand and Levitan in 1951 and an independent approach by Krein \cite{Kr51} in the same year, followed by a seminal
contribution \cite{Ma73} by Marchenko in 1952. The
Borg--Marchenko uniqueness result was first published by Marchenko \cite{Ma50} in 1950 but Borg apparently had it in 1949 and it was independently published by Borg
\cite {Bo52} and again by Marchenko \cite{Ma73} in 1952.
Both results, the uniqueness theorem and the Gel'fand--Levitan (reconstruction) formalism remained pillars of the inverse spectral theory that withstood any reformulation or improvement for nearly fifty years. Hence it was an incredible
achievement by Barry Simon to have changed the inverse spectral landscape
by offering such a reformulation of inverse spectral theory and in the
very same paper \cite{Si99} to have been able to substantially improve the
Borg--Marchenko uniqueness theorem from a global to a local version.
We start by highlighting the approach in Simon's paper \cite{Si99} and then switch
to a more detailed treatment of some aspects of the theory by borrowing from
\cite{GS00a}.
Inverse spectral methods have been actively studied in the past
years both via their relevance in a variety of applications and due to
their connection with integrable evolution equations such as the KdV equation. In this section, however, we will not deal with the full-line inverse spectral approach relevant to integrable equations but exclusively focus on inverse spectral theory for half-line
Schr\"odinger operators. In this particular context, a major role is played by
the Gel'fand--Levitan equations \cite{GL55} (see also,
\cite[Chs.\ 3, 4]{CCPR97}, \cite{CS89}, \cite{Kr53}, \cite{Kr53a}, \cite{Kr54},
\cite[Ch.\ 2]{Le87}, \cite{Ma73}, \cite[Ch.\ 2]{Ma86}, \cite[Ch.\ VIII]{Na68}, \cite{Re02},
\cite{Sy79}, \cite{Th79}). The goal in Barry Simon's paper \cite{Si99} was
to present a new approach to their basic results. In particular, he
introduced a new basic object, the $A$-function (see \eqref{7.1.24}
below), the remarkable equation \eqref{7.1.28} it satisfies, and
illustrated its fundamental importance with several new results including
improved asymptotic expansions of the Weyl--Titchmarsh $m$-function in
the high-energy regime and the local uniqueness result.
To present some of these new results, we will first describe the
major players in this game. One is concerned with self-adjoint differential
operators on either $L^2 ([0,b])$ with $b<\infty$, or $L^2 ([0,\infty))$
associated with differential expressions of the form
\begin{equation}
-\frac{d^2}{dx^2} + V(x), \quad x\in (0,b). \label{7.1.1}
\end{equation}
If $b$ is finite, we suppose
$$
\int_0^b dx \, |V(x)| < \infty
$$
and place a boundary condition
\begin{equation}
u'(b) + hu(b) =0 \label{7.1.3}
\end{equation}
at $b$, where $h\in{\mathbb{R}} \cup \{\infty\}$ with $h=\infty$ shorthand
for the Dirichlet boundary condition $u(b)=0$. If $b=\infty$, we suppose
$$
\int_y^{y+1} dx \, |V(x)| < \infty \, \text{ for all } \, y\geq 0
$$
and
\begin{equation}
\sup_{y>0} \int_y^{y+1} dx \, \max(V(x), 0) <
\infty . \label{7.1.5}
\end{equation}
Under condition \eqref{7.1.5}, it is known that \eqref{7.1.1} is limit
point at infinity \cite[App.\ to Sect.\ X.1]{RS75}. In addition, a fixed self-adjoint
boundary condition at $x=0$ is assumed when talking about the self-adjoint operator associated with \eqref{7.1.1}.
In either case, for each $z\in{\mathbb{C}} \backslash [\beta, \infty)$
with $-\beta$ sufficiently large, there is a unique solution (up
to an overall constant), $u(z,x)$, of $-u''+Vu =zu$ which satisfies
\eqref{7.1.3} at $b$ if $b<\infty$ or which is $L^2$ at $\infty$ if
$b=\infty$. The principal $m$-function $m(z)$ is defined by
$$
m(z) =\frac{u'(z,0)}{u(z,0)}.
$$
If we replace $b$ by $b_1 = b-x_0$ with $x_0\in (0,b)$ and let
$V(s) = V(x_0+s)$ for $s\in (0,b_1)$, we get a new $m$-function
we will denote by $m(z,x_0)$. It is given by
$$
m(z,x) = \frac{u'(z,x)}{u(z,x)}.
$$
$m(z,x)$ satisfies the Riccati-type equation
\begin{equation}
\frac{d}{dx}m(z,x) = V(x) - z - m^2 (z,x). \label{7.1.8}
\end{equation}
Obviously, $m(z,x)$ depends only on $V$ on $(x,b)$ (and on $h$ if
$b<\infty)$. A basic result of the inverse spectral theory says that the
converse is true as was shown independently by Borg \cite{Bo52} and
Marchenko \cite{Ma73} in 1952:
\begin{theorem} \label{t7.1.1} $m$ uniquely determines $V$. Explicitly, if $V_j$
are potentials with corresponding $m$-functions $m_j$, $j=1,2$, and
$m_1 =m_2$, then $V_1=V_2$ a.e.\ $($including $h_1 = h_2$$)$.
\end{theorem}
In 1999, Simon \cite{Si99} spectacularly improved this to obtain a local
version of the Borg--Marchenko uniqueness result as follows:
\begin{theorem} \label{t7.1.2} If $(V_1, b_1, h_1)$, $(V_2, b_2, h_2)$
are two potentials and $a <\min(b_1, b_2)$ and if
\begin{equation}
V_1 (x) = V_2 (x) \, \text{ on } \, (0,a), \label{7.1.9}
\end{equation}
then as $\kappa\to\infty$,
\begin{equation}
m_1 (-\kappa^2) - m_2 (-\kappa^2) = \tilde O(e^{-2\kappa a}).
\label{7.1.10}
\end{equation}
Conversely, if \eqref{7.1.10} holds, then \eqref{7.1.9} holds.
\end{theorem}
In \eqref{7.1.10}, we use the symbol $\tilde O$ defined by
\begin{align*}
&f=\tilde O(g) \, \text{
as $x\to x_0$ (where $\lim_{x\to x_0} g(x)=0$)} \\
& \quad \text{if and only if
$\lim_{x\to x_0} \frac{|f(x)|}{|g(x)|^{1-\varepsilon}} =0$ for
all $\varepsilon >0$.}
\end{align*}
From a results point of view, this local version of the
Borg--Marchenko uniqueness theorem was the most significant new
result in Simon's paper \cite{Si99}, but a major thrust of this paper was
the new set of methods introduced which led to a new approach of the inverse spectral problem. Theorem \ref{t7.1.2} implies that $V$ is
determined by the asymptotics of $m(-\kappa^2)$ as $\kappa \to \infty$. One
can also read off differences of the boundary condition from these
asymptotics. Moreover, the following result is proved in \cite{Si99}:
\begin{theorem} \label{t7.1.3} Let $(V_1, b_1, h_1)$, $(V_2, b_2, h_2)$
be two potentials and suppose that
\begin{equation}
b_1 = b_2 \equiv b <\infty, \quad |h_1| + |h_2| < \infty,
\quad V_1(x) = V_2 (x) \, \text{ on } \, (0,b). \label{7.1.11}
\end{equation}
Then
\begin{equation}
\lim_{\kappa\to\infty} e^{2b\kappa}
|m_1 (-\kappa^2) - m_2 (-\kappa^2)| = 4(h_1 - h_2). \label{7.1.12}
\end{equation}
Conversely, if \eqref{7.1.12} holds for some $b <\infty$ with a
limit in $(0,\infty)$, then \eqref{7.1.11} holds.
\end{theorem}
\medskip
To understand Simon's new approach, it is useful to recall briefly
the two approaches to the inverse problem for Jacobi matrices on
$\ell^2 ({\mathbb{N}}_0)$ \cite[Ch.\ VII]{Be68}, \cite{GS97a},
\cite{St95}:
$$
A = \begin{pmatrix} b_0 & a_0 & 0 & 0 & \dots \\
a_0 & b_1 & a_1 & 0 & \dots \\
0 & a_1 & b_2 & a_2 & \dots \\
\dots & \dots & \dots & \dots & \dots \end{pmatrix}
$$
with $a_j >0, b_j\in{\mathbb{R}}$. Here the $m$-function is just $(\delta_0,
(A-z)^{-1}\delta_0) = m(z)$ and, more generally, $m_n(z)=
(\delta_n, (A^{(n)}-z)^{-1}\delta_n)$ with $A^{(n)}$ on
$\ell^2 (\{n, n+1, \dots\})$ obtained by truncating the
first $n$ rows and $n$ columns of $A$. Here $\delta_n$ is the
Kronecker vector, that is, the vector with $1$ in slot $n$ and
$0$ in other slots. The fundamental theorem in this case is that
$m(z) \equiv m_0(z)$ determines the $b_n$'s and $a_n$'s.
$m_n(z)$ satisfies an analog of the Riccati equation \eqref{7.1.8}:
\begin{equation}
a^2_n m_{n+1}(z) = b_n - z - \frac{1}{m_n(z)}\, . \label{7.1.13}
\end{equation}
One solution of the inverse problem is to turn \eqref{7.1.13} around to
see that
\begin{equation}
m_n (z)^{-1} = -z + b_n - a^2_n m_{n+1} (z) \label{7.1.14}
\end{equation}
which, first of all, implies that as $z\to\infty$, $m_n(z) =
-z^{-1} + O(z^{-2})$, so \eqref{7.1.14} implies
\begin{equation}
m_n (z)^{-1} = -z + b_n + a^2_n z^{-1} + O(z^{-2}). \label{7.1.15}
\end{equation}
Thus, \eqref{7.1.15} for $n=0$ yields $b_0$ and $a^2_0$ and so $m_1(z)$
by \eqref{7.1.13}, and then an obvious induction yields successive
$b_k$, $a^2_k$, and $m_{k+1}(z)$.
A second solution involves orthogonal polynomials. Let $P_n(z)$
be the eigensolutions of the formal $(A-z)P_n =0$ with boundary
conditions $P_{-1}(z)=0$, $P_0(z)=1$. Explicitly,
\begin{equation}
P_{n+1}(z) = a^{-1}_n [(z-b_n)P_n (z)] -a_{n-1} P_{n-1}.
\label{7.1.16}
\end{equation}
Let $d\rho$ be the spectral measure for $A$ and vector $\delta_0$ so
that
\begin{equation}
m(z) = \int_{{\mathbb{R}}} \frac{d\rho (\lambda)}{\lambda-z}. \label{7.1.17}
\end{equation}
Then one can show that
\begin{equation}
\int_{{\mathbb{R}}} d\mu(\lambda) \, P_n(\lambda) P_m(\lambda) =\delta_{n,m}, \quad
n,m\in{\mathbb{N}}_0. \label{7.1.18}
\end{equation}
Thus, $P_n(z)$ is a polynomial of degree $n$ with positive
leading coefficients determined by \eqref{7.1.18}. These orthonormal
polynomials are determined via Gram--Schmidt from $\rho$ and by
\eqref{7.1.17} from $m$. Once one has the polynomials $P_n$, one can
determine the $a$'s and $b$'s from equation \eqref{7.1.16}.
Of course, these approaches via Riccati equation and orthogonal
polynomials are not completely disjoint. The Riccati solution
gives the $a_n$'s and $b_n$'s as continued fractions and the
connection between continued fractions and orthogonal polynomials
played a fundamental role in Stieltjes' work \cite{St95} on the moment problem
in 1895.
The Gel'fand--Levitan approach to the continuum case (cf.\ \cite{GL55},
\cite[Ch.\ 2]{Le87}, \cite{Ma73}, \cite[Ch.\ 2]{Ma86}) is a direct analog
of this orthogonal polynomial case. One looks at solutions $U(k,x)$ of
\begin{equation}
-U''(k,x) + V(x)U(k,x) = k^2 U(k,x) \label{7.1.19}
\end{equation}
satisfying $U(k,0)=1$, $U'(k,0)=ik$, and proves that they satisfy a
representation
\begin{equation}
U(k,x)=e^{ikx} + \int_{-x}^x dy \, K(x,y) e^{iky}, \label{7.1.20}
\end{equation}
the analog of $P_n(z)=cz^n +$ lower order. One defines $s(k,x)
=(2ik)^{-1} [U(k,x) - U(-k,x)]$ which satisfies \eqref{7.1.19} with
$s(k,0_+) =0$, $s'(k,0_+)=1$.
The spectral measure $d\rho$ associated to $m(z)$ by
$$
d\rho (\lambda) = (2\pi)^{-1} \lim_{\varepsilon\downarrow 0} [
\Im (m(\lambda + i\varepsilon))\, d\lambda]
$$
satisfies
\begin{equation}
\int_{{\mathbb{R}}} d\rho(k^2) \, s(k,x) s(k,y) = \delta (x-y), \label{7.1.21}
\end{equation}
at least formally. \eqref{7.1.20} and \eqref{7.1.21} yield an integral
equation for $K$ depending only on $d\rho$ and then once one has $K$,
one can find $U$ and hence $V$ via \eqref{7.1.19} (or via another relation
between $K$ and $V$).
The principal goal in \cite{Si99} was to present a new approach to the
continuum case, that is, an analog of the Riccati equation
approach to the discrete inverse problem. The simple idea for
this is attractive but has a difficulty to overcome. $m(z,x)$
determines $V(x)$, at least if $V$ is continuous by the known
asymptotics (\cite{DL91}, \cite{Ry02}):
\begin{equation}
m(-\kappa^2, x) = -\kappa - \frac{V(x)}{2\kappa} + o(\kappa^{-1}).
\label{7.1.22}
\end{equation}
We can therefore think of \eqref{7.1.8} with $V$ defined by \eqref{7.1.22}
as an evolution equation for $m$. The idea is that using a suitable
underlying space and uniqueness theorem for solutions of
differential equations, \eqref{7.1.8} should uniquely determine $m$ for
all positive $x$, and hence $V(x)$ by \eqref{7.1.22}.
To understand the difficulty, consider a potential $V(x)$ on
the whole real line. There are then functions $u_\pm (z,x)$
defined for $z\in{\mathbb{C}}\backslash [\beta, \infty)$ which are
$L^2$ at $\pm\infty$ and two $m$-functions $m_\pm (z,x) =
u'_\pm (z,x)/u_\pm (z,x)$. Both satisfy \eqref{7.1.8}, yet $m_+
(z,0)$ determines and is determined by $V$ on $(0,\infty)$
while $m_- (z,0)$ has the same relation to $V$ on $(-\infty, 0)$.
Put differently, $m_+ (z,0)$ determines $m_+ (z,x)$ for $x>0$
but not at all for $x<0$. $m_-$ is the reverse. So uniqueness
for \eqref{7.1.8} is one-sided and either side is possible! That this
does not make the scheme hopeless is connected with the fact
that $m_-$ does not satisfy \eqref{7.1.22}, but rather
\begin{equation}
m_- (-\kappa^2, x)=\kappa + \frac{V(x)}{2\kappa} +
o(\kappa^{-1}). \label{7.1.23}
\end{equation}
We will see the one-sidedness of the solvability is intimately
connected with the sign of the leading $\pm\kappa$ term in
\eqref{7.1.22}, \eqref{7.1.23}.
The key object in this new approach is a function $A(\alpha)$
defined for $\alpha \in (0,b)$ related to $m$ by
\begin{equation}
m(-\kappa^2) = -\kappa - \int_0^a d\alpha\,A(\alpha) e^{-2\alpha\kappa}
+\tilde O(e^{-2a\kappa}) \label{7.1.24}
\end{equation}
as $\kappa\to\infty$. We have written $A(\alpha)$ as a function
of a single variable but we will allow similar dependence on
other variables. Since $m(-\kappa^2, x)$ is also an $m$-function,
\eqref{7.1.24} has an analog with a function $A(\alpha, x)$.
By uniqueness of inverse Laplace transforms (see
\cite[Appendix 2, Theorem A.2.2]{Si99}), \eqref{7.1.24} and $m$ near
$-\infty$ uniquely determine $A(\alpha)$.
Not only will \eqref{7.1.24} hold but, in a sense, $A(\alpha)$ is close
to $V(\alpha)$. Explicitly, one can prove the following result:
\begin{theorem} \label{t7.1.4} Let $m$ be the $m$-function of the
potential $V$. Then there is a function $A \in L^1
([0,b])$ if $b< \infty$ and $A \in L^1 ([0,a])$ for all
$a<\infty$ if $b=\infty$ so that \eqref{7.1.24} holds for any
$a\leq b$ with $a<\infty$. $A(\alpha)$ only depends on $V(y)$
for $y\in [0,\alpha]$. Moreover, $A(\alpha)=V(\alpha) +
E(\alpha)$ where $E(\alpha)$ is continuous and satisfies
$$
|E(\alpha)|\leq \biggl( \int_0^\alpha dy \, |V(y)|\biggr)^2
\exp\biggl(\alpha \int_0^\alpha dy \,|V(y)|\biggr).
$$
\end{theorem}
Restoring the $x$-dependence, we see that $A(\alpha, x) =
V(\alpha + x) + E(\alpha, x)$ where
$$
\lim_{\alpha\downarrow 0}\, \sup_{0\leq x\leq a} |E(\alpha, x)|=0
$$
for any $a>0$, so
\begin{equation}
\lim_{\alpha\downarrow 0} A(\alpha, x) = V(x), \label{7.1.26}
\end{equation}
where this holds in general in the $L^1$-sense. If $V$ is
continuous, \eqref{7.1.26} holds pointwise. In general, \eqref{7.1.26} will
hold at any point of right Lebesgue continuity of $V$.
Because $E$ is continuous, $A$ determines any discontinuities
or singularities of $V$. More is true. If $V$ is $C^k$, then
$E$ is $C^{k+2}$ in $\alpha$, and so $A$ determines
$k^{\text{th}}$ order kinks in $V$. Much more is true and one can also
prove the following result:
\begin{theorem} \label{t7.1.5} $V$ on $[0,a]$ is only a function of
$A$ on $[0,a]$. Explicitly, if $V_1, V_2$ are two potentials,
let $A_1, A_2$ be their $A$-functions. If $a<b_1$, $a<b_2$, and
$A_1(\alpha)=A_2(\alpha)$ for $\alpha \in [0,a]$, then
$V_1 (x)=V_2 (x)$ for $x\in[0,a]$.
\end{theorem}
Theorems \ref{t7.1.4} and \ref{t7.1.5} imply Theorem
\ref{t7.1.2}.
As noted, the singularities of $V$ come from singularities of
$A$. A boundary condition is a kind of singularity, so one
might hope that boundary conditions correspond to very
singular $A$. In essence, we will see that this is the case -- there are
delta-function and delta-prime singularities at $\alpha =b$. Explicitly,
one can prove the following result:
\begin{theorem} \label{t7.1.6} Let $m$ be the $m$-function for a
potential $V$ with $b<\infty$. Then for $a<2b$,
\begin{equation}
m(-\kappa^2)=-\kappa -\int_0^a d\alpha \, A(\alpha) e^{-2\alpha\kappa}
- A_1 \kappa e^{-2\kappa b} - B_1 e^{-2\kappa b}
+\tilde O(e^{-2a\kappa}), \label{7.1.27}
\end{equation}
where the following facts hold: \\
$(a)$ If $h=\infty$, then $A_1=2, \quad B_1 = -2
\int_0^b V(y)\, dy$. \\
$(b)$ If $|h|<\infty$, then $A_1=-2, \quad B_1 =
2[2h+\int_0^b V(y)\, dy]$.
\end{theorem}
This implies Theorem \ref{t7.1.3}.
The reconstruction theorem, Theorem \ref{t7.1.5}, depends on the
differential equation that $A(\alpha, x)$ satisfies. Remarkably, $V$
drops out of the translation of \eqref{7.1.8} to the equation for $A$:
\begin{equation}
\frac{\partial A(\alpha,x)}{\partial x} =
\frac{\partial A(\alpha,x)}{\partial\alpha} +
\int_0^\alpha d\beta \, A(\beta,x) A(\alpha-\beta,x). \label{7.1.28}
\end{equation}
If $V$ is $C^1$, the equation holds in classical sense. For
general $V$, it holds in a variety of weaker senses. Either way,
$A(\alpha,0)$ for $\alpha\in [0,a]$ determines $A(\alpha,x)$ for
all $x,\alpha$ with $\alpha >0$ and $0<x+\alpha <a$. \eqref{7.1.26}
then determines $V(x)$ for $x\in [0,a)$. That is the essence
from which uniqueness comes. We will return to this circle of ideas later on when discussing Simon's approach to the inverse spectral problem in detail.
\smallskip
\begin{center}
$\bigstar$ \qquad $\bigstar$ \qquad $\bigstar$
\end{center}
\smallskip
Now we switch to \cite{GS00a} and take a closer look at some of the concepts introduced in \cite{Si99}. In particular, we continue the
study of the $A$-amplitude associated to half-line Schr\"odinger
operators, $-\frac{d^2}{dx^2}+ V$ in $L^2 ([0,b))$, $b\leq \infty$. $A$ is
related to the Weyl--Titchmarsh $m$-function via
$m(-\kappa^2) =-\kappa - \int_0^a d\alpha \, A(\alpha) e^{-2\alpha\kappa}
+O(e^{-(2a -\varepsilon)\kappa})$ for all $\varepsilon >0$. Three main issues will be
discussed:
\medskip
\noindent
$\bullet$ First, we describe how to extend the theory to general $V$
in $L^1 ([0,a])$ for all $a>0$, including $V$'s which are limit circle at
infinity. \\
$\bullet$ Second, the following relation between the
$A$-amplitude and the spectral measure $\rho$:
$$
A(\alpha) = -2\int_{-\infty}^\infty d\rho(\lambda) \, \lambda^{-\frac12}
\sin (2\alpha \sqrt{\lambda}),
$$
will be discussed. (Since the integral is
divergent, this formula has to be properly interpreted.) \\
$\bullet$ Third, a Laplace transform representation for $m$ without error term
in the case $b<\infty$ will be presented.
\medskip
We consider Schr\"odinger operators
\begin{equation} \label{1.1}
-\frac{d^2}{dx^2} + V
\end{equation}
in $L^2 ([0,b))$ for $0<b<\infty$ or $b=\infty$ and real-valued locally
integrable $V$. There are essentially four distinct cases.
\medskip
\noindent{\bf Case 1.} $b<\infty$. We suppose $V\in L^1 ([0,b])$. We then
pick $h\in {\mathbb{R}} \cup \{\infty\}$ and add the boundary condition at $b$
\begin{equation}
\label{1.2}
u'(b_-) + hu(b_-)=0,
\end{equation}
where $h=\infty$ is shorthand for the Dirichlet boundary condition
$u(b_-)=0$.
For Cases 2--4, $b=\infty$ and
\begin{equation}
\label{1.3}
\int_0^a dx \, |V(x)| < \infty \quad\text{for all }a<\infty.
\end{equation}
\medskip
\noindent{\bf Case 2.} $V$ is ``essentially'' bounded from below in the
sense that
\begin{equation} \label{1.3a}
\sup_{a>0} \left(\int_a^{a+1} dx\, \max (-V(x),0)\right) < \infty.
\end{equation}
Examples include $V(x) = c(x+1)^\beta$ for $c>0$ and all $\beta\in{\mathbb{R}}$
or $V(x) = -c (x+1)^\beta$ for all $c>0$ and $\beta \leq 0$.
\medskip
\noindent{\bf Case 3.} \eqref{1.3a} fails but \eqref{1.1} is limit point at
$\infty$ (see, e.g., \cite[Ch.\ 9]{CL55}, \cite[Sect.\ X.1]{RS75} for a
discussion of limit point/limit circle), that is, for each $z\in{\mathbb{C}}_+ =
\{z\in{\mathbb{C}} \,|\, \Im (z)>0\}$,
\begin{equation}
\label{1.4}
-u'' + Vu = zu
\end{equation}
has a unique solution, up to a multiplicative constant, which is $L^2$ at
$\infty$. An example is $V(x) = -c(x+1)^\beta$ for $c>0$ and
$0< \beta\leq 2$.
\medskip
\noindent{\bf Case 4.} \eqref{1.1} is limit circle at infinity, that is,
every solution of \eqref{1.4} is $L^2 ([0,\infty))$ at infinity if
$z\in{\mathbb{C}}_+$. We then pick a boundary condition by choosing a nonzero
solution $u_0$ of \eqref{1.4} for $z=i$. Other functions $u$ satisfying
the associated boundary condition at infinity then are supposed to satisfy
\begin{equation}
\label{1.5}
\lim_{x\to\infty} W(u_0,u)(x)=\lim_{x\to\infty} [u_0 (x) u'(x) - u'_0 (x) u(x)] = 0.
\end{equation}
Examples include $V(x) = -c(x+1)^\beta$ for $c>0$ and $\beta >2$.
\medskip
The Weyl--Titchmarsh $m$-function, $m(z)$, is defined for $z\in{\mathbb{C}}_+$ as
follows. Fix $z\in{\mathbb{C}}_+$. Let $u(x,z)$ be a nonzero solution of
\eqref{1.4} which satisfies the boundary condition at $b$. In Case 1, that
means $u$ satisfies \eqref{1.2}; in Case 4, it satisfies \eqref{1.5}; and
in Cases 2--3, it satisfies
$\int_R^\infty |u (z,x)|^2\, dx<\infty$ for some (and hence for all)
$R\geq 0$. Then,
\begin{equation}
\label{1.6}
m(z) = \frac{u'(z,0_+)}{u(z,0_+)}
\end{equation}
and, more generally,
\begin{equation}\label{1.7}
m(z,x) = \frac{u'(z,x)}{u(z,x)}\,.
\end{equation}
$m(z,x)$ satisfies the Riccati equation
(with $m'=\partial m/\partial x$),
\begin{equation}\label{1.8a}
m'(z,x) = V(x) - z - m(z,x)^2.
\end{equation}
$m$ is an analytic function of $z$ for $z\in{\mathbb{C}}_+$, and the following properties hold:
\medskip
\noindent{\bf Case 1.} $m$ is meromorphic in ${\mathbb{C}}$ with a discrete set
$\lambda_1 < \lambda_2 < \cdots$ of poles on ${\mathbb{R}}$ (and none on $(-\infty,
\lambda_1)$).
\smallskip
\noindent{\bf Case 2.} For some $\beta\in{\mathbb{R}}$, $m$ has an analytic continuation
to ${\mathbb{C}}\backslash[\beta, \infty)$ with $m$ real on $(-\infty, \beta)$.
\smallskip
\noindent{\bf Case 3.} In general, $m$ cannot be continued beyond ${\mathbb{C}}_+$
(there exist $V$'s where $m$ has a dense set of polar singularities on
${\mathbb{R}}$).
\smallskip
\noindent{\bf Case 4.} $m$ is meromorphic in ${\mathbb{C}}$ with a discrete set of
poles (and zeros) on ${\mathbb{R}}$ with limit points at both $+\infty$ and $-\infty$.
\smallskip
Moreover,
\[
\text{if } z\in{\mathbb{C}}_+ \text{ then } m(z,x)\in{\mathbb{C}}_+ ,
\]
so $m$ admits the Herglotz representation,
\begin{equation}
\label{1.8b}
m(z) = \Re(m(i)) + \int_{\mathbb{R}} d\rho(\lambda)
\left[ \frac{1}{\lambda-z} -\frac{\lambda}{1+\lambda^2} \right], \quad z\in{\mathbb{C}}\backslash{\mathbb{R}},
\end{equation}
where $\rho$ is a positive measure called the spectral measure, which satisfies
\begin{gather}
\int_{\mathbb{R}} \frac{d\rho(\lambda)}{1+|\lambda|^2} < \infty , \label{1.9} \\
d\rho(\lambda) = \wlim_{\varepsilon\downarrow 0}
\frac{1}{\pi} \Im (m(\lambda + i\varepsilon))\, d\lambda , \label{1.10}
\end{gather}
where $\wlim$ is meant in distributional sense.
\smallskip
All these properties of $m$ are well known (see, e.g.\
\cite[Ch.\ 2]{LS75}).
\medskip
In \eqref{1.8b}, the constant $\Re (m(i))$ is determined
by the result of Everitt \cite{Ev72} that for each $\varepsilon>0$,
\begin{equation}
\label{1.11}
m(-\kappa^2) = -\kappa +o(1) \quad\text{as } |\kappa| \to \infty
\text{ with } -\frac{\pi}{2} + \varepsilon < \arg (\kappa) < - \varepsilon < 0.
\end{equation}
Atkinson \cite{At81} improved \eqref{1.11} to read,
\begin{equation}
\label{1.12}
m(-\kappa^2) = -\kappa + \int_0^{a_0} d\alpha \, V(\alpha)
e^{-2\alpha\kappa} + o(\kappa^{-1})
\end{equation}
again as $|\kappa|\to\infty$ with $- \frac{\pi}{2} +\varepsilon < \arg(\kappa) <
- \varepsilon <0$ (actually, he allows $\arg(\kappa) \to 0$ as
$|\kappa| \to \infty$
as long as $\Re (\kappa) >0$ and $\Im (\kappa) > -\exp(-D|\kappa|)$ for
suitable $D$). In \eqref{1.12}, $a_0$ is any fixed $a_0 >0$.
One of the main results in \cite{GS00a} was to go way beyond the two
leading orders in \eqref{1.12}.
\begin{theorem} \label{t1} There exists a function $A(\alpha)$ for $\alpha
\in [0,b)$ so that $A\in L^1 ([0,a])$ for all $a<b$ and
\begin{equation}
\label{1.13}
m(-\kappa^2) = -\kappa - \int_0^a d\alpha \,A(\alpha) e^{-2\alpha\kappa}
+ \tilde O(e^{-2\alpha\kappa})
\end{equation}
as $|\kappa|\to\infty$ with $-\frac{\pi}{2} + \varepsilon<\arg(\kappa)< - \varepsilon <0$.
Here we say $f=\tilde O(g)$ if $g\to 0$ and for all
$\varepsilon >0$, $\big(\frac{f}{g}\big)|g|^\varepsilon\to 0$ as
$|\kappa| \to \infty$. Moreover, $A-q$ is continuous and
\begin{equation}
\label{1.14}
|(A-q)(\alpha)|\leq \left[\int_0^\alpha dx \,|V(x) \right]^2 \exp
\left(\alpha \int_0^\alpha dx \, |V(x)|\right).
\end{equation}
\end{theorem}
This result was proved in Cases 1 and 2 in \cite{Si99}. The proof of this
result if one only assumes \eqref{1.3} (i.e., in Cases~3 and 4) has been
provided in \cite{GS00a}.
Actually, in \cite{Si99}, \eqref{1.13} was proved in Cases~1 and 2 for
$\kappa$ real with $|\kappa| \to\infty$. The proof in \cite{GS00a}, assuming
only condition \eqref{1.3},
includes Case~2 in the general $\kappa$-region $\arg(\kappa)\in (-\frac{\pi}{2}
+ \varepsilon, -\varepsilon)$ and, as can be shown, the proof also holds in this region
for Case~1.
\begin{remark} At first sight, it may appear that Theorem \ref{t1},
as stated, does not imply the $\kappa$ real result of \cite{Si99}, but
if the spectral measure $\rho$ of \eqref{1.8b} has $\text{\rm{supp}}(\rho)\subseteq [a,\infty)$
for some $a\in {\mathbb{R}}$, \eqref{1.13} extends to all $\kappa$ in $|\arg(\kappa)|
< \frac{\pi}{2} - \varepsilon$, $|\kappa| \geq a +1$. To see this, one notes by
\eqref{1.8b} that $m'(z)$ is bounded away from $[a,\infty)$ so one has
the a priori bound $|m(z)| \leq C|z|$ in the region $\Re (z) < a -1$.
This bound and a Phragm\'en--Lindel\"of argument let one extend \eqref{1.13}
to the real $\kappa$ axis.
\end{remark}
The following is a basic result from \cite{Si99}:
\begin{theorem} \label{t2} \mbox{\rm (Theorem~2.1 of \cite{Si99})} Let
$V\in L^1 ([0,\infty))$. Then there exists a function $A$ on
$(0,\infty)$ so that $A-V$ is continuous and satisfies \eqref{1.14} such
that for $\Re (\kappa) > \|V\|_1/2$,
\begin{equation} \label{1.15}
m(-\kappa^2) = -\kappa - \int_0^\infty d\alpha\, A(\alpha)
e^{-2\alpha\kappa}.
\end{equation}
\end{theorem}
\begin{remark} In \cite{Si99}, this is only stated for $\kappa$ real with
$\kappa > \|V\|_1/2$, but \eqref{1.14} implies that
$|A(\alpha) -V(\alpha)|
\leq \|V\|_1^2 \exp (\alpha \|V\|_1)$ so the right-hand side of
\eqref{1.15} converges to an analytic function in $\Re (\kappa) >
\|V\|_1/2$. Since $m(z)$ is analytic in ${\mathbb{C}}\backslash [\alpha, \infty)$
for suitable $\alpha$, we have equality in $\{ \kappa\in{\mathbb{C}} \,|\, \Re
(\kappa) > \|V\|_1/2\}$ by analyticity.
\end{remark}
Theorem \ref{t1} in all cases follows from Theorem \ref{t2} and the following
result which was proved in \cite{GS00a}:
\begin{theorem} \label{t3} Let $V_1, V_2$ be potentials defined on $(0, b_j)$
with $b_j >a$ for $j=1,2$. Suppose that $V_1 =V_2$ on $[0,a]$. Then in the
region $\arg (\kappa) \in (-\frac{\pi}{2}+\varepsilon, -\varepsilon)$, $|\kappa| \geq
K_0$, we have that
\begin{equation} \label{1.16}
|m_1 (-\kappa^2) - m_2 (-\kappa^2)| \leq C_{\varepsilon,\delta} \exp
(-2a \Re (\kappa)),
\end{equation}
where $C_{\varepsilon,\delta}$ depends only on $\varepsilon$, $\delta$, and
$\sup_{0\leq x\leq a} \big(\int_x^{x+\delta} dy\, |V_j (y)|\big)$, where $\delta>0$
is any number so that $a+\delta \leq b_j$, $j=1,2$.
\end{theorem}
\begin{remark} $(i)$ An important consequence of Theorem \ref{t3} is that
if $V_1(x) = V_2 (x)$ for $x\in [0,a]$, then $A_1(\alpha) = A_2(\alpha)$
for $\alpha\in [0,a]$. Thus, $A(\alpha)$ is only a function of $V$ on
$[0,\alpha]$. (We emphasize that, conversely, one can show that
also $V(x)$ is only a function of $A$ on $[0,x]$.) \\
$(ii)$ This implies Theorem \ref{t1} by taking $V_1 = V$ and $V_2 =
V\chi_{[0,a]}$ and using Theorem \ref{t2} on $V_2$. \\
$(iii)$ The actual proof implies \eqref{1.16} on a larger region than $\arg
(\kappa) \in (-\frac{\pi}{2}+\varepsilon, -\varepsilon)$. Basically, one needs $\Im
(\kappa) \geq -C_1 \exp(-C_2|\kappa|)$ as $\Re (\kappa) \to\infty$.
\end{remark}
The basic connection between the spectral
measure $d\rho$ and the $A$-amplitude established in \cite{GS00a} says
\begin{equation} \label{1.18}
A(\alpha) = -2 \int_{-\infty}^\infty d\rho(\lambda) \,\lambda^{-\f12}
\sin(2\alpha \sqrt{\lambda}\,).
\end{equation}
However, the integral in \eqref{1.18} is not convergent. Indeed, the asymptotics
\eqref{1.11} imply that $\int_0^R d\rho(\lambda)\sim\frac{2}{3\pi} R^{\f32}$
so \eqref{1.18} is never absolutely convergent. Thus, \eqref{1.18} has to be suitably
interpreted.
We will indicate how to demonstrate \eqref{1.18} as a distributional
relation, smeared in $\alpha$ on both sides by a function $f\in
C_0^\infty ((0,\infty))$. This holds for all $V$'s in Cases~1--4.
Finally, we will discuss an Abelianized version of \eqref{1.18}, namely,
\begin{equation} \label{1.19}
A(\alpha) = -2\lim_{\varepsilon\downarrow 0} \int_{-\infty}^\infty
d\rho(\lambda) \, e^{-\varepsilon\lambda}
\lambda^{-\f12} \sin(2\alpha\sqrt{\lambda}\,)
\end{equation}
at any point, $\alpha$, of Lebesgue continuity for $V$. \eqref{1.19} is
proved only for a restricted class of $V$'s including Case~1, 2 and those
$V$'s satisfying
$$
V(x) \geq -Cx^2, \quad x\geq R
$$
for some $R>0$, $C>0$, which are always in the limit point case at infinity.
Subsequently, we will use \eqref{1.19} as the point of departure for
relating $A(\alpha)$ to scattering data.
In order to prove \eqref{1.18} for finite $b$, one needs to analyze the
finite $b$ case extending \eqref{1.13} to all $a$ including $a=\infty$
(by allowing $A$ to have $\delta$ and $\delta'$ singularities at
multiples of $b$). This was originally done in
\cite{Si99} for $\kappa$ real and positive and $a<\infty$. We now need
results in the entire region $\Re (\kappa)\geq K_0$. Explicitly, the following
was proved in \cite{GS00a}:
\begin{theorem} \label{t6} In Case~1, there are $A_n, B_n$ for $n=1,2,\dots$, and a
function $A(\alpha)$ on $(0,\infty)$ with
\begin{align*}
& |A_n|\leq C, \quad |B_n| \leq Cn, \\
& \int_0^a d\alpha \, |A(\alpha)| \leq C\exp (K_0 |a|) \,
\text{ so that for $\Re (\kappa) > \f12 K_0$:} \\
& m(-\kappa^2) = -\kappa - \sum_{n=1}^\infty A_n \kappa e^{-2\kappa bn}
-\sum_{n=1}^\infty B_n e^{-2\kappa bn} - \int_0^\infty
d\alpha \, A(\alpha)
e^{-2\alpha\kappa}.
\end{align*}
\end{theorem}
\eqref{1.18} can be used to obtain a priori bounds on $\int_{-R}^0
d\rho(\lambda)$ as $R\to\infty$.
\medskip
Now we turn to more details and start by illustrating how to use the
Riccati equation and a priori control on $m_j$ to obtain exponentially
small estimates on $m_1 - m_2$.
\begin{lemma} \label{p2.1} Let $m_1, m_2$ be two absolutely
continuous functions on $[a,b]$ so that for some $Q\in L^1 ([a,b])$,
\begin{equation} \label{2.1}
m'_j (x) = Q(x) - m_j (x)^2, \quad j=1,2, \ x\in (a,b).
\end{equation}
Then
\[
[m_1(a) - m_2(a)] = [m_1(b) - m_2(b)] \exp
\left(\int_a^b dy \, [m_1(y) + m_2(y)] \right).
\]
\end{lemma}
As an immediate corollary, one obtains the following result (which implies
Theorem~\ref{t3}):
\begin{theorem} \label{t2.2} Let $m_j (-\kappa^2,x)$ be functions defined for
$x\in [a,b]$ and $\kappa\in K$ some region of ${\mathbb{C}}$. Suppose that for each
$\kappa$ in $K$\!, $m_j$ is absolutely continuous in $x$ and satisfies
{\rm (}note that $V$ is the same for $m_1$ and $m_2${\rm)},
\[
m'_j(-\kappa^2,x) = V(x) + \kappa^2 - m_j (-\kappa^2,x)^2, \quad j=1,2.
\]
Suppose $C$ is such that for each $x\in [a,b]$ and $\kappa\in K$\!,
\begin{equation} \label{2.2}
|m_j (-\kappa^2,x) + \kappa| \leq C, \quad j=1,2,
\end{equation}
then
\begin{equation} \label{2.3}
|m_1 (-\kappa^2,a) - m_2 (-\kappa^2,a)| \leq 2C \exp [-2(b-a)
[\Re (\kappa) -C]].
\end{equation}
\end{theorem}
Theorem~\ref{t2.2} places importance on a priori bounds of the form
\eqref{2.2}. Fortunately, by modifying ideas of Atkinson \cite{At81}, we
can obtain estimates of this form as long as $\Im (\kappa)$ is bounded
away from zero.
Atkinson's method allows one to estimate $|m(-\kappa^2)+\kappa|$ in two
steps. We will fix some $a<b$ finite and define $m_0(-\kappa^2)$ by
solving
\begin{subequations} \label{3.3}
\begin{align}
m'_0 (-\kappa^2, x) &= V(x) + \kappa^2 - m_0 (-\kappa^2, x)^2, \label{3.3a} \\
m_0(-\kappa^2, a) &= - \kappa \label{3.3b} \\
\intertext{and then setting}
m_0 (-\kappa^2) & := m_0 (-\kappa^2, 0_+). \label{3.3c}
\end{align}
\end{subequations}
One then proves the following result.
\begin{lemma} \label{p3.1} There is a $C>0$ depending only on $V$ and a
universal constant $E>0$ so that if $\Re (\kappa) \geq C$ and $\Im
(\kappa)
\neq 0$, then
\begin{equation} \label{3.4}
|m(-\kappa^2) -m_0(-\kappa^2)| \leq E\, \frac{|\kappa|^2}{|\Im (\kappa)|}\,
e^{-2a\Re (\kappa)}.
\end{equation}
In fact, one can take
\[
C =\max \left(a^{-1} \ln (6), \, 4 \int_0^a dx\, |V(x)| \right), \quad
E = \frac{3 \cdot 2 \cdot 12^2}{5}\, .
\]
\end{lemma}
\begin{lemma} \label{p3.2} There exist constants $D_1$ and $D_2$
{\rm(}depending only on $a$ and $V${\rm)}, so that for $\Re (\kappa)
>D_1$,
\[
|m_0 (-\kappa^2) + \kappa| \leq D_2.
\]
Indeed, one can take
\[
D_1 = D_2 = 2\int_0^a dx \, |V(x)|.
\]
\end{lemma}
These Lemmas together with Theorem \ref{t2} yield the following
explicit form of Theorem \ref{t3}.
\begin{theorem} \label{t3.3} Let $V_1, V_2$ be defined on $(0, b_j)$ with
$b_j >a$ for $j=1,2$. Suppose that $V_1 = V_2$ on $[0,a]$. Pick $\delta$
so that $a + \delta \leq \min(b_1, b_2)$ and let $\eta=\sup_{0\leq x \leq
a; j=1,2} (\int_x^{x+\delta} dy\, |V_j (y)|)$. Then if $\Re (\kappa) \geq
\max(4\eta,
\delta^{-1}\ln (6))$ and $\Im (\kappa)\neq 0$, one obtains
\[
|m_1 (-\kappa^2) - m_2(-\kappa^2)|\leq 2 g(\kappa) \exp (-2a[\Re (\kappa)
- g(\kappa)]),
\]
where
\[
g(\kappa) = 2\eta + \frac{864}{5}\, \frac{|\kappa|^2}{|\Im (\kappa)|}\,
e^{-2\delta\Re (\kappa)}\, .
\]
\end{theorem}
\begin{remark} $(i)$ To obtain Theorem~\ref{t3}, we need only note that
in the region $\arg (\kappa) \in (-\frac{\pi}{2}+\varepsilon, -\varepsilon)$,
$|\kappa|\geq K_0$, $g$ is bounded. \\
$(ii)$ We need not require that $\arg(\kappa)< -\varepsilon$ to obtain $g$
bounded. It suffices, for example, that $\Re (\kappa) \geq |\Im
(\kappa)| \geq e^{-\alpha\Re (\kappa)}$ for some $\alpha < 2\delta$. \\
$(iii)$ For $g$ to be bounded, we need not require that $\arg(\kappa) > -
\frac{\pi}{2} +\varepsilon$. It suffices that $|\Im (\kappa)| \geq \Re (\kappa)
\geq \alpha\ln [|\Im (\kappa)|]$ for some $\alpha >(2\delta)^{-1}$.
Unfortunately, this does not include the region $\Im (-\kappa^2)=c$,
$\Re (-\kappa^2)\to\infty$, where
$\Re (\kappa)$ goes to zero as $|\kappa|^{-1}$. However, as
$\Re (-\kappa^2)
\to\infty$, we only need that $|\Im(-\kappa^2)|\geq 2\alpha |\kappa| \ln
(|\kappa|)$.
\end{remark}
Next, we turn to finite $b$ representations with no
errors: Theorem \ref{t2} implies that if $b=\infty$ and $V\in L^1
([0,\infty))$, then
\eqref{1.15} holds, a Laplace transform representation for $m$ without
errors. It is, of course, of direct interest that such a formula holds,
but we are especially interested in a particular consequence of it --
namely, that it implies that the formula \eqref{1.13} with error holds in
the region $\Re (\kappa) > K_0$ with error uniformly bounded in
$\Im(\kappa)$; that is, one proves the following result:
\begin{theorem}\label{t4.1} If $V\in L^1 ([0, \infty))$ and $\Re (\kappa) >
\|V\|_1/2$, then for all $a>0$:
\begin{equation} \label{4.1}
\left| m(-\kappa^2) + \kappa + \int_0^a d\alpha \,
A(\alpha) e^{-2\alpha\kappa} \right| \leq \left[ \|V\|_1 + \frac{\|V\|^2_1
e^{a\|V\|_1}} {2\Re (\kappa) - \|V\|_1}\right] e^{-2a\Re (\kappa)}.
\end{equation}
\end{theorem}
The principal goal is to prove an analog of this result in the case
$b<\infty$. To do so, we will need to first prove an analog of
\eqref{1.15} in case $b<\infty$ -- something of interest in its own
right. The idea will be to mimic the proof of Theorem~2 from \cite{Si99}
but use the finite $b$, $V^{(0)} (x)=0$, $x\geq0$ Green's function where
\cite{Si99} used the infinite $b$ Green's function. The basic idea is
simple, but the arithmetic is a bit involved.
We will start with the $h=\infty$ case. Three functions for
$V^{(0)}(x)=0$, $x \geq 0$ are significant. First, the kernel of the
resolvent $(-\frac{d^2}{dx^2} + \kappa^2)^{-1}$ with $u(0_+)=u(b_-)=0$
boundary conditions. By an elementary calculation (see, e.g.,
\cite[Sect.\ 5]{Si99}), it has the form
\begin{equation} \label{4.2}
G^{(0)}_{h=\infty} (-\kappa^2,x,y) = \frac{\sinh(\kappa x_<)}{\kappa}
\left[ \frac{e^{-\kappa x_>} - e^{-\kappa (2b-x_>)}}{1-e^{-2\kappa b}}\right],
\end{equation}
with $x_<=\min(x,y)$, $x_> = \max(x,y)$.
The second function is
\begin{equation} \label{4.3}
\psi^{(0)}_{h=\infty} (-\kappa^2,x) = \lim_{y\downarrow 0}
\frac{\partial G^{(0)}_{h=\infty}}{\partial y}\, (-\kappa^2,x,y) =
\frac{e^{-\kappa x} - e^{-\kappa (2b-x)}}{1-e^{-2\kappa b}}
\end{equation}
and finally (notice that $\psi^{(0)}_{h=\infty}(-\kappa^2,0_+)=1$ and
$\psi^{(0)}_{h=\infty}$ satisfies the equations
$-\psi'' =-\kappa^2 \psi$ and $\psi(-\kappa^2,b_-)=0$):
\begin{equation} \label{4.4}
m^{(0)}_{h=\infty} (-\kappa^2)
= \psi^{(0)\prime}_{h=\infty} (-\kappa^2,0_+)
= -\frac{\kappa + \kappa e^{-2\kappa b}}{1-e^{-2\kappa b}}\, .
\end{equation}
In \eqref{4.4}, prime means $d/dx$.
Now fix $V\in C^\infty_0 ((0,b))$. The pair of formulas
\begin{equation*}
\begin{split}
&\left(-\frac{d^2}{dx^2} + V + \kappa^2 \right)^{-1} \\
&\qquad \qquad = \sum_{n=0}^\infty
(-1)^n \left( -\frac{d^2}{dx^2} + \kappa^2\right)^{-1} \left[ V
\left( -\frac{d^2}{dx^2} + \kappa^2\right)^{-1}\right]^n
\end{split}
\end{equation*}
and
\[
m(-\kappa^2) = \lim_{x<y;\, y\downarrow 0}
\frac{\partial^2 G (-\kappa^2,x,y)}{\partial x\partial y}
\]
yields the following expansion for the $m$-function of $-\frac{d^2}{dx^2} +
V$ with $u(b_-)=0$ boundary conditions.
\begin{lemma} \label{p4.2} Let $V\in C^\infty_0 ((0, b))$, $b<\infty$. Then
\begin{equation} \label{4.5}
m(-\kappa^2) = \sum_{n=0}^\infty M_n (-\kappa^2; V),
\end{equation}
where
\begin{align}
M_0 (-\kappa^2; V) &= m^{(0)}_{h=\infty} (-\kappa^2), \label{4.6} \\
M_1 (-\kappa^2; V) &= -\int_0^b V(x) \psi^{(0)}_{h=\infty}
(-\kappa^2,x)^2 \, dx , \label{4.7}
\end{align}
and for $n\geq 2$,
\begin{align}
\begin{split}
M_n (-\kappa^2; V) &= (-1)^n \int_0^b dx_1 \dots \int_0^b
dx_n \, V(x_1) \dots V(x_n) \\
& \quad \times \psi^{(0)}_{h=\infty}(-\kappa^2,x_1) \psi^{(0)}_{h=\infty}
(-\kappa^2,x_n) \prod_{j=1}^{n-1}
G^{(0)}_{h=\infty} (-\kappa^2, x_j, x_{j+1}) . \label{4.8}
\end{split}
\end{align}
\end{lemma}
The precise region of convergence is unimportant since one can
expand regions by analytic continuation. For now, we note it certainly converges
in the region $\kappa$ real with $\kappa^2 > \|V\|_\infty$.
Writing each term in \eqref{4.5} as a Laplace transform then yields the
following result:
\begin{theorem} \label{t4.3} {\rm{(Theorem~\ref{t6} for $h =\infty$)}}
Let $b<\infty$, $h =\infty$, and $V\in L^1 ([0,b])$. Then for $\Re
(\kappa) > \|V\|_1/2$, we have that
\begin{equation} \label{4.22}
m(-\kappa^2) = -\kappa - \sum_{j=1}^\infty A_j \kappa e^{-2\kappa bj}
-\sum_{j=1}^\infty B_j e^{-2\kappa bj} - \int_0^\infty d\alpha\, A(\alpha)
e^{-2\alpha\kappa} ,
\end{equation}
where
\begin{align*}
& A_j =2, \quad B_j = -2j\int_0^b dx \, V(x), \quad j\in{\mathbb{N}}, \\
& |A(\alpha) - A_1 (\alpha)| \leq \frac{(2\alpha
+ b)(2\alpha +2b)}
{2b^2} \|V\|^2_1 \exp (\alpha \|V\|_1)\, \text{ with $A_1$ given by} \\
& A_1(\alpha) = \begin{cases}
V(\alpha), & 0 \leq\alpha < b, \\
(n+1) V(\alpha - nb) + nq ((n+1)b-\alpha), & nb\leq\alpha < (n+1)b, \;
n\in{\mathbb{N}} .\end{cases}
\end{align*}
In particular, for all $a\in(0,b)$,
\[
\int_0^a d\alpha \,|A(\alpha)| \leq C (b, \|V\|_1) (1+a^2) \exp(a\|V\|_1).
\]
\end{theorem}
This implies the following estimate:
\begin{corollary} \label{c4.4} If $V\in L^1 ([0,\infty))$ and $\Re (\kappa)
\geq\f12 \|V\|_1 + \varepsilon$, then for all $a\in(0,b)$, $b<\infty$, we have
that
\[
\left| m(-\kappa^2) + \kappa + \int_0^a d\alpha \,
A(\alpha) e^{-2\alpha\kappa} \right| \leq C(a,\varepsilon) e^{-2a\Re (\kappa)},
\]
where $C(a,\varepsilon)$ depends only on $a$ and $\varepsilon$ {\rm(}and
$\|V\|_1$\rm{)} but not on $\Im(\kappa)$.
\end{corollary}
Next, we turn to the case $h\in{\mathbb{R}}$. Then \eqref{4.2}--\eqref{4.4} become
\begin{align}
G^{(0)}_{h}(-\kappa^2,x,y) &= \frac{\sinh (\kappa x_<)}{\kappa}\,
\psi^{(0)}_{h} (-\kappa^2,x_>) , \label{4.26} \\
\psi^{(0)}_{h}(-\kappa^2,x) &= \left[ \frac{e^{-\kappa x} + \zeta(h,\kappa)
e^{-\kappa (2b-x)}}{1+\zeta (h,\kappa) e^{-2b\kappa}}\right], \label{4.27} \\
m^{(0)}_{h}(-\kappa^2) &= -\kappa + 2\kappa\, \frac{\zeta(h,\kappa) e^{-2\kappa b}}
{1+\zeta(h,\kappa) e^{-2\kappa b}}\, , \label{4.28}
\end{align}
where
\begin{equation} \label{4.29}
\zeta(h,\kappa) = \frac{\kappa-h}{\kappa+h}\,.
\end{equation}
This then leads to the following result:
\begin{theorem} \label{t4.7} {\rm{(Theorem~\ref{t6} for general
$h\in{\mathbb{R}}$)}} Let $b<\infty$, $|h|<\infty$, and $V\in L^1 ([0,b])$. Then
for
$\Re (\kappa) > \f12 D_1 [\|V\|_1 + |h|+b^{-1} +1]$ for a suitable
universal constant $D_1$,
\eqref{4.22} holds, where
\begin{align}
& A_j = 2(-1)^j, \quad B_j = 2(-1)^{j+1}
j\bigg[2h + \int_0^b dx \, V(x) \bigg], \\
& |A(\alpha) - V(\alpha)| \leq \|V\|^2_1\exp(\alpha \|V\|_1) \,
\text{ if $|\alpha| <b$, and for any $a>0$}, \\
& \quad \int_0^a d\alpha \,|A(\alpha)| \leq D_2 (b, \|V\|_1,h) \exp(D_1 a
(\|V\|_1 + |h|+b^{-1}+1)).
\end{align}
\end{theorem}
Hence, one obtains the following estimate:
\begin{corollary} \label{c4.8} Fix $b<\infty$, $V\in L^1 ([0,b])$, and $|h|
<\infty$. Fix $a<b$. Then there exist positive constants $C$ and $K_0$ so
that for all complex $\kappa$ with $\Re (\kappa) > K_0$,
\[
\left| m(-\kappa^2) + \kappa + \int_0^a d\alpha \,
A(\alpha) e^{-2\alpha\kappa} \right| \leq Ce^{-2a\kappa}.
\]
\end{corollary}
Next we return to the relation between $A$ and $\rho$ and
discuss a first distributional form of this relation: Our primary goal in
the following is to discuss a formula which formally
says that
\begin{equation} \label{5.1}
A(\alpha) = -2\int_{-\infty}^\infty d\rho(\lambda) \,
\lambda^{-\f12} \sin(2\alpha
\sqrt{\lambda}\,),
\end{equation}
where for $\lambda \leq 0$, we define
\[
\lambda^{-\f12}\sin(2\alpha\sqrt{\lambda}\,) = \begin{cases}
2\alpha &\text{if } \lambda =0, \\
(-\lambda)^{-\f12} \sinh(2\alpha \sqrt{-\lambda}\, ) &\text{if } \lambda <0.
\end{cases}
\]
In a certain sense which will become clear, the left-hand side of
\eqref{5.1} should be $A(\alpha) - A(-\alpha) + \delta'(\alpha)$.
To understand \eqref{5.1} at a formal level, note the basic formulas,
\begin{align}
m(-\kappa^2) &= -\kappa - \int_0^\infty d\alpha \,
A(\alpha) e^{-2\alpha\kappa},
\label{5.2} \\ m(-\kappa^2) &= \Re (m(i)) + \int_{-\infty}^\infty
d\rho(\lambda) \left[
\frac{1} {\lambda+\kappa^2} - \frac{\lambda}{1+\lambda^2}\right], \label{5.3}
\end{align}
and
\begin{equation} \label{5.4}
(\lambda + \kappa^2)^{-1} = 2\int_0^\infty d\alpha \,\lambda^{-\f12}
\sin(2\alpha \sqrt{\lambda}\,) e^{-2\alpha\kappa},
\end{equation}
which is an elementary integral if $\kappa >0$ and $\lambda >0$. Plug
\eqref{5.4} into \eqref{5.3}, formally interchange the order of integrations,
and \eqref{5.2} should only hold if \eqref{5.1} does. However, a closer
examination of this procedure reveals that the interchange of the order of
integrations is not justified and indeed \eqref{5.1} is not true as a
simple integral since,
$\int_0^R d\rho (\lambda) \simlim\limits_{R\to\infty}\frac{2}{3\pi} R^{\f32}$,
which implies that \eqref{5.1} is not absolutely convergent. We will even
see later that the integral sometimes fails to be conditionally convergent.
Our primary method for understanding \eqref{5.1} is as a distributional
statement, that is, it will hold when smeared in $\alpha$ for $\alpha$ in
$(0,b)$. We discuss this next if $V\in L^1 ([0,\infty))$ or if
$b<\infty$. Later it will be extended to all $V$ (i.e.,
all Cases 1--4) by a limiting argument. Subsequently, we will study
\eqref{5.1} as a pointwise statement, where the integral is defined as an
Abelian limit.
Suppose $b<\infty$ or $b=\infty$ and $V\in L^1 ([0,b))$. Fix $a<b$ and $f
\in C^\infty_0 ((0,a))$. Define
\begin{equation} \label{5.5}
m_a (-\kappa^2) := -\kappa - \int_0^a d\alpha \, A(\alpha)
e^{-2\alpha\kappa}
\end{equation}
for $\Re (\kappa) \geq 0$. Fix $\kappa_0$ real and let
\[
g(y,\kappa_0, a) := m_a (-(\kappa_0 + iy)^2),
\]
with $\kappa_0, a$ as real parameters and $y\in{\mathbb{R}}$ a variable. As usual,
define the Fourier transform by (initially for smooth functions and then
by duality for tempered distributions \cite[Ch.\ IX]{RS75})
\begin{equation} \label{5.6}
\hat F(k) = \frac{1}{\sqrt{2\pi}} \int_{\mathbb{R}} dy \, e^{-iky} F(y), \quad
\check F(k) = \frac{1}{\sqrt{2\pi}} \int_{\mathbb{R}} dy \, e^{iky} F(y).
\end{equation}
Then by \eqref{5.5},
\begin{equation}
\widehat{\bar g} (k, \kappa_0, a) = -\sqrt{2\pi}\, \kappa_0 \delta(k) -
\sqrt{2\pi}\, \delta'(k) - \frac{\sqrt{2\pi}}{2}\, e^{-k\kappa_0}
A\left(\frac{k}{2}\right) \chi_{(0,2a)} (k).
\end{equation}
Thus, since $f(0_+)=f'(0_+)=0$, in fact, $f$ has support away from $0$ and
$a$,
\begin{align}
\int_0^a d\alpha \,A(\alpha) f(\alpha) &= -\frac{2}{\sqrt{2\pi}} \int_0^a
d\alpha \,
\overline{\widehat{\bar g}}\, (2\alpha, \kappa_0, a) e^{2\alpha\kappa_0}
f(\alpha) \notag \\
&=-\frac{1}{\sqrt{2\pi}} \int_0^{2a} d\alpha \,
\overline{\widehat{\bar g}} (\alpha, \kappa_0, a)
e^{\alpha\kappa_0} f\left(\frac{\alpha}{2}\right) \notag \\
&= -\frac{1}{\sqrt{2\pi}} \int_{\mathbb{R}} dy \,
g(y, \kappa_0, a) \check F (y, \kappa_0),
\label{5.8}
\end{align}
where we have used the unitarity of $\,\, \widehat{} \,\,$ and
\begin{equation}
\check F (y, \kappa_0) = \frac{1}{\sqrt{2\pi}} \int_0^{2a} d\alpha \,
e^{\alpha (\kappa_0 +iy)} f\left(\frac{\alpha}{2}\right)
= \frac{2}{\sqrt{2\pi}} \int_0^a d\alpha \,
e^{2\alpha (\kappa_0 +iy)} f(\alpha).
\label{5.9}
\end{equation}
Notice that
\begin{equation} \label{5.10}
|\check F (y, \kappa_0)| \leq Ce^{2(a-\varepsilon)\kappa_0} (1+|y|^2)^{-1}
\end{equation}
since $f$ is smooth and supported in $(0, a-\varepsilon)$ for some $\varepsilon >0$.
By Theorem~\ref{t4.1} and Corollary~\ref{c4.8},
\begin{equation} \label{5.11}
|m_a (-(\kappa_0 + iy)^2) - m(-(\kappa_0 + iy)^2)| \leq Ce^{-2a\kappa_0}
\end{equation}
for large $\kappa_0$, uniformly in $y$. From \eqref{5.8}, \eqref{5.10}, and
\eqref{5.11}, one concludes the following fact:
\begin{lemma} \label{l5.1} Let $f\in C_0^\infty ((0,a))$ with $0<a<b$ and
$V\in L^1 ([0,b))$. Then
\begin{equation} \label{5.12}
\int_0^a d\alpha \, A(\alpha) f(\alpha)
= \lim_{\kappa_0 \uparrow \infty} \left[-\frac{1}{\pi} \int_{\mathbb{R}} dy \,
m(-(\kappa_0 + iy)^2)\int_0^a d\alpha \,e^{2\alpha(\kappa_0 +
iy)} f(\alpha) \right].
\end{equation}
\end{lemma}
As a function of $y$, for $\kappa_0$ fixed, the alpha integral is
$O((1+y^2)^{-N})$ for all $N$ because $f$ is $C^\infty$. Now define
\begin{equation} \label{5.13}
\tilde m_R (-\kappa^2)=\left[ c_R + \int_{\lambda\leq R}
\frac{d\rho(\lambda)}{\lambda + \kappa^2}\right],
\end{equation}
where $c_R$ is chosen so that $\tilde m_R\arrow\limits_{R\to\infty} m$. Because
$\int_{\mathbb{R}} \frac{d\rho(\lambda)}{1+\lambda^2} <\infty$, the convergence is uniform
in $y$ for $\kappa_0$ fixed and sufficiently large. Thus in \eqref{5.12} we can
replace $m$ by $m_R$ and take a limit (first $R\to\infty$ and then $\kappa_0
\uparrow\infty$). Since $f(0_+)=0$, the $\int dy\, c_R\, d\alpha$-integrand is
zero. Moreover, we can now interchange the $dy\,d\alpha$ and $d\rho(\lambda)$
integrals. The result is that
\begin{align}
\begin{split}
\int_0^a d\alpha\,A(\alpha) f(\alpha) &= \lim_{\kappa_0\uparrow\infty}\,
\lim_{R\to\infty} \int_{\lambda\leq R} d\rho(\lambda) \\
& \quad \times \left[ \int_0^a d\alpha\, e^{2\alpha\kappa_0} f(\alpha)
\left[ -\frac{1}{\pi} \int_{\mathbb{R}}
\frac{dy \,e^{2\alpha iy}}{(\kappa_0 + iy)^2 +\lambda}
\right]\right]. \label{5.14}
\end{split}
\end{align}
In the case at hand, $d\rho$ is bounded below, say $\lambda \geq -K_0$.
As long as we take $\kappa_0 > K_0$, the poles of $(\kappa_0 + iy)^2 + \lambda$
occur in the upper half-plane
\[
y_\pm = i\kappa_0 \pm\sqrt{\lambda}\, .
\]
Closing the contour in the upper plane, we find that if $\lambda \geq -K_0$,
\[
-\frac{1}{\pi} \int_{\mathbb{R}} \frac{dy\,e^{2\alpha iy}}{(\kappa_0 + iy)^2 + \lambda}
=
-2e^{-2\alpha\kappa_0}\, \frac{\sin(2\alpha\sqrt{\lambda}\,)}{\sqrt\lambda}\,.
\]
Thus \eqref{5.14} becomes
$$
\int_0^a d\alpha \, A(\alpha) f(\alpha)
= -2 \lim_{\kappa_0\uparrow\infty}\, \lim_{R\to\infty} \int_{\lambda\leq
R} d\rho(\lambda)
\left[\int_0^a d\alpha \, f(\alpha)\,
\frac{\sin(2\alpha\sqrt\lambda\,)}{\sqrt\lambda}\right].
$$
$\kappa_0$ has dropped out and the $\alpha$ integral is bounded by $C
(1+\lambda^2)^{-1}$, so one can take the limit as $R\to\infty$ since $\int_{\mathbb{R}}
\frac{d\rho(\lambda)}{1+\lambda^2} <\infty$. One is therefore led to the
following result.
\begin{theorem} \label{t5.2} Let $f\in C^\infty_0 ((0,a))$ with $a<b$ and either
$b<\infty$ or $V\in L^1 ([0, \infty))$ with $b=\infty$. Then
\begin{equation} \label{5.15}
\int_0^a d\alpha \, A(\alpha) f(\alpha) = -2 \int_{\mathbb{R}}
d\rho(\lambda) \left[ \int_0^a
d\alpha \,
f(\alpha)\, \frac{\sin(2\alpha\sqrt\lambda\,)}{\sqrt\lambda}\right].
\end{equation}
\end{theorem}
One can strengthen this in two ways. First, one wants to allow $a>b$
if $b<\infty$. As long as $A$ is interpreted as a distribution with
$\delta$ and $\delta'$ functions at $\alpha=nb$, this is easy. One also
wants to allow $f$ to have a nonzero derivative at $\alpha =0$. The net
result is described in the next theorem:
\begin{theorem} \label{t5.3} Let $f\in C^\infty_0 ({\mathbb{R}})$ with $f(-\alpha) =
-f(\alpha)$, $\alpha\in{\mathbb{R}}$ and either $b<\infty$ or
$V\in L^1 ([0,\infty))$ with $b=\infty$. Then
\begin{equation} \label{5.16}
-2\int_{\mathbb{R}} d\rho(\lambda) \left[ \int_{-\infty}^\infty
d\alpha \, f(\alpha) \,
\frac{\sin(2\alpha\sqrt\lambda\, )}{\sqrt\lambda} \right]
= \int_{-\infty}^\infty d\alpha \, \tilde A(\alpha) f(\alpha),
\end{equation}
where $\tilde A$ is the distribution
\begin{subequations} \label{5.17}
\begin{align}
\tilde A(\alpha) &= \chi_{(0,\infty)} (\alpha) A(\alpha) -
\chi_{(-\infty, 0)} (\alpha) A(-\alpha) + \delta'(\alpha) \label{5.17a} \\
\intertext{if $b=\infty$ and}
\tilde A(\alpha) &= \chi_{(0,\infty)} (\alpha) A(\alpha) - \chi_{(-\infty,
0)} (\alpha) A(-\alpha) + \delta' (\alpha) \notag \\
&\quad + \sum_{j=1}^\infty B_j [\delta (\alpha - 2bj) - \delta (\alpha +
2bj)] \notag \\
&\quad + \sum_{j=1}^\infty \tfrac12\, A_j [\delta' (\alpha - 2bj) +
\delta' (\alpha + 2bj)] \label{5.17b}
\end{align}
\end{subequations}
if $b<\infty$, where $A_j, B_j$ are $h$ dependent and given in
Theorems \ref{t4.3} and \ref{t4.7}.
\end{theorem}
\medskip
Next we change the subject temporarily and turn to bounds on $\int_0^{\pm R} d\rho(\lambda)$ which are of independent interest:
As we will see, \eqref{1.11} implies asymptotic results on $\int_{-R}^R
d\rho(\lambda)$, and \eqref{5.1} will show that $\int_{-\infty}^0
e^{b\sqrt{-\lambda}}d\rho(\lambda)<\infty$ for all $b>0$ and more. It
follows from \eqref{5.3} that
\[
\Im (m(ia)) = a\int_{\mathbb{R}} \frac{d\rho(\lambda)}{\lambda^2 + a^2}\,,
\quad a>0.
\]
Thus, Everitt's result \eqref{1.11} implies that
\[
\lim_{a\to\infty} a^{\f12} \int_{\mathbb{R}} \frac{d\rho(\lambda)}{\lambda^2 + a^2} =
2^{-\f12}.
\]
Standard Tauberian arguments (see, e.g., in \cite[Sect.\ III.10]{Si79},
which in this case shows that on even functions $R^{\f32} d\rho
(\lambda/R)\arrow\limits_{R\to\infty} (2 \pi)^{-1}
|\lambda|^{\f12}\, d\lambda$) then imply the following result:
\begin{theorem} \label{t6.1}
\begin{equation} \label{6.1}
\lim_{R\to\infty} R^{-\f32} \int_{-R}^R d\rho(\lambda) = \frac{2}{3\pi}\, .
\end{equation}
\end{theorem}
\begin{remark} $(i)$ This holds in all cases (1--4) we consider here,
including some with $\text{\rm{supp}}(d\rho)$ unbounded below. \\
$(ii)$ Since one can show that $\int_{-\infty}^0 d\rho$ is bounded, one can
replace $\int_{-R}^R$ by $\int_0^R$ in \eqref{6.1}.
\end{remark}
Next, we recall the following a priori bound that follows from
Lemmas \ref{p3.1} and \ref{p3.2}:
\begin{lemma} \label{p6.2} Let $d\rho$ be the spectral measure for a
Schr\"odinger operator in Cases 1--4. Fix $a <b$. Then there is a constant
$C_a$ depending only on $a$ and $\int_0^a dy \, |V(y)|$ so that
\begin{equation} \label{6.2}
\int_{\mathbb{R}} \frac{d\rho(\lambda)}{1+\lambda^2} \leq C_a.
\end{equation}
\end{lemma}
The goal is to bound $\int_{-\infty}^0 e^{2\alpha\sqrt{-\lambda}}$ $d\rho
(\lambda)$ for any $\alpha < b$ and to find an explicit bound in terms of
$\sup_{0\leq x \leq\alpha + 1}$ $[-V(y)]$ when that $\sup$ is finite. As
a preliminary, we need the following result from the standard limit
circle theory \cite[Sect.\ 9.4]{CL55}.
\begin{lemma}\label{p6.3} Let $b=\infty$ and let $d\rho$ be the spectral
measure for some Schr\"odinger operator in Cases 2--4. Let $d\rho_{R,h}$ be the spectral measure
for the problem with $b=R<\infty, h$ and potential equal to $V(x)$ for
$x\leq R$. Then there exists $h(R)$ so that
\[
d\rho_{R, h(R)}\arrow_{R\to\infty} d\rho,
\]
when smeared with any function $f$ of compact support. In particular, if
$f\geq 0$, then
\[
\int_{\mathbb{R}} d\rho(\lambda) \, f(\lambda) \leq
\operatornamewithlimits{\varlimsup}_{R\to\infty}
\int_{\mathbb{R}} d\rho_{R, h(R)} (\lambda) \, f(\lambda).
\]
\end{lemma}
This result implies that we need only obtain bounds for $b<\infty$ (where
we have already proved \eqref{5.15}).
\begin{lemma} \label{l6.4} If $\rho_1$ has support in $[-E_0, \infty)$, $E_0 >0$,
then
\begin{equation} \label{6.3}
\int_{-\infty}^0 e^{\gamma\sqrt{-\lambda}} \, d\rho_1 (\lambda) \leq
e^{\gamma\sqrt{E_0}} (1+ E_0^2) \int_{-\infty}^0 \frac{d\rho_1 (\lambda)}
{1+\lambda^2}\,.
\end{equation}
\end{lemma}
Lemmas {\ref{p6.2}, \ref{p6.3} and Lemma \ref{l6.4} imply the following result.
\begin{theorem} \label{t6.5} Let $\rho$ be the spectral measure for some Schr\"odinger operator in Cases 2--4. Let
\begin{align*}
&E(\alpha_0)
:= -\inf \bigg\{ \int_0^{\alpha_0 +1} dx \, (|\varphi'_n(x)|^2 + V(x)
|\varphi(x)|^2) \biggm| \varphi\in
C^\infty_0 ((0,\alpha_0 +1)), \\
&\hspace*{2.6cm} \int_0^{\alpha_0 +1} dx\,|\varphi(x)|^2 \leq 1 \bigg\}.
\end{align*}
Then for all $\delta >0$ and $\alpha_0 >0$,
\begin{equation}
\alpha_0 \delta \int_{-\infty}^0 e^{2(1-\delta)\alpha_0\sqrt{-\lambda}}
\, d\rho (\lambda)
\leq \biggl[ C_1 (1+\alpha_0) + C_2 (1+E(\alpha_0)^2)
e^{2(\alpha_0 +1)\sqrt{E(\alpha_0)}} \, \biggr], \label{"6.7"}
\end{equation}
where $C_1, C_2$ only depend on $\int_0^1 dx\,|V(x)|$. In particular,
\begin{equation} \label{"6.8"}
\int_{-\infty}^0 e^{B\sqrt{-\lambda}} \, d\rho(\lambda) <\infty
\end{equation}
for all $B<\infty$.
\end{theorem}
As a special case, suppose $V(x) \geq -C(x+1)^2$. Then $E(\alpha_0) \geq
-C(\alpha_0 + 2)^2$ and we see that
\begin{equation} \label{6.7}
\int_{-\infty}^0 e^{B \sqrt{-\lambda}} \, d\rho(\lambda) \leq D_1
e^{D_2 B^2}.
\end{equation}
This implies the next result.
\begin{theorem} \label{t6.6} If $d\rho$ is the spectral measure for a potential
which satisfies
\begin{equation} \label{6.8}
V(x) \geq - Cx^2, \quad x\geq R
\end{equation}
for some $R>0$, $C>0$, then for $\varepsilon>0$ sufficiently small,
\begin{equation} \label{6.9}
\int_{-\infty}^0 e^{-\varepsilon\lambda} \, d\rho(\lambda) < \infty.
\end{equation}
\end{theorem}
If in addition $V\in L^1([0,\infty))$, then the
corresponding Schr\"odinger operator is bounded from below and hence
$d\rho$ has compact support on $(-\infty, 0]$. This fact will be useful
later in the scattering-theoretic context.
The estimate \eqref{"6.8"}, in the case of non-Dirichlet boundary
conditions at $x=0_+$, appears to be due to Marchenko \cite{Ma73}. Since
it is a fundamental ingredient in the inverse spectral problem, it
generated considerable attention; see, for instance, \cite{GL55},
\cite{Le73}, \cite{Le73a}, \cite{Le77}, \cite{LG64}, \cite{Ma73},
\cite[Sect.\ 2.4]{Ma86}. The case of a Dirichlet boundary at $x=0_+$
was studied in detail by Levitan \cite{Le77}. These authors, in addition
to studying the spectral asymptotics of $\rho(\lambda)$ as
$\lambda\downarrow -\infty$, were also particularly interested in the
asymptotics of $\rho(\lambda)$ as $\lambda\uparrow \infty$ and
established Theorem \ref{t6.1}. In the latter context,
we also refer to Bennewitz \cite{Be89}, Harris \cite{Ha97}, and the
literature cited therein. In contrast to these activities, we were not
able to find estimates of the type \eqref{"6.7"} (which implies
\eqref{"6.8"}) and \eqref{6.9} in the literature.
At this point one can return to the relation between $A$ and $\rho$ and
discuss a second distributional form of this relation which extends
Theorem \ref{t5.2} to all four cases.
\begin{theorem} \label{t7.1} Let $f\in C^\infty_0 ((0, \infty))$ and suppose
$b=\infty$. Assume $V$ satisfies \eqref{1.3} and let $d\rho$ be the
associated spectral measure and $A$ the associated $A$-function. Then
\eqref{5.16} and \eqref{5.17} hold.
\end{theorem}
Next we establish a third relation between $A$ and $\rho$ and turn to
Abelian limits:
For $f\in C^\infty_0 ({\mathbb{R}})$, define for $\lambda\in{\mathbb{R}}$,
\begin{equation} \label{8.1}
Q(f)(\lambda) = \int_{-\infty}^\infty d\alpha \, f(\alpha)
\frac{\sin(2\alpha \sqrt\lambda\,)}{\sqrt\lambda}
\end{equation}
and then
\begin{align}
T(f) &= -2 \int_{\mathbb{R}} d\rho(\lambda) \,Q(f)(\lambda) \label{8.2} \\
&= \int_{-\infty}^\infty d\alpha \, \tilde A(\alpha) f(\alpha). \label{8.3}
\end{align}
Relations \eqref{5.16}, \eqref{5.17} show that for
$f\in C^\infty_0 ({\mathbb{R}})$, the two expressions \eqref{8.2}, \eqref{8.3}
define the same $T(f)$. This was proved for odd $f$'s but both
integrals vanish for even $f$'s. Now one wants to use \eqref{8.2} to extend to a
large class of $f$, but needs to exercise some care not to use \eqref{8.3},
except for $f\in C^\infty_0 ({\mathbb{R}})$.
$Q(f)$ can be defined as long as $f$ satisfies
\begin{equation} \label{8.4}
|f(\alpha)| \leq C_k e^{-k|\alpha|}, \quad \alpha\in{\mathbb{R}}
\end{equation}
for all $k >0$. In particular, a simple calculation shows that
\begin{equation} \label{8.5}
f(\alpha) = (\pi\varepsilon)^{-\f12} \left[ e^{-(\alpha-\alpha_0)^2/ \varepsilon}
\right] \, \text{ implies } \,
Q(f)(\lambda) = \frac{\sin(2\alpha_0 \sqrt\lambda\, )}
{\sqrt\lambda}\, e^{-\varepsilon\lambda}.
\end{equation}
We use $f(\alpha, \alpha_0, \varepsilon)$ for the function $f$ in \eqref{8.5}.
For $\lambda \geq 0$, repeated integrations by parts show that
\begin{equation} \label{8.6}
|Q(f)(\lambda)| \leq C(1+\lambda^2)^{-1} \left[ \|f\|_1 +
\left\| \frac{d^3f}{d\alpha^3} \right\|_1 \right],
\end{equation}
where $\| \,\cdot\,\|_1$ represents the $L^1 ({\mathbb{R}})$-norm. Moreover,
essentially by repeating the calculation that led to \eqref{8.5}, one sees
that for $\lambda
\leq 0$,
\begin{equation} \label{8.7}
|Q(f)(\lambda)| \leq Ce^{\varepsilon |\lambda|} \bigl\| e^{+\alpha^2/ \varepsilon}
f\bigr\|_\infty.
\end{equation}
One then concludes the following result.
\begin{lemma} \label{p8.1} If $\int_{\mathbb{R}} (1+\lambda^2)^{-1} \, d\rho(\lambda) <
\infty$ {\rm(}always true{\rm!)} and $\int_{-\infty}^0
e^{-\varepsilon_0 \lambda} \, d\rho(\lambda) <\infty$ {\rm(}see Theorem~\ref{t6.6}
and the remark following its proof{\rm)}, then using \eqref{8.2},
$T(\,\cdot\,)$ can be extended to functions $f\in C^3({\mathbb{R}})$ that satisfy
$e^{\alpha^2/\varepsilon_0}f\in L^\infty ({\mathbb{R}})$ for some $\varepsilon_0 >0$ and
$\frac{d^3 f}{d\alpha^3}\in L^1 ({\mathbb{R}})$, and moreover,
\begin{equation}
|T(f)| \leq C \left[ \, \biggl\| \frac{d^3 f}{d\alpha^3}\biggr\|_1 +
\bigl\| e^{\alpha^2/ \varepsilon_0} f\bigr\|_\infty \right]
:= C ||| f |||_{\varepsilon_0}. \label{8.8}
\end{equation}
\end{lemma}
Next, fix $\alpha_0$ and $\varepsilon_0>0$ so that
$\int_{-\infty}^0 e^{-\varepsilon_0 \lambda} \,
d\rho(\lambda) < \infty$. If $0<\varepsilon < \varepsilon_0$,
$f(\alpha, \alpha_0, \varepsilon)$
satisfies $||| f|||_{\varepsilon_0} <\infty$ so we can define $T(f)$. Fix $g\in
C^\infty_0 ({\mathbb{R}})$ with $g:= 1$ on $(-2\alpha_0, 2\alpha_0)$. Then
$||| f(\,\cdot\,, \alpha_0, \varepsilon)(1-g)|||_{\varepsilon_0} \to 0$ as
$\varepsilon \downarrow 0$. So
\[
\lim_{\varepsilon\downarrow 0} T(f(\,\cdot\, , \alpha_0, \varepsilon))
= \lim_{\varepsilon \downarrow 0}
T(gf(\,\cdot\,, \alpha_0, \varepsilon)).
\]
For $gf$, we can use the expression \eqref{8.3}. $f$ is approximately
$\delta (\alpha - \alpha_0)$ so standard estimates show if $\alpha_0$ is a
point of Lebesgue continuity of $\tilde A(\alpha)$, then
\[
\int_{-\infty}^\infty d\alpha \, f(\alpha, \alpha_0, \varepsilon) g(\alpha)
\tilde A(\alpha) \arrow_{\varepsilon\downarrow 0} \tilde A (\alpha_0).
\]
Since $A-q$ is continuous, points of Lebesgue continuity of $A$ exactly are
points of Lebesgue continuity of $V$. Thus, one obtains the following
theorem.
\begin{theorem} \label{t8.2} Suppose either $b<\infty$ and $V\in L^1 ([0,b])$
or $b=\infty$, and then either $V\in L^1 ([0,\infty))$ or
$V\in L^1 ([0,a])$ for all $a>0$ and
\[
V(x) \geq -Cx^2, \quad x\geq R
\]
for some $R>0$, $C>0$. Let $\alpha_0 \in (0,b)$ and be a point of Lebesgue
continuity of $V$. Then
\begin{equation} \label{8.9}
A(\alpha_0) = - 2\lim_{\varepsilon\downarrow 0} \int_{\mathbb{R}} d\rho(\lambda) \,
e^{-\varepsilon\lambda}
\frac{\sin(2\alpha_0 \sqrt\lambda \,)}{\sqrt\lambda}.
\end{equation}
\end{theorem}
\medskip
Finally, we specialize \eqref{8.9} to the scattering-theoretic setting.
Assuming $V\in L^1 ([0,\infty); (1+x)\, dx)$, the corresponding Jost
solution $f(z,x)$ is defined by
\begin{equation} \label{8.11}
f(z,x) = e^{i\sqrt{z}\,x} - \int_x^\infty dx' \, \frac{\sin(\sqrt{z}\, (x-x'))}
{\sqrt z}\, V(x') f(z,x'), \quad \Im(\sqrt z\, )\geq 0,
\end{equation}
and the corresponding Jost function, $F(\sqrt z\,)$, and scattering matrix,
$S(\lambda)$, $\lambda \geq 0$, then read
\begin{align}
F(\sqrt z\, ) &= f(z,0_+), \label{8.12} \\
S(\lambda) &= \overline{F(\sqrt\lambda\, )}/F(\sqrt\lambda\,), \quad
\lambda \geq 0. \label{8.13}
\end{align}
The spectrum of the Schr\"odinger operator $H$ in $L^2 ([0,\infty))$
associated with the differential expression $-\frac{d^2}{dx^2} + V(x)$ and a
Dirichlet boundary condition at $x=0_+$ (cf.\ \eqref{8.1.5} for precise details) is simple and of the type
\[
\sigma(H) = \{ -\kappa^2_j <0\}_{j\in J} \cup [0,\infty).
\]
Here $J$ is a finite (possibly empty) index set, $\kappa_j >0$, $j\in J$,
and the essential spectrum is purely absolutely continuous. The
corresponding spectral measure explicitly reads
\begin{equation} \label{8.14}
d\rho(\lambda) = \begin{cases}
\pi^{-1} |F(\sqrt\lambda\,)|^{-2}\sqrt\lambda\, d\lambda,
&\lambda \geq 0, \\
\sum_{j\in J} c_j \delta (\lambda + \kappa^2_j)\, d\lambda, &\lambda <0,
\end{cases}
\end{equation}
where
\begin{equation} \label{8.15}
c_j = \|\varphi (-\kappa^2_j,\,\cdot\,)\|^{-2}_2, \quad j\in J
\end{equation}
are the norming constants associated with the eigenvalues $\lambda_j =
-\kappa^2_j <0$. Here the regular solution $\varphi(z,x)$ of
$-\psi''(z,x)+[V(x)-z]\psi(z,x)=0$ (defined by
$\varphi(z,0_+)=0$, $\varphi'(z,0_+)=1$) and $f(z,x)$ in \eqref{8.11} are
linearly dependent precisely for $z=-\kappa^2_j$, $j\in J$.
Since
\[
|F(\sqrt\lambda\, )| = \prod_{j\in J}
\left( 1+\frac{\kappa^2_j}{\lambda}\right)
\exp\left(\frac{1}{\pi}\, P\int_0^\infty \frac{d\lambda' \, \delta(\lambda')}
{\lambda - \lambda'}\right), \quad \lambda \geq 0,
\]
where $P\int_0^\infty$ denotes the principal value symbol and
$\delta(\lambda)$
the corresponding scattering phase shift, that is,
$S(\lambda) =\exp(2i \delta
(\lambda))$, $\delta(\lambda)\arrow\limits_{\lambda\uparrow\infty} 0$, the
scattering data
\[
\{-\kappa^2_j, c_j\}_{j\in J} \cup \{S(\lambda)\}_{\lambda \geq 0}
\]
uniquely determine the spectral measure \eqref{8.14} and hence $A(\alpha)$.
Inserting \eqref{8.14} into \eqref{8.9} then yields the following expression for
$A(\alpha)$ in terms of scattering data.
\begin{theorem} \label{t8new} Suppose that $V\in L^1 ([0,\infty);
(1+x)dx)$. Then
\begin{equation} \label{8.16}
\begin{split}
A(\alpha) &= -2\sum_{j\in J} c_j \kappa^{-1}_j \sinh(2\alpha\kappa_j) \\
&\quad -2\pi^{-1} \lim_{\varepsilon\downarrow 0}\int_0^\infty
d\lambda \,
e^{-\varepsilon\lambda} |F(\sqrt\lambda)|^{-2} \sin(2\alpha \sqrt\lambda\,)
\end{split}
\end{equation}
at points $\alpha \geq 0$ of Lebesgue continuity of $V$.
\end{theorem}
\begin{remark} In great generality $|F(k)|\to 1$ as $k\to\infty$,
so one cannot take the limit in $\varepsilon$ inside the integral in \eqref{8.16}.
In general, though, one can can replace $|F(\sqrt\lambda\,)|^{-2}$ by
$(|F(\sqrt\lambda\,)|^{-2} -1) \equiv X(\lambda)$ and ask if one can take
a limit there. As long as $V$ is $C^2 ((0,\infty))$ with $V'' \in L^1
([0,\infty))$, it is not hard to see that as $\lambda\to\infty$
\[
X(\lambda) = -\frac{V(0)}{2\lambda} + O(\lambda^{-2}).
\]
Thus, if $V(0) = 0$, then
\begin{equation} \label{8.17}
\begin{split}
A(\alpha) &= -2 \sum_{j\in J} c_j \kappa^{-1}_j \sinh (2\alpha \kappa_j)
\\ &\quad -2\pi^{-1} \int_0^\infty d\lambda \, (|F(\sqrt\lambda\,)|^{-2}-1)
\sin(2\alpha \sqrt\lambda\,).
\end{split}
\end{equation}
The integral in \eqref{8.17} is only conditionally convergent if
$V(0)\neq 0$.
\end{remark}
We note that in the present case, where $V\in L^1 ([0,\infty);
(1+x)\,dx)$, the representation \eqref{1.15} of the $m$-function in terms
of the $A(\alpha)$-amplitude was considered in a paper by Ramm \cite{Ra87}
(see also \cite[p.\ 288--291]{Ra92}).
We add a few more remarks in the scattering-theoretic setting. Assuming
$V\in L^1 ([0,\infty); (1+x)\, dx)$, one sees that
\begin{equation} \label{"9.17"}
|F(k)| \eqlim_{k\uparrow \infty} 1 + o(k^{-1})
\end{equation}
(cf.\ \cite[eq.\ II.4.13]{CS89} and apply the Riemann-Lebesgue lemma;
actually, one only needs $V\in L^1 ([0,\infty))$ for the asymptotic
results on $F(k)$ as
$k\uparrow\infty$ but we will ignore this refinement in the following). A
comparison of \eqref{"9.17"} and \eqref{8.16} then clearly demonstrates the
necessity of an Abelian limit in \eqref{8.16}. Even replacing $d\rho$ in
\eqref{8.9} by $d\sigma = d\rho - d\rho^{(0)}$, that
is, effectively replacing $|F(\sqrt\lambda\,)|^{-2}$ by
$[|F(\sqrt\lambda\,)|^{-2} -1]$ in \eqref{8.16}, still does not
necessarily produce an absolutely convergent integral in \eqref{8.16}.
The latter situation changes upon increasing the smoothness properties of
$V$ since, for example, assuming $V\in L^1 ([0,\infty); (1+x)\, dx)$,
$V'\in L^1 ([0,\infty))$, yields
\[
|F(k)|^{-2} - 1 \eqlim_{k\uparrow \infty} O(k^{-2})
\]
as detailed high-energy considerations (cf.~\cite{GPT80}) reveal. Indeed,
if $V'' \in L^1 ([0,\infty))$, then the integral one gets is absolutely
convergent if and only if
$V(0)=0$.
\medskip
As a final issue related to the representation \eqref{5.1}, we discuss the
issue of bounds on $A$ when $|V(x)| \leq Cx^2$. One has two general
bounds on $A$: the estimate of \cite{Si99} (see \eqref{1.14}),
\begin{equation} \label{9.16}
|A(\alpha) - V(\alpha)| \leq \left[ \int_0^\alpha dy \, |V(y)| \right]^2
\exp \left[ \alpha\int_0^\alpha dy \, |V(y)| \right] ,
\end{equation}
and the estimate in Theorem~\ref{t10.2},
\begin{equation} \label{9.17}
|A(\alpha)| \leq \frac{\gamma(\alpha)}{\alpha} \,
I_1 (2\alpha \gamma(\alpha)),
\end{equation}
where $|\gamma(\alpha)| = \sup_{0\leq x\leq\alpha} |V(x)|^{1/2}$ and
$I_1 (\,\cdot\,)$ is the modified Bessel function of order one (cf., e.g.,
\cite{AS}, Ch.~9). Since (\cite{AS}, p.~375)
\begin{equation} \label{9.18}
0\leq I_1 (x) \leq e^x , \quad x\geq 0,
\end{equation}
one concludes that
\begin{equation} \label{9.19}
|A(\alpha)| \leq \sqrt C\, e^{2\sqrt C\, \alpha^2}
\end{equation}
if $|V(x)|\leq Cx^2$.
We continue with a discussion of the case of constant $V$:
\begin{example} If $b=\infty$ and $V(x)=V_0$, $x\geq 0$, then if $V_0 >0$,
\begin{equation} \label{10.1}
A(\alpha) = \frac{V^{1/2}_0}{\alpha}\, J_1 (2\alpha V^{1/2}_0),
\end{equation}
where $J_1 (\,\cdot\,)$ is the Bessel function of order one {\rm(}cf., e.g.,
\cite{AS}, Ch.~9{\rm )}; and if $V_0 < 0$,
\begin{equation} \label{10.2}
A(\alpha) = \frac{(-V_0)^{1/2}}{\alpha}\, I_1 (2\alpha (-V_0)^{1/2}),
\end{equation}
with $I_1 (\,\cdot\,) $ the corresponding modified Bessel function.
\end{example}
This example is important because of the following monotonicity
property:
\begin{theorem} \label{t10.2} Let $|V_1(x)| \leq - V_2(x)$ on $[0,a]$ with
$a\leq \min(b_1, b_2)$. Then,
$$
|A_1 (\alpha)| \leq -A_2 (\alpha) \, \text{ on } \, [0,a].
$$
In particular, for any $V$ satisfying $\sup_{0\leq x \leq\alpha} |V(x)| <
\infty$, one obtains
\begin{equation} \label{10.5a}
|A(\alpha)| \leq \frac{\gamma(\alpha)}{\alpha}\, I_1 (2\alpha \gamma(\alpha)),
\end{equation}
where
\begin{equation}\label{10.5b}
\gamma(\alpha) = \sup_{0\leq x\leq\alpha} (|V(x)|^{1/2}).
\end{equation}
In particular, \eqref{9.18} implies
\begin{equation} \label{10.6}
|A(\alpha)| \leq \alpha^{-1} \gamma(\alpha) e^{2\alpha\gamma(\alpha)},
\end{equation}
and if $V$ is bounded,
\begin{equation} \label{10.7}
|A(\alpha)| \leq \alpha^{-1} \|V\|^{1/2}_\infty
\exp(2\alpha \|V\|^{1/2}_\infty).
\end{equation}
\end{theorem}
For $\alpha$ small, \eqref{10.6} is a poor estimate and one should use
\eqref{9.16} which implies that $|A(\alpha) \leq \|V\|_\infty + \alpha^2
\|V\|^2_\infty e^{\alpha^2\|V\|_\infty}$.
This lets one prove the following result:
\begin{theorem} \label{t10.3} Let $h=\infty$ and $V\in L^\infty
([0,\infty))$. Suppose
$\kappa^2 > \|V\|_\infty$. Then
\begin{equation} \label{10.9}
m(-\kappa^2) = -\kappa - \int_0^\infty d\alpha \, A(\alpha) e^{-2\alpha\kappa}
\end{equation}
{\rm (}with an absolutely convergent integral and no error term{\rm )}.
\end{theorem}
\begin{remark}
We recall (cf.\ \eqref{1.15}) that the representation \eqref{10.9} also
holds with $A\in L^1([0,a])$ for all $a>0$ and as an absolutely convergent
integral for $\Re(\kappa) > \|V\|_1/2$ if $V\in L^1([0,\infty))$. This
fact will be used below.
\end{remark}
The case of Bargmann potentials has
been worked out in
\cite[Sect.\ 11]{GS00a} and explicit formulas for the $A$-function have
been obtained.
We end this survey of \cite{GS00a} and \cite{Si99} by recalling the major thrust
of \cite{Si99} -- the connection between $A$ and the inverse spectral
theory. Namely, there is an $A(\alpha, x)$ function associated to $m(z,x)$ by
\begin{equation} \label{1.20}
m(-\kappa^2, x) = -\kappa - \int_0^a d\alpha \,
A(\alpha, x) e^{-2\alpha\kappa} +
\tilde O(e^{-2\alpha\kappa})
\end{equation}
for $a<b -x$. This, of course, follows from Theorem \ref{t1} by
translating the origin. The point is that $A$ satisfies the simple
differential equation in distributional sense
\begin{equation} \label{1.21}
\frac{\partial A}{\partial x}\, (\alpha, x)
= \frac{\partial A}{\partial\alpha} \,
(\alpha, x) + \int_0^a d\beta \, A(\alpha-\beta, x) A(\beta,x).
\end{equation}
This is proved in \cite{Si99} for $V\in L^1 ([0, a])$ (and some other
$V$'s) and so holds in the generality of \cite{GS00a} since Theorem \ref{t3}
implies $A(\alpha,x)$ for $\alpha+x \leq a$ is only a function of $V(y)$
for $y\in[0,a]$.
Moreover, by \eqref{1.14}, one has
\begin{equation} \label{1.22}
\lim_{\alpha\downarrow 0} |A(\alpha, x) - V(\alpha + x)| =0
\end{equation}
uniformly in $x$ on compact subsets of the real line, so by the uniqueness
theorem for solutions of \eqref{1.21} (proved in \cite{Si99}), $A$ on
$[0,a]$ determines $V$ on $[0,a]$.
In the limit circle case, there is an additional issue to discuss. Namely,
that $m(z, x=0)$ determines the boundary condition at $\infty$. This is
because, as we just discussed, $m$ determines $A$ which determines $V$ on
$[0,\infty)$. $m(z, 0_+)$ and $V$ determine $m(z,x)$ by the Riccati
equation. Once we know $m$, we can recover $u(i, x) = \exp
\big(\int_0^x m(i, y)\, dy\big)$, and so the particular solution that
defined the boundary condition at $\infty$.
Thus, the inverse spectral theory aspects of the framework easily extend to
the general case of potentials considered in \cite{GS00a}.
To turn this into an inverse spectral approach alternative to and fully equivalent to that of Gel'fand and Levitan, one needs to settle necessary and sufficient conditions for solvability of the differential equation \eqref{1.21} in terms of an initial condition
$A(\alpha, 0_+)=A_0(\alpha)$, that is, in terms of properties of $A_0$. This final step was accomplished by Remling \cite{Re03} and we briefly describe its major elements next.
Remling's first result is of local nature and determines a necessary and sufficient condition on $A$ to be the $A$-function of a potential $V$. Assuming $V\in L^1([0,b])$ for all $b>0$, he introduces the set
\begin{equation}
{\mathcal A}_b=\{A\in L^1([0,b])\,|\, \text{$A$ real-valued}, \, I+{\mathcal K}_A >0 \}, \label{1.23}
\end{equation}
where
\begin{align*}
& ({\mathcal K}_A f)(\alpha)=\int_0^b d\beta \, K(\alpha,\beta) f(\beta), \quad
\alpha \in [0,b], \; f\in L^2([0,b]), \\
& K(\alpha, \beta)=[\phi(\alpha-\beta)-\phi(\alpha+\beta)]/2, \quad
\phi(\alpha)=\int_0^{|\alpha|/2} d\gamma \, A(\gamma), \; \alpha, \beta \in [0,b].
\end{align*}
Based on his reformulation of the Gel'fand--Levitan approach in terms of de Branges spaces in \cite{Re02}, Remling obtained the following characterization of $A$-functions:
\begin{theorem}
${\mathcal A}_b$ is precisely the set of $A$-functions in
$$
m(-\kappa^2) = -\kappa - \int_0^a d\alpha \, A(\alpha) e^{-2\alpha\kappa} +
\tilde O(e^{-2\alpha\kappa}) \, \text{ for all $a<b$.}
$$
Equivalently, given $A_0\in L^1([0,b])$, there exists a potential
$V\in L^1([0,b])$ such that $A_0$ is the $A$-function of $V$ if and only if
$A_0\in{\mathcal A}_b$.
\end{theorem}
(We recall that all potentials $V$ in this survey are assumed to be real-valued.)
As a second result, Remling also proved in \cite{Re03} that the positivity condition in \eqref{1.23} is necessary and sufficient to solve \eqref{1.21} on
$\Delta_b=\{(\alpha,x)\in{\mathbb{R}}^2\,|\, \alpha\in [0,b-x], \, x\in [0,b] \}$ given an initial condition $A(\cdot,0_+)=A_0 \in L^1([0,b])$. The potential $V$ can then be read off from
\begin{equation}
V(x)=A(0_+,x) \, \text{ for $x\in [0,b]$.} \label{1.24}
\end{equation}
Necessity of this positivity condition had been established independently by Keel and Simon (unpublished). To make this precise, it pays to slightly rewrite
\eqref{1.21} as follows: Let
\begin{equation}
B(\alpha,x)=A(\alpha-x,x)-A_0(\alpha), \quad (\alpha, x) \in\widetilde \Delta_b, \label{1.25}
\end{equation}
where
$$
\widetilde \Delta_b = \{(\alpha,x)\in{\mathbb{R}}^2\,|\, 0\leq x\leq \alpha \leq b\}.
$$
Then \eqref{1.21} together with the initial condition $A(\cdot,0_+)=A_0 \in L^1([0,b])$, becomes
\begin{align}
& B(\alpha,x)=\int_0^x dy \int_0^{\alpha -y} d\beta \, [B(y+\beta,y)+A_0(y+\beta)]
[B(\alpha-\beta,y)+A_0(\alpha-\beta)] \notag \\
& B(\alpha,0_+)=0, \quad (\alpha,x)\in \widetilde \Delta_b. \label{1.26}
\end{align}
If $A$ is actually the $A$-function of a potential, then $B\in C(\widetilde \Delta_b)$ by \cite[Theorem\ 2.1]{Si99}. Remling \cite{Re03} then proves the following result:
\begin{theorem}
Suppose $A_0\in L^1([0,b])$. Then \eqref{1.26} has a solution $B\in C(\widetilde \Delta_b)$ if and only if $A_0\in {\mathcal A}_b$.
\end{theorem}
This brings Simon's inverse approach to full circle and one can envision the following two scenarios. First, Simon's inverse $A$-function approach, as complemented by Remling \cite{Re03}:
\vspace{5pt}
\fbox{\begin{minipage}{330pt}
\begin{align}
& A_0\in {\mathcal A}_b
\xrightarrow{\text{by \eqref{1.26}}} B(\alpha,x), \; (\alpha,x)\in\widetilde\Delta_b
\xrightarrow{\text{by \eqref{1.25}}} A(\alpha,x), \; (\alpha,x)\in\Delta_b \notag \\
& \quad \xrightarrow{\text{by \eqref{1.24}}} V=A(0_+,\cdot)\in L^1([0,b]). \label{1.27}
\end{align}
\end{minipage}}
\vspace{10pt}
Second, denote by ${\mathcal R}$ the set of spectral functions $\rho$ associated with self-adjoint half-line Schr\"odinger operators with a Dirichlet boundary condition at $x=0$ and a self-adjoint boundary condition \eqref{1.5} at infinity (if any, i.e., if \eqref{1.1} is in the limit circle case at $\infty$). For characterizations of ${\mathcal R}$ we refer, for instance, to
\cite{LG64}, \cite[Ch.\ 2]{Le87}, \cite[Ch.\ 2]{Ma86}. Then combining \eqref{1.27} with \eqref{5.16} yields Simon's inverse spectral approach as an alternative to that by Gel'fand and Levitan:
\vspace{5pt}
\fbox{\begin{minipage}{330pt}
\begin{align}
\begin{split}
& \rho \in {\mathcal R} \xrightarrow{\text{by \eqref{5.16}}} A_0\in {\mathcal A}_b \, \text{ for all $b>0$ } \\
& \quad \xrightarrow{\text{by \eqref{1.26}}} B(\alpha,x), \; (\alpha,x)\in\widetilde\Delta_b
\, \text{ for all $b>0$} \\
& \qquad \xrightarrow{\text{by \eqref{1.25}}} A(\alpha,x), \; (\alpha,x)\in\Delta_b
\, \text{ for all $b>0$} \\
& \qquad \quad \xrightarrow{\text{by \eqref{1.24}}} V=A(0_+,\cdot)\in L^1([0,b])
\, \text{ for all $b>0$}.
\end{split}
\end{align}
\end{minipage}}
\vspace{10pt}
\medskip
{\it More recent references:} Local solvability and a necessary condition for global solvability of the $A$-equation \eqref{1.21} were recently discussed by Zhang \cite{Zh03}, \cite{Zh06}. Connections between the $A$-amplitude and the scattering transform for Schr\"odinger operators on the real line have been discussed by Hitrik \cite{Hi01}.
\smallskip
\begin{center}
$\bigstar$ \qquad $\bigstar$ \qquad $\bigstar$
\end{center}
\smallskip
Next we briefly quote the main results by Ramm and Simon \cite{RS00}. The
primary goal in this paper was to study $A$ as an interesting object in
its own right and, in particular, using ideas implicit in Ramm
\cite{Ra87} to obtain detailed information on the behavior of $A(\alpha)$
as $\alpha\to\infty$ when $V$ decays sufficiently fast as $x\to\infty$.
Indeed, for potentials decaying rapidly enough, Ramm \cite{Ra87} stated
the representation \eqref{10.9} (actually, \eqref{1.15}), but no proof was
given (nor was there any connection of the function $A$ to the inverse
problem for $V$). In \cite{Ra87} the inverse problem of finding the
potential from the knowledge of the $m$-function has been solved for
short-range potentials. A more detailed discussion of the result in
\cite{Ra87} can be found in \cite{Ra00}, \cite{Ra02}.
Throughout \cite{RS00} it is assumed that
\begin{equation} \label{8.1.4}
\int_0^\infty (1+x) \, dx \, |V(x)| < \infty
\end{equation}
and the Dirichlet-type Schr\"odinger operator $H$ in
$L^2([0,\infty))$ defined by
\begin{align}
\begin{split}
&H f=-f''+Vf, \quad
f\in \text{\rm{dom}}(H)=\{u\in L^2([0,\infty))\,|\, u, u' \in AC_{\text{\rm{loc}}}([0,b]) \\
& \hspace*{3.4cm}
\text{for all $b>0$}; \, u(0_+)=0; \, (-g''+Vg)\in L^2([0,\infty))\} \label{8.1.5}
\end{split}
\end{align}
is considered.
More generally, for $n\in{\mathbb{N}}_0$, $B\leq 0$ and $\ell
\geq 0$, the space $C^{B,\ell}_n$ of all functions $q$ with $n-1$
classical derivatives and $q^{(n)}\in L^1 ([0,\infty))$ so that
$$
\int_0^\infty (1 +x)^\ell\, e^{-Bx} \, dx \, |q^{(j)}(x)| < \infty
$$
for $j=0,1,\dots, n$. Thus, \eqref{8.1.4} says $V\in C^{B=0,
\ell=1}_{n=0}$.
Under condition \eqref{8.1.4}, general principles (see, e.g.,
\cite[Ch.\ 3]{Ma86}) imply that for all $\kappa\in{\mathbb{C}}$ with $\Re(\kappa)
\geq 0$, there is a unique solution $F(\kappa,x)$ of $-f'' +
Vf=-\kappa^2 f$ normalized so that $F(\kappa,x) = e^{-\kappa x} (1
+ o(1))$ as $x\to
\infty$. We set $F(\kappa):= F(\kappa,0_+)$. Except
for the change of variables $\kappa = -ik$, $F(\kappa,x)$ and $F(\kappa)$
are the standard Jost solution and Jost function. Both $F(\kappa,x)$ and
$F(\kappa)$ are analytic with respect to $\kappa$ in $\{\kappa\in{\mathbb{C}}\,|\,
\Re(\kappa) > 0\}$. If $V\in C^{B,\ell}_n$ for any $n,\ell$ and $B<0$,
then $F(\kappa,x)$ and $F(\kappa)$ have analytic continuations into the
region $\Re (\kappa) > B/2$.
The following is easy to see and well known (cf.\ \cite[Ch.\ 3]{Ma86}):
\begin{enumerate}
\item[(1)] The zeros of $F$ in $\{\kappa\in {\mathbb{C}}\,|\,\Re(\kappa) >0\}$
occur precisely at those points $\kappa_j$ with $-\kappa^2_j$ an
eigenvalue of the operator $H$ and each such zero is simple.
\item[(2)] $F$ has no zeros in $\{\kappa\in{\mathbb{C}}\,|\,\Re(\kappa)
=0,\,\kappa\neq 0\}$.
\item[(3)] If $F(0)=0$ and $V\in C^{B=0, \ell=2}_{n=0}$, then $F$ is $C^1$
and $F'(0)\neq 0$. If $F(0)=0$, we say that $H$ has a {\it zero energy
resonance}.
\end{enumerate}
If $F$ can be analytically continued to $\{\kappa\in{\mathbb{C}}\,|\,\Re(\kappa)
>B/2\}$ for $B<0$, then zeros of $F$ in $\{\kappa\in{\mathbb{C}}\,|\,\Re(\kappa)
<0\}$ are called {\it resonances} of $H$. They occur in complex
conjugate pairs (since $F$ is real on the real axis). If
$F'(\kappa_0)\neq 0$ at a zero
$\kappa_0$, we say that
$\kappa_0$ is a simple resonance. Resonances need not be simple if $\Re
(\kappa_0) <0$ although they are generically simple.
The result stated in \cite{Ra87} can be phrased as follows:
\begin{theorem} \label{t8.8.1} Suppose that $V$ satisfies \eqref{8.1.4}
{\rm(}i.e., it lies in
$C^{B=0, \ell=1}_{n=0}${\rm)} and that $H$ does not have a zero energy
resonance. Let $\{-\kappa^2_j\}^J_{j=1}$ be the negative eigenvalues of
$H$ with $\kappa_j >0$. Then
\begin{equation} \label{8.1.6}
A(\alpha) = \sum_{j=1}^J B_j \, e^{2\alpha\kappa_j} + g(\alpha),
\end{equation}
where $g\in L^1 ([0,\infty))$. In particular, if $H$ has no eigenvalues
and no zero energy resonance {\rm(}e.g., if $V\geq 0${\rm)}, then $A\in
L^1([0,\infty))$.
\end{theorem}
\begin{remark} $(i)$ The result stated in \cite{Ra87} assumes
implicitly that there is no zero energy resonance.
Details can be found in \cite{Ra00}. \\
$(ii)$ If $A\in L^1([0,\infty))$, then the representation \eqref{10.9}
(resp., \eqref{1.15}) can be analytically continued to the entire region
$\Re(\kappa)
\geq 0$. \\
$(iii)$ If $u_j$ is the eigenfunction of $H$ corresponding to the
eigenvalue $-\kappa^2_j$, normalized by $\|u_j\|_2 = 1$, then
$$
B_j = -\frac{|u'_j (0_+)|^2}{\kappa_j}\, .
$$
This follows from \eqref{8.16}
and the fact that
$$
d\rho(\lambda)\restriction (-\infty, 0) = \sum_{j=1}^J
|u'_j(0_+)|^2 \delta (\lambda + \kappa^2_j)\, d\lambda.
$$
\end{remark}
To handle zero energy resonances of $H$,
one needs an extra two powers of decay
(just as \eqref{1.4} says more or less that $|V(x)|$ is bounded by
$O(x^{-2-\varepsilon})$, the condition in the next theorem says that
$|V(x)|$ is more or less $O(x^{-4-\varepsilon})$):
\begin{theorem} \label{t8.8.2} Let $V\in C^{B=0, \ell=3}_{n=0}$. Suppose
that $H$ has a zero energy
resonance and negative eigenvalues at $\{-\kappa^2_j\}^J_{j=1}$ with
$\kappa_j >0$. Then
\begin{equation} \label{8.1.7b}
A(\alpha) = B_0 + \sum_{j=1}^J B_j \, e^{2\alpha\kappa_j} + g(\alpha),
\end{equation}
where $g\in L^1 ([0,\infty))$.
\end{theorem}
These results are special cases of the following theorem:
\begin{theorem} \label{t8.8.3} Let $V\in C^{B=0, \ell}_n$ where $\ell \geq
1$ and, if $H$ has a zero energy resonance, then $\ell\geq 3$. Then
\eqref{8.1.6} {\rm(}resp., \eqref{8.1.7b} if there is a zero energy
resonance{\rm)} holds, where
$g\in C^{B=0, \ell-1}_n$ {\rm(}resp., $C^{B=0, \ell-3}_n${\rm)}.
\end{theorem}
Finally, for $B<0$, the following result was proved in \cite{RS00}:
\begin{theorem} \label{t8.8.4} Let $V\in C^{B, \ell=0}_n$ with $B<0$. Let
$\tilde B\in (B, 0)$ and let $\{-\kappa^2_j\}_{j=1}^J$ with $\kappa_j >0$
be the negative eigenvalues of $H$, $\{\lambda_j\}_{j=1}^M$ with
$\lambda_j\leq 0$ the real resonances {\rm(}a.k.a.~anti-bound
states{\rm)} of $H$, and $\{\mu_j
\pm i\nu_j\}_{j=1}^N$ the complex resonances of $H$ with $\tilde B \leq
\mu_j <0$ and $\nu_j > 0$. Suppose each resonance is simple. Then for
suitable
$\{B_j\}_{j=1}^J$, $\{C_j\}_{j=1}^M$, $\{D_j\}_{j=1}^N$,
$\{\theta_j\}_{j=1}^N$, one obtains
\[
A(\alpha) = \sum_{j=1}^J B_j \, e^{2\alpha\kappa_j} + \sum_{j=1}^M
C_j \, e^{2\alpha\lambda_j} + \sum_{j=1}^N D_j \, e^{2\mu_j\alpha}
\cos(2\nu_j\alpha + \theta_j) + \tilde g(\alpha),
\]
where $\tilde g \in C^{\tilde B, \ell=0}_n$. In particular, if
$H$ has no negative eigenvalues, the rate of decay of $A$ is
determined by the resonance with the least negative value of $\lambda$ or
$\mu$.
\end{theorem}
\smallskip
\begin{center}
$\bigstar$ \qquad $\bigstar$ \qquad $\bigstar$
\end{center}
\smallskip
We conclude this section with a brief look at the principal results in \cite{GS00b}.
Let $H_j = -\frac{d^2}{dx^2} + V_j$, $V_j\in L^1 ([0,b])$ for
all $b>0$, $V_j$ real-valued, $j=1,2$, be two self-adjoint operators in
$L^2 ([0,\infty))$ with a Dirichlet boundary condition at $x=0_+$. Let $m_j(z)$,
$z\in{\mathbb{C}}\backslash{\mathbb{R}}$ be the Weyl--Titchmarsh $m$-functions associated with
$H_j$, $j=1,2$. The main purpose of \cite{GS00b} was to provide a short proof of the following local uniqueness theorem in the spectral theory of one-dimensional
Schr\"odinger operators, originally obtained by Simon \cite{Si99}, but under slightly more general assumptions than in \cite{Si99}.
We summarize the principal results of \cite{GS00b} as follows:
\begin{theorem}\label{t9.1.1} $(i)$ Let $a>0$, $0<\varepsilon<\pi/2$ and
suppose that
$$
|m_1 (z) - m_2(z)| \eqlim_{|z|\to\infty}
O(e^{-2\Im (z^{1/2})a})
$$
along the ray $\arg(z) = \pi-\varepsilon$. Then
$$
V_1(x) = V_2 (x) \, \text{ for a.e.\ } \, x \in [0,a].
$$
$(ii)$ Conversely, let $\arg(z)\in (\varepsilon, \pi -\varepsilon)$
for some $0<\varepsilon<\pi$ and suppose $a>0$. If
$$
V_1(x) = V_2(x) \, \text{ for a.e.\ } \, x \in [0,a],
$$
then
\begin{equation} \label{9.2.13}
|m_1 (z) - m_2(z)| \eqlim_{|z|\to\infty}
O(e^{-2\Im (z^{1/2})a}).
\end{equation}
$(iii)$ In addition, suppose that $H_j$, $j=1,2$, are bounded
from below. Then \eqref{9.2.13} extends to all $\arg(z) \in (\varepsilon,\pi]$.
\end{theorem}
\begin{corollary}\label{c9.2.6} Let $0<\varepsilon <\pi/2$ and suppose
that for all $a>0$,
$$
|m_1(z) - m_2(z)| \eqlim_{|z|\to\infty}
O(e^{-2\Im (z^{1/2})a})
$$
along the ray $\arg(z) = \pi -\varepsilon$. Then
$$
V_1(x) = V_2(x) \, \text{ for a.e.\ } \, x\in [0,\infty).
$$
\end{corollary}
Theorem \ref{t9.1.1} and Corollary \ref{c9.2.6} follow by combining some of the Riccati equation methods in \cite{GS00a} with properties of transformation operators (cf.\
\cite[Sect.\ 3.1]{Ma86}) and a uniqueness theorem for finite Laplace transforms
\cite[Lemma\ A.2.1]{Si99}.
In particular, Corollary \ref{c9.2.6} represents a considerable strengthening of the original Borg--Marchenko uniqueness result \cite{Bo52}, \cite{Ma50}, \cite{Ma73}:
\begin{theorem} \label{t9.1.2}
Suppose
$$
m_1(z) = m_2(z), \quad z\in{\mathbb{C}}\backslash{\mathbb{R}},
$$
then
$$
V_1(x) = V_2(x) \, \text{ for a.e.\ } \, x\in [0,\infty).
$$
\end{theorem}
\begin{remark} \label{r9.2.7}
$(i)$ Marchenko \cite{Ma50} first published Theorem \ref{t9.1.2} in 1950. His extensive treatise on spectral theory of one-dimensional Schr\"odinger operators
\cite{Ma73}, repeating the proof of his uniqueness theorem, then appeared in 1952, which also marked the appearance of Borg's proof of the uniqueness theorem \cite{Bo52} (apparently, based on his lecture at the 11th Scandinavian Congress of Mathematicians held at Trondheim, Norway, in 1949). \\
We emphasize that Borg and Marchenko also treat the general case of non-Dirichlet
boundary conditions at $x=0_+$ (see also item $(vi)$ below). Moreover, Marchenko simultaneously discussed the half-line and finite interval case (cf.\ item $(vii)$ below). \\
$(ii)$ As pointed out by Levitan \cite{Le87} in his Notes to Chapter\ 2, Borg and
Marchenko were actually preceded by Tikhonov \cite{Ti49} in 1949, who
proved a special case of Theorem~\ref{t9.1.2} in connection with the string equation (and hence under certain additional hypotheses on $V_j$). \\
$(iii)$ Since Weyl--Titchmarsh functions $m$ are uniquely related to the spectral measure $d\rho$ of $H$ by the standard Herglotz representation theorem,
\eqref{1.8b}, Theorem~\ref{t9.1.2} is equivalent to the following statement: Denote by
$d\rho_j$ the spectral measures of $H_j$, $j=1,2$. Then
$$
d\rho_1 = d\rho_2 \, \text{ implies } \, V_1 = V_2 \, \text{ a.e.~on } \, [0,\infty).
$$
In fact, Marchenko took the spectral measures $d\rho_j$ as his point of
departure while Borg focused on the Weyl--Titchmarsh functions $m_j$. \\
$(iv)$ The Borg--Marchenko uniqueness result, Theorem~\ref{t9.1.2} (but not the strengthened version, Corollary \ref{c9.2.6}), under the additional condition of
short-range potentials $V_j$ satisfying $V_j\in L^1 ([0,\infty); (1+x)\, dx)$, $j=1,2$, can also be proved using Property~C, a device used by Ramm \cite{Ra99}, \cite{Ra00} in a variety of uniqueness results. \\
$(v)$ The ray $\arg(z) = \pi -\varepsilon$, $0<\varepsilon < \pi/2$ chosen in Theorem \ref{t9.1.1}\,$(i)$ and Corollary \ref{c9.2.6} is of no particular importance. A limit taken along any non-self-intersecting curve ${\mathcal C}$ going to infinity in the sector
$\arg(z)\in ((\pi/2)+\varepsilon, \pi -\varepsilon)$ will do since we can apply the
Phragm\'en--Lindel\"of principle (\cite[Part~III, Sect.~6.5]{PS72}) to the region enclosed by ${\mathcal C}$ and its complex conjugate $\overline{\mathcal C}$. \\
$(vi)$ For simplicity of exposition, we only discussed the Dirichlet boundary condition
$$
u(0_+)=0
$$
in the Schr\"odinger operator $H$. Everything extends to the the general boundary condition
$$
u'(0_+) + h u(0_+) = 0, \quad h\in{\mathbb{R}},
$$
and we refer to \cite[Remark\ 2.9]{GS00b} for details. \\
$(vii)$ Similarly, the case of a finite interval problem on
$[0,b]$, $b\in (0,\infty)$, instead of the half-line $[0,\infty)$ in Theorem~\ref{t9.1.1}\,$(i)$, with $0<a<b$, and a self-adjoint boundary condition at $x=b_-$ of
the type
$$
u'(b_-) + h_b u(b_-) = 0, \quad h_b \in {\mathbb{R}},
$$
can be discussed (cf.\ \cite[Remark\ 2.10]{GS00b}).
\end{remark}
While we have separately described a few extensions in
Remarks \ref{r9.2.7}\,$(v)$--$(vii)$, it is clear that they can all be combined at once.
Without going into further details, we also mention that \cite{GS00b} contains the analog of the local Borg--Marchenko uniqueness result, Theorem \ref{t9.1.1}\,$(i)$ for Schr\"odinger operators on the real line. In addition, the case of half-line Jacobi operators and half-line matrix-valued Schr\"odinger operators was dealt with in \cite{GS00b}.
\medskip
{\it More recent references:} An even shorter proof of Theorem \ref{t9.1.1}\,$(i)$, close in spirit to Borg's original paper \cite{Bo52}, was found by Bennewitz \cite{Be01}. Still other proofs were presented by Horv\'ath
\cite{Ho01} and Knudsen \cite{Kn01}. Various local and global uniqueness results for matrix-valued Schr\"odinger, Dirac-type, and Jacobi operators were considered in \cite{GKM02}. The analog of the local
Borg--Marchenko theorem for certain Dirac-type systems was also studied by A.\ Sakhnovich \cite{Sa02}. The matrix-valued weighted Sturm--Liouville case has further been studied by Andersson \cite{An03}. He also studied uniqueness questions for certain scalar higher-order differential operators in \cite{An05}. A local Borg--Marchenko theorem for complex-valued potentials has been proved by Brown, Peacock, and Weikard \cite{BPW02}. The case of semi-infinite Jacobi operators with complex-valued coefficients was studied by Weikard \cite{We04}. A (global) uniqueness result for trees in terms of the (generalized) Dirichlet-to-Neumann map was found by Brown and Weikard \cite{BW05}.
\medskip
\noindent {\bf Acknowledgments.}
I'm grateful to Mark Asbaugh and the anonymous referee for a critical reading of this manuscript. I am also indebted to T.\ Tombrello for the kind invitation to visit the
California Institute of Technology, Pasadena, CA, during the fall
semester of 2005, where parts of this paper were written. The great
hospitality of the Department of Mathematics at Caltech is gratefully
acknowledged. Moreover, I thank the Research Council and the
Office of Research of the University of Missouri-Columbia for granting me
a research leave for the academic year 2005--2006.
|
\section{Introduction and notations}
Given $p>2$ a prime and $a,b,m\in\ZZ/p\ZZ$ such that $\leg
mp=\leg{a^2-mb^2}p=1$, the Legendre's symbol $\leg{a+b\k m}p$ is
defined as $\leg{a+b\a}p$, where $\a\in\ZZ/p\ZZ$ satisfies
$\a^2=m$. The definition is independent of the choice of the square
root $\a$ of $m$ because $(a+b\a )(a-b\a )=a^2-b^2m$ so
$\leg{a+b\a}p\leg{a-b\a}p=\leg{a^2-mb^2}p=1$ so
$\leg{a+b\a}p=\leg{a-b\a}p$. One may aslo define $\leg{a+b\k m}p$ if
$a^2-mb^2=0$ and $a\neq 0$. In this case the two radicals of $m$ are
$\a =\pm\h ab$. If we take $\a =-\h ab$ then $\leg{a+b\a}p=\leg 0p$,
which is not convenient. Hence we will take $\a =\h ab$ and we get
$\leg{a+b\k m}p=\leg{a+b\a}p=\leg{2a}p$.
In the particular case when $a=0$, $b=1$ we get $\leg{\k m}p$, defined
when $\leg mp=\leg{0^2-m\cdot 1^2}p=1$, i.e. when $p\ev 1\pmod 4$ and
$\leg mp=1$. This coincides with the rational quartic residue symbol
$\leg mp_4$.
The Legendre symbol $\leg{a+b\k m}p$ can also be defined when $a,b,m$
are $p$-adic integers satisfying $\leg mp=\leg{a^2-mb^2}p=1$ or $p\mid
a^2-mb^2$, $p\nmid a$.
For any prime $p$ of $\QQ$, including the archimedian prime $p=\j$, we
denote by $\leg{\cdot,\cdot}p:\QQ_p^\ti/(\QQ_p^\ti
)^2\ti\QQ_p^\ti/(\QQ_p^\ti )^2\to\{\pm 1\}$ the Hilbert symbol. (At
$p=\j$ we have $\QQ_\j =\RR$.)
Similarly as for $\leg{a+b\k m}p$ we can introduce the Hilbert symbol
$\leg{a+b\k m,c}p$. It is defined for $a,b,c,m\in\QQ_p$ satisfying
$m\in (\QQ_p^\ti )^2$ and $\leg{a^2-mb^2,c}p=1$. If
$\pm\a\in\QQ_p^\ti$ are the two square roots of $m$ then $\leg{(a+b\a
)(a-b\a ),c}p=\leg{a^2-mb^2,c}p=1$ so $\leg{a+b\a,c}p=\leg{a-b\a ,c}p$
so there is no ambiguity in this definition. If $a^2-mb^2=0$ and
$a\neq 0$, same as for $\leg{a+b\k m}p$, we define $\leg{a+b\k
m,c}p=\leg{2a,c}p$. Throughout this paper we will consider
$\leg{a+b\k m,c}p$ only for $p=2$.
Various mathematicians have obtained results involving Legendre
symbols of the type $\leg {a+b\k m}p$. Most of this results involve
the biquadratic symbols $\leg mp_4$ and they are called biquadratic or
quartic reciprocity laws. In this paper we will give some very general
reciprocity laws, which generalize all the existing results. In \S3 we
show how many of these results can be obtained as a consequence of
Theorem 2.5, which we only state in this paper.
\bdf For any $A\in\QQ^\ti/(\QQ^\ti )^2$ we denote by $\( A$ the only
squarefree integer such that $A=\( A(\QQ^\ti )^2$,
\edf
\subsection*{The algebras $T(V)$, $S(V)$ and $S'(V)$}${}$
For convenience we denote by $V:=\QQ^\ti/(\QQ^\ti )^2$ and for any
prime $p$ we denote $V_p:=\QQ_p^\ti/(\QQ_p^\ti )^2$. In particular,
$V_\j =\RR^\ti/(\RR^\ti )^2$.
Now $V$ and $V_p$ are $\ZZ/2\ZZ$-vector spaces. Therefore we may
consider the corresponding tensor algebras $T(V)$ and $T(V_p)$ and the
symmetric algebras $S(V)$ and $S(V_p)$. Note that, while $V$ has a
multiplicative notation, $T(V)$ and $S(V)$ have an additive one. In
order to prevent confusion for any $A_1,\ldots,A_n\in V$ we will
denote by $A_1\bu A_2\bu\cdots\bu A_n$ their product in $S^n(V)$
(instead of simply $A_1\cdots A_n$). For example, the distributivity
law on $S^2(V)$ will be written as $AB\bu C=A\bu C+B\bu C$.
We also define the algebra
$$S'(V)=T(V)/\la A_1\ot\cdots\ot A_n-A_{\pi (1)}\ot\cdots\ot A_{\pi
(n)}\mid\pi\in A_n\ra.$$
(Same definition as for $S(V)$ but this time the products are
invariant only under even permutations of the factors.)
We denote by $A_1\odot\cdots\odot A_n$ the image in $S'(V)$ of
$A_1\ot\cdots\ot A_n\in T(V)$. We will only be concerned with the
homogenous component of degree $3$, $S'^3(V)=T^3(V)/\la A\ot B\ot
C-B\ot C\ot A\ra$. We have $A\odot B\odot C=B\odot C\odot A=C\odot
A\odot B$ and $B\odot A\odot C=A\odot C\odot B=C\odot B\odot
A$. However $A\odot B\odot C\ne B\odot A\odot C$ (unless $A,B,C$ are
linearly dependent). Similarly we define $S'(V_p)$.
For any $\xi\in V$, $T(V)$, $S(V)$ or $S'(V)$ and any prime $p$ we
denote by $\xi_p$ the image of $\xi$ in $V_p$, $T(V_p)$, $S(V_p)$ or
$S'(V_p)$. When there is no danger of confusion we simply write $\xi$
instead of $\xi_p$.
\blm We have an exact sequence
$$S^3(V)\xz\tau S'^3(V)\xz\rho S^3(V)\to 0,$$
where $\tau$ is given by $A\bu B\bu C\mapsto A\odot B\odot C+B\odot
A\odot C$ and $\rho$ by $A\odot B\odot C\mapsto A\bu B\bu C$.
\elm
\pf The mapping $\rho$ is well defined and surjective because
$S^3(V)=T^3(V)/I$ and $S'^3(V)=T^3(V)/I'$ where $I,I'$ are bubgroups
of $T^3(V)$ with $I'\sb I$. For $\tau$ we note that the mapping
$:V^3\to S'^3(V)$ given by $(A,B,C)\mapsto A\odot B\odot C+B\odot
A\odot C$ is trilinear and also symmetric. (Recall that $A\odot B\odot
C=B\odot C\odot A=C\odot A\odot B$ and $B\odot A\odot C=A\odot C\odot
B=C\odot B\odot A$.) Hence $\tau$ is well defined.
We have $\rho (\tau (A\bu B\bu C)))=A\bu B\bu C+B\bu A\bu C=2(A\bu
B\bu C)=0$ so $\rho\circ\tau=0$ so $\im\tau\sbq\ker\rho$. For the
reverse inclusion take $\xi =\sum_iA_i\odot B_i\odot
C_i\in\ker\rho$. Then $0=\rho (\xi )=\sum_iA_i\bu B_i\bu C_i$ so there
are $A_{j,1},A_{j,2},A_{j,3}\in V$ and $\pi_j\in S_3$ such that
$\sum_iA_i\ot B_i\ot C_i=\sum_j(A_{j,1}\ot A_{j,2}\ot
A_{j,3}-A_{j,\pi_j(1)}\ot A_{j,\pi_j(2)}\ot A_{j,\pi_j(3)})$. This
implies that $\xi =\sum_iA_i\odot B_i\odot C_i=\sum_j(A_{j,1}\odot
A_{j,2}\odot A_{j,3}-A_{j,\pi_j(1)}\odot A_{j,\pi_j(2)}\odot
A_{j,\pi_j(3)})$. But $A_{j,\pi_j(1)}\odot A_{j,\pi_j(2)}\odot
A_{j,\pi_j(3)}$ equals $A_{j,1}\odot A_{j,2}\odot A_{j,3}$ or
$A_{j,2}\odot A_{j,1}\odot A_{j,3}$ if $\pi_j$ is even or odd,
respectively. So $A_{j,1}\odot A_{j,2}\odot
A_{j,3}-A_{j,\pi_j(1)}\odot A_{j,\pi_j(2)}\odot A_{j,\pi_j(3)}$ is $0$
if $\pi_j\in A_3$ and it is $A_{j,1}\odot A_{j,2}\odot
A_{j,3}-A_{j,2}\odot A_{j,1}\odot A_{j,3}=\tau (A_{j,1}\bu A_{j,2}\bu
A_{j,3})$ otherwise. It follows that $\xi =\tau (\eta )$, where $\eta
=\sum_{j,\pi_j\notin A_3}A_{j,1}\bu A_{j,2}\bu A_{j,3}$. \qed
\section{Main result}
\bdf We denote by $\mathcal D$ the set of all $(B,A,C)\in V^3$ such that
1) $\leg{A,B}p=\leg{A,C}p=\leg{B,C}p=1$ for every prime $p$, including
the archimedian prime $p=\j$.
2) $(\( A,\( B,\( C)=1$.
3) At least one of $\( A,\( B,\( C$ is $\ev 1\pmod 4$.
\edf
\bdf We define $f_1:{\mathcal D}\to\{\pm 1\}$ as follows. If
$(B,A,C)\in\mathcal D$ we take $x,y,z\in\ZZ$ with $(x,y,z)=1$ such
that $x^2-\( Ay^2=\( Bz^2$. (The existence of such $x,y,z$ is ensured
by Minkovski's theorem since $\leg{A,B}p=1$ for all $p$.) Then define
$f_1(B,A,C)=\a_\j\prod_{p\mid 2\( C}\a_p$, where
$$\a_\j =\begin{cases}1&\text{if }C>0\\
sgn(x)&\text{if }C<0\end{cases},$$
$$\a_p=\begin{cases}\leg{x+y\k{\( A}}p&\text{if }p\nmid\( A\\
\leg{2(x+z\k{\( B})}p&\text{if }p\nmid\( B\end{cases}$$
if $p\mid\( C$, $p>2$ and
$$\a_2=\begin{cases}1&\text{if }\( C\ev 1\pmod 8\\
\leg{x+y\k{\( A},C}2&\text{if }\( A\ev 1\pmod 8\\
\leg{2(x+z\k{\( B}),C}2&\text{if }\( B\ev 1\pmod 8\\
\leg{z,5}2&\text{if }\( A\ev\( C\ev 5\pmod 8\\
-\leg{y,5}2&\text{if }\( B\ev\( C\ev 5\pmod 8\\
-\leg{3x+z\k{5\( B},-1}2&\text{if }\( A\ev\( B\ev 5\pmod 8,~\( C\ev
-1\pmod 8\\
-\leg{5x+z\k{5\( B},3}2&\text{if }\( A\ev\( B\ev 5\pmod 8,~\( C\ev
3\pmod 8\
\end{cases}.$$
\edf
\bdf We define $f_2:{\mathcal D}\to\{\pm 1\}$ as follows. If
$(B,A,C)\in\mathcal D$ we take $x,y,z\in\ZZ$ with $(x,y,z)=1$ such
that $x^2-\( By^2=-\({ABC}z^2$. (The existence of such $x,y,z$ is
ensured by Minkovski's theorem since $\leg{A,B}p=\leg{C,B}p=1$ so
$\leg{-ABC,B}p=1$ for all $p$.) Then define
$f_2(B,A,C)=\b_\j\prod_{p\mid 2\( C}\b_p$, where
$$\b_\j =\begin{cases}1&\text{if }C>0\\
sgn(x)&\text{if }C<0\end{cases},$$
$$\b_p=\begin{cases}\leg xp&\text{if }p\nmid\( A\\
\leg{2(x+y\k{\( B})}p&\text{if }p\nmid\( B
\end{cases}$$
if $p\mid\( C$, $p>2$ and
$$\b_2=\begin{cases}1&\text{if }\( C\ev 1\pmod 8\\
\leg{x,C}2&\text{if }\( A\ev 1\pmod 8\\
\leg{2(x+y\k{\( B}),C}2&\text{if }\( B\ev 1\pmod 8\\
\leg{y+z\k{\h{\({ABC}}{\( B}},5}2&\text{if }\( A\ev\( C\ev 5\pmod 8\\
-\leg{z,5}2&\text{if }\( B\ev\( C\ev 5\pmod 8\\
\leg{-2(5x+y\k{5\( B}),C}2&\text{if }\( A\ev\( B\ev 5\pmod 8
\end{cases}.$$
\edf
\bff{\bf Note} If $x=0$ and $\( A\ev 1\pmod 8$ then we make the
convention that in the definition of $\b_2$ we don't use the option
$\b_2=\leg{x,C}2=\leg{0,C}2$, which is not defined. More precisely, if
$x=0$ then $0^2-\( By^2=-\({ABC}z^2$, which implies that $B=ABC$ so
$A=C$. Hence if $\( A\ev 1\pmod 8$ then also $\( C\ev 1\pmod 8$ so we
will choose the option $\b_2=1$. (See the special case 2.)
\eff
\bff{\bf Remark} In the definition of $f_1$ and $f_2$ we may replace
the condition that $x,y,z$ are relatively prime integers by the
condition that $x,y,z$ are relatively prime elements of $\ZZ [\h
12]$. Indeed, if we replace $x,y,z$ by $2^kx,2^ky,2^kz$ for some
$k\in\ZZ$ then $\a_\j$ is not changed, $\a_2$ is replaced by
$\leg{2^k,C}2\a_2$ and for any $p>2$, $p\mid\( C$ $\a_p$ is replced by
$\leg{2^k}p\a_p$. Since $\leg{2^k,C}2\prod_{p\mid\(
C,p>2}\leg{2^k}p=\prod_p\leg{2^k,C}p=1$ the product $\a_\j\prod_{p\mid
2\( C}\a_p$ is not changed. Similarly for the product
$\b_\j\prod_{p\mid 2\( C}\b_p$, which defines $f_2(B,A,C)$.
\eff
\bff{\bf Approximations of 2-adic square roots} In the definition of
$\a_2,\b_2$ for $\( C\not\ev 1\pmod 8$ we have formulas of the type
$\leg{a+b\k m,C}2$, where $a,b\in\QQ_2$ and $m\in 1+8\ZZ_2$. Here we
have the liberty of choosing either of the two quadratic roots of
$m$. Let $m=1+8x$. We make the convention that $\k m$ is the quare
root of $m$ that is $\ev 1\pmod 4$. We have $\k m+(1-4x)\ev 2\pmod 4$
and $(\k m)^2-(1-4x)^2=-16x(x-1)\vdots 32$ so $\k
m-(1-4x)\vdots 16$ so $\k m\ev 1-4x=\h{3-m}2\pmod{16}$.
Consider the $\pm$ sign such that
$$\ord_2(a\pm b\cdot\h{3-m}2)=
\min\{\ord_2(a+b\cdot\h{3-m}2),\ord_2(a-b\cdot\h{3-m}2)\}.$$
It follows
that $\ord_2(a\pm b\cdot\h{3-m}2)\leq\ord_2((a+b\cdot\h{3-m}2)+
(a-b\cdot\h{3-m}2))=\ord_22b\cdot\h{3-m}2=\ord_22b$. (Recall that
$\h{3-m}2=1-4x$ is an odd 2-adic integer.) Hence $a\pm
b\cdot\h{3-m}2\mid 2b$. Since also $b(\k m-\h{3-m}2)\vdots 16b$ we get
$$\h{a\pm b\k m}{a\pm b\cdot\h{3-m}2}-1=\h{\pm b(\k
m-\h{3-m}2)}{a\pm b\cdot\h{3-m}2}\vdots \h{16b}{2b}=8.$$
It follows that $\h{a\pm b\k m}{a\pm b\cdot\h{3-m}2}$ is a square so
$\leg{a\pm b\k m,C}2=\leg{a\pm b\cdot\h{3-m}2,C}2$.
If $\( C\ev 5\pmod 8$ then our formula becomes $\leg{a\pm b\k
m,5}2$. By the same reasoning as above, since $\k m\ev 1\pmod 4$, if
we take the $\pm$ sign such that $\ord_2(a\pm
b)=\min\{\ord_2(a+b),\ord_2(a-b)\}$ then $\h {a\pm b\k m}{a\pm b}\ev
1\pmod 2$ so $\leg{\h {a\pm b\k m}{a\pm b},5}2=1$, which implies that
$\leg{a\pm b\k m,5}2=\leg{a\pm b,5}2$.
\eff
Although calculating the Hasse symbols from the definition of
$\a_2,\b_2$ can be done very easily in numerical cases by using
2.3, proving general results may be quite labourios. Readers who want
to skip this part may check the short version of this paper.
{\bf Special cases}
{\bf 1.} $B=-A$ and we want to calculate $f_1(-A,A,C)$. Condition 1)
from the definition of $\mathcal D$ means that $\leg{A,C}p=\leg{-1,C}p=1$
$\forall p$ and condition 2) means that $(\( A,\( C)=1$. We have $\(
B=-\( A$ and we can take $x=0$, $y=z=1$. Since $\leg{-1,C}\j =1$ we
have $C>0$ so $\a_\j =1$. If $p\mid\( C$, $p>2$ then $\a_p=\leg{0+\k{\(
A}}p=\leg{\( A}p_4$. Also
$$\a_2=\begin{cases}1&\text{if }\( C\ev 1\pmod 8\\
\leg{\k{\( A},C}2&\text{if }\( A\ev 1\pmod 8\\
\leg{2\k{-\( A},C}2&\text{if }\( A\ev -1\pmod 8\\
1&\text{if }\( A\ev\( C\ev 5\pmod 8\\
-1&\text{if }\( A\ev 3\pmod 8,~\( C\ev 5\pmod 8\\
\end{cases}.$$
(The proof follows straight-forward from the definition of
$\a_2$. Since $\( B=-\( A$ the cases with $\( A\ev\( B\ev 5\pmod 8$ do
not occur. The condition $\( B\ev 1\pmod 8$ means $\( A\ev -1\pmod 8$
and the condition $\( B\ev\( C\ev 5\pmod 8$ means $\( A\ev 3\pmod 8$,
$\( C\ev 5\pmod 8$.)
{\bf 2.} $A=C$ and we want to calculate $f_2(B,C,C)$. Condition 1) in
the definition of $\mathcal D$ means $\leg{B,C}p=\leg{-1,C}p=1$ $\forall
p$ and condition 2) means that $(\( B,\( C)=1$. We have
$\({ABC}=\({BC^2}=\( B$ so we are looking for $x,y,z$ relatively
prime, such that $x^2-\( By^2=-\(B z^2$. One obvious choice is
$(x,y,z)=(0,1,1)$ but it is more convenable to choose $(x,y,z)=(0,\h
12,\h 12)$. (See Remark 2.2.) Since $\leg{-1,C}\j =1$ we have $C>0$ so
$\b_\j=1$. For $p\mid\( C$, $p>2$ we have $p\nmid\( B$ so we take
$\b_p=\leg{2(x+y\k{\( B})}p=\leg{2(0+\h 12\k{\( B})}p=\leg{\k{\(
B}}p=\leg{\( B}p_4$. If $\( A=\( C$ is odd then $\leg{C,-1}2=1$
implies $\( C\ev 1\pmod 4$. If $\( C\ev 1\pmod 8$ then $\b_2=1$. If
$\( A=\( C\ev 5\pmod 8$ then $\b_2=\leg{y+z\k{\h{\({ABC}}{\(
B}},5}2=\leg{\h 12+\h 12\k{\h{\( B}{\( B}},5}2=\leg{1,5}2=1$. If $\(
A=\( C$ is even then by the definition of $\mathcal D$ we have $\( B\ev
1\pmod 4$, which, together with $\leg{B,C}2=1$, implies $\( B\ev
1\pmod 8$. (We have $2\|\( C$.) It follows that $\b_2=\leg{2(x+y\k{\(
B}),C}2=\leg{2(0+\h 12\k{\( B}),C}2=\leg{\k{\( B},C}2$. Note that
$\k{\( B}$ is an odd $2$-adic integer so by multiplying with $\pm 1$
we may assume that $\k{\( B}\ev 1\pmod 4$. Hence $\b_2=\leg{\k{\(
B},C}2$ is $1$ or $-1$ according as $\k{\( B}\ev 1$ or $5\pmod
8$. But by 2.3 $\k{\( B}\ev\h{3-\( B}2\pmod 8$ so $\b_2=\leg{\k{\(
B},C}2$ is $1$ if $\( B\ev 1\pmod{16}$ and it is $-1$ if $\( B\ev
9\pmod{16}$. For short, if $\( C$ is even then $\b_2=\leg{\(
B}2_4$, where $\leg\cdot 2_4:1+8\ZZ_2\to\{\pm 1\}$ is given by $\leg
a2_4=1$ if $a\ev 1\pmod{16}$ and $\leg a2_4=-1$ if $a\ev
9\pmod{16}$. If $\( C$ is odd then $\b_2=1$. In conclusion:
$$f_2(B,C,C)=\prod_{p\mid\( C}\leg{\( B}p_4=:\leg{\( B}{\( C}_4.$$
{\bf 3.} $ABC=-1$, i.e. $A=-BC$, and we want to calculate
$f_2(B,-BC,C)$. In this case condition 1) in the definition of $\mathcal
D$ is equivalent to $\leg{B,C}p=1$ $\forall p$, while condition 2) is
vacuous. We have $\({ABC}=1$ so we need $x,y,z$ such that $x^2-\(
By^2=z^2$. An obvious choice is $x=z=1$, $y=0$. Since
$x>0$ we have $\b_\j =1$. If $p\mid\( C$, $p>2$
then $p\nmid\( A=-\({BC}$ is equivalent to $p\mid\( B$. In this case
$\b_p=\leg xp=1$. If $p\nmid\( B$ then $\b_2=\leg{2(x+y\k{\( B})}p=\leg
2p$. Also
$$\b_2=\begin{cases}\leg{2,C}2&\text{if }\( B\ev 1\pmod 8\\
\leg{-2,C}2&\text{if }\( B\ev 5\pmod 8\\
1&\text{otherwise }\end{cases}.$$
(Since $x=z=1$, $y=0$ and $\({ABC}=-1$ we get $\b_2=1$ if $\( C\ev
1\pmod 8$ or $\( A\ev 1\pmod 8$; $\b_2=\leg{2,C}2$ if $\( B\ev 1\pmod
8$; $\b_2=\leg{\k{\h{-1}{\( B}},5}2=1$ if $\( A\ev\( C\ev 5\pmod 8$
(in this case $\( B\ev -1\pmod 8$ so $\h{-1}{\( B}$ is a unit in $\ZZ_2$
and so is $\k{\h{-1}{\( B}}$); $\b_2=-1$ if $\( B\ev\( C\ev 5\pmod 8$;
$\b_2=\leg{-10,-1}2=-1$ if $\( A\ev\( B\ev 5\pmod 8$ (in this case
$\( C\ev -1\pmod 8$). Note that if $\( B\ev 5\pmod 8$ from
$\leg{B,C}2=1$ we get that $\( C$ is odd. If $\( B\ev 5,\( C\ev 1\pmod
8$ then $\b_2=1$; if $\( B\ev 5,\( C\ev 5\pmod 8$ then $\b_2=-1$; if
$\( B\ev 5,\( C\ev -1\pmod 8$, which is equivalent to $\( B\ev\( A\ev
5\pmod 8$, then $\b_2=-1$; if $\( B\ev
5,\( C\ev 3\pmod 8$, which is equivalent to $\( B\ev 5\pmod 8$, $\(
A\ev 1\pmod 8$, then $\b_2=1$. Hence if $\( B\ev 5\pmod 8$ then
$\b_2=\leg{-2,C}2$. If $\( B\ev 1\pmod 8$ then $\b_2=\leg{2,C}2$. In
all the other cases $\b_2=1$.)
Note that $\prod_{p\mid\( C,p>2}\b_p=\prod_{p\mid\( C,p\nmid 2\(B}\leg
2p=(-1)^{\h{n^2-1}8}$, where $n=\h{\( C}{(\( C,2\( B)}$. Since also
$\b_\j =1$ we get $f_2(B,-BC,C)=(-1)^{\h{n^2-1}8}\b_2$.
\bff{\bf Remark} If $p\mid\( C$, $p>2$ and $p\nmid\( A,\( B$ then in
the definition of $\a_p$ and $\b_p$ we can choose the formula from
both the $p\nmid\( A$ case and the $p\nmid\( B$ case. However the
outcome is the same.
Indeed, for $\a_p$ we have $x^2-\( Ay^2=\( Bz^2$ so $(x+y\k{\(
A})\cdot 2(x+z\k{\( B})=(x+y\k{\( A}+z\k{\ B})^2$. It follows that
$\leg{(x+y\k{\( A})\cdot 2(x+z\k{\( B})}p=1$ so $\leg{x+y\k{\(
A}}p=\leg{2(x+z\k{\( B})}p$.
For $\b_p$ we have $x^2-\(
By^2=\({ABC}z^2$ so $x\cdot 2(x+y\k{\( B})=(x+y\k{\(
B})^2-\({ABC}z^2\ev (x+y\k{\( B})^2\pmod p$. It follows that
$\leg{x\cdot 2(x+y\k{\( B})}p=1$ so $\leg xp=\leg{2(x+y\k{\( B})}p$.
We have similar redundancies in the definition of $\a_2,\b_2$ but
again all the cases of the definition which apply produce the same
outcome.
\eff
We now state our main result.
\btm (i) The functions $f_1$ and $f_2$ are well defined, i.e. they are
independent of the choice of $x,y,z$, and they are equal. We denote
$f=f_1=f_2$.
(ii) $f$ is symmetric in all three variables.
(iii) If $(B_i,A_i,C_i)\in\mathcal D$ for $1\leq i\leq n$ and
$\sum_iB_i\odot A_i\odot C_i=0$ then $\prod_if(B_i,A_i,C_i)=1$.
\etm
Theorem 2.5 justifies the following definition.
\bdf We denote by $W$ the subspace of $S'^3(V)$ generated by $B\odot
A\odot C$ with $(B,A,C)\in\mathcal D$. We define the group morphism $\c
:W\to\{\pm 1\}$ by $B\odot A\odot C\mapsto f(B,A,C)$.
\edf
\bff{\bf Remark} Part of Theorem 2.5(ii) follows from Theorem
2.5(iii). Namely, since $B\odot A\odot C=A\odot C\odot B=C\odot B\odot
A$, we have $f(B,A,C)=f(A,C,B)=f(C,B,A)$, i.e. $f$ has a circular
symmetry. However from Theorem 2.5(ii) we know that $f$ is symmetric,
not merely circular symmetric. Hence one may assume that we have a
more precise result, namely $\prod_if(B_i,A_i,C_i)=1$ whenever
$\sum_iB_i\bu A_i\bu C_i=0$ in $S^3(V)$. The reason that this
doesn't happen is the following. Assume that $(B_i,A_i,C_i)\in\mathcal D$
and $\sum_iB_i\bu A_i\bu C_i=0$. Then $\xi :=\sum_i B_i\odot A_i\odot
C_i$ can be written as $\xi =\sum_j(B_j'\odot A_j'\odot C_j'+A_j'\odot
B_j'\odot C_j')$ for some $A_j',B_j',C_j'\in V$. We have
$\prod_if(B_i,A_i,C_i)=\c (\xi )$. If $(B_j',A_j',C_j')\in\mathcal D$ for
all $j$ then $\c (\xi
)=\prod_jf(B_j',A_j',C_j')f(A_j',B_j',C_j')$. But from Theorem 2.5(ii)
we have $f(B_j',A_j',C_j')=f(A_j',B_j',C_j')$. Hence we get $\c (\xi
)=1$, as expected. The reason why this ``proof'' is wrong is that not
always $B_j',A_j',C_j'$ can be chosen such that
$(B_j',A_j',C_j')\in\mathcal D$.
\eff
\bff{\bf Example} Assume that $p,q,r,s,a,b,c,d$ are odd primes with
$p\ev q\ev r\ev s\ev 1\pmod 4$, $\leg ap=\leg bp=\leg cp=\leg dp=\leg
aq=\leg dq=\leg br=\leg dr=\leg cs=\leg ds=-1$ and $\leg bq=\leg
cq=\leg ar=\leg cr=\leg as=\leg bs=1$. Then
$(ad,pq,pq),(bd,pr,pr),(cd,ps,ps),(a,rs,rs),(b,qs,qs),\\ (c,qr,qr),
(abcd,pqrs,pqrs)\in\mathcal D$. Indeed, in all cases the condition 2)
from the definition of $\mathcal D$ is trivial and the conditon 3)
follows from $p\ev q\ev r\ev s\ev 1\pmod 4$. Condition 1) for the
first triplet means $\leg{ad,pq}t=1$ and $\leg{pq,-1}t=1$ for any
prime $t$. The second condition follows from the fact that $pq$ is a
sum of two squares. The first condition at $t=\j$ follows from $pq>0$
and at $t=2$ from the fact that $pq\ev 1\pmod 4$ and $ad$ is odd. At
$t=p,q,a$ and $d$ it means $\leg{ad}p=\leg{ad}q=\leg{pq}a=\leg{pq}d=1$
and it follows from $\leg ap=\leg pa=\leg dp=\leg pd=\leg aq=\leg
qa=\leg dq=\leg qd=-1$. For $t\neq\j,2,p,q,a,d$ our statement is
trivial. Similarly for the other triplets.
We also have $ad\bu pq\bu pq+bd\bu pr\bu pr+cd\bu ps\bu ps+a\bu rs\bu
rs+b\bu qs\bu qs+c\bu qr\bu qr+abcd\bu pqrs\bu pqrs=0$. However the
product
$f(ad,pq,pq)f(bd,pr,pr)f(cd,ps,ps)f(a,rs,rs)f(b,qs,qs)f(c,qr,qr)\\
f(abcd,pqrs,pqrs)$ is $-1$. Indeed, by the special case 2 our product
is equal to $\leg{ad}{pq}_4\leg{bd}{pr}_4\leg{cd}{ps}_4\leg
a{rs}_4\leg b{qs}_4\leg c{qr}_4\leg{abcd}{pqrs}_4=\\
\leg{a^2b^2c^2d^4}p_4\leg{a^2b^2c^2d^2}q_4\leg{a^2b^2c^2d^2}r_4
\leg{a^2b^2c^2d^2}s_4=\\ \leg{abc}p\leg{abcd}q\leg{abcd}r\leg{abcd}s
=(-1)\cdot 1\cdot 1\cdot 1=-1$.
\eff
\blm If $U=\tau\1 (W)$ then for every $p$ the image $U_p$ of $U\sbq
S^3(V)$ in $S^3(V_p)$ is generated by elements of the form $A\bu A\bu
B$ with $A,B\in V_p$.
\elm
\bff{\bf Remark} With the exception of the case $p=2$ we have
$\dim_{\ZZ/2\ZZ}V_p\leq 2$ so $U_p=S^3(V_p)$. If $p>2$ then
$V_p=\la\D_p,p\ra$, where $\D_p$ is a nonsquare unit in $\ZZ_p$ so
$S^3(V_p)=\la\D_p\bu\D_p\bu\D_p,\D_p\bu\D_p\bu p,p\bu p\bu\D_p,p\bu
p\bu p\ra =U_p$. If $p=\j$ then $V_\j =\RR^\ti/(\RR^\ti )^2=\la -1\ra$
so $S^3(V_\j )=\la (-1)\bu (-1)\bu (-1)\ra =U_\j$.
If $p=2$ then $V_2=\la -2,3,6\ra$ so a basis for $S^3(V_2)$ is $\{
A\bu B\bu C\mid A,B,C\in\{ -2,3,6\}\}$. A basis of $U_2$ is made of
all the elements of the basis for $S^3(V_2)$ except $(-2)\bu 3\bu 6$
so $U_2\sb S^2(V_2)$. (The dimensions of $S^3(V_2)$ and $U_2$ are $10$
and $9$, respectively.)
\eff
\bdf We define $\d :U\to\{\pm 1\}$ by $\d (\xi )=\prod_p\d_p(\xi_p)$,
where the product is taken over all primes, including $p=\j$, and $\d_p
:U_p\to\{\pm 1\}$ is given by $A\bu A\bu B\mapsto\leg{A,B}p$.
\edf
\btm $\c\circ\tau =\d$.
\etm
Theorem 2.10 provides a generalization of Theorem 2.5(iii). Namely, if
$(B_i,A_i,C_i)\in\mathcal D$ and $\sum_iB_i\bu A_i\bu C_i=0$ then we write
$\xi =\sum_iB_i\odot A_i\odot C_i\in W$ and we have $\rho (\xi
)=0$. It follows that $\xi\in\ker\rho =\im\tau$. (See Lemma 1.1.) Then
$\xi =\tau (\eta )$ for some $\eta\in\tau\1 (W)=U$ and we have
$\prod_if(B_i,A_i,C_i)=\c (\xi )=\c (\tau (\eta ))=\d (\eta )$. Note
that, in principle, calculating $\d (\eta )$ is easier than $\c (\xi
)$ as it only involves usual Legendre symbols. In order to calculate
$\c (\xi )$ one has to compute $f(B_i,A_i,C_i)$, which involves
finding nontrivial zeros for some ternary quadratic form and
calculating Legendre symbols of the type $\leg{a+b\k m}p$.
Some more explicit formulas for $\d_p$, so for $\d$ are given bellow:
\bff If $p=\j$ then $\d_\j :U_\j\to\{\pm 1\}$ is given by
$$\d_\j (A\bu B\bu C)=\begin{cases}-1&\text{if }A,B,C<0\\
1&\text{otherwise}\end{cases}.$$
If $p>2$ then $\d_p:U_p\to\{\pm 1\}$ is obtained as follows. For any
$A,B,C\in V$ if we write $\( A=p^ra$, $\( B=p^sb$ and $\( C=p^tc$,
where $r,s,t\in\{ 0,1\}$ and $p\nmid abc$ then
$$\begin{array}{ll}\d_p(A\bu B\bu C)=&\leg{-1}p^{rst}\leg ap^{st}\leg
bp^{rt}\leg cp^{rs}\cdot\\ {}&\cdot\left(\leg bp*\leg cp\right)^r
\left(\leg cp*\leg ap\right)^s\left(\leg ap*\leg
bp\right)^t,\end{array}$$
where $*:\{\pm 1\}\ti\{\pm 1\}\to\{\pm 1\}$ is given by
$$\e *\eta =\begin{cases}-1&\text{if }\e =\eta =-1\\
1&\text{otherwise}\end{cases}.$$
If $p=2$ and $\eta\in U$ then $\eta_2=\sum_iA_i\bu B_i\bu C_i$ with
$A_i,B_i,C_i\in V_2$. We write $A_i,B_i,C_i$ in terms of the basis
$-2,3,6$ of $V_2$. We have $A_i=(-2)^{r_{i,1}}3^{r_{i,2}}6^{r_{i,3}}$,
$B_i=(-2)^{s_{i,1}}3^{s_{i,2}}6^{s_{i,3}}$ and
$C_i=(-2)^{t_{i,1}}3^{t_{i,2}}6^{t_{i,3}}$ with
$r_{i,j},s_{i,j},t_{i,j}\in\{ 0,1\}$. Then
$$\d_2 (\eta )=\prod_i\prod_{j=1}^3(-1)^{r_{i,j}s_{i,j}t_{i,j}}.$$
Note that we may extend $\d_2:U_2\to\{\pm 1\}$ to the whole $S^3(V_2)$
by setting arbitrarily $\d_2 ((-2)\bu 3\bu 6)=1$. This way $\d
:U\to\{\pm 1\}$ extends to the whole $S^3(V)$. The formula above for
$\d_2$ will hold for for all $\eta\in S^3(V)$. If
$A=(-2)^{r_1}3^{r_2}6^{r_3}$, $B=(-2)^{s_1}3^{s_2}6^{s_3}$ and
$C=(-2)^{t_1}3^{t_2}6^{t_3}$ then $\d_2 (A\bu B\bu
C)=\prod_{j=1}^3(-1)^{r_js_jt_j}$.
\eff
In most applications we won't need Theorem 2.10. In fact in all proofs
from the next section we don't even need the full strength of Theorem
2.5(iii). It is enough to use a weaker version of Theorem 2.5(iii)
where the condition $\sum_iB_i\odot A_i\odot C_i=0$ is replaced by
$\sum_iB_i\ot A_i\ot C_i=0$. (Note that $\sum_iB_i\bu A_i\bu
C_i=0\Longrightarrow\sum_iB_i\odot A_i\odot
C_i=0\Longrightarrow\sum_iB_i\ot A_i\ot C_i=0$.) The only symmetry
properties we need are the ones following from Theorem 2.5(ii).
\section{Applications}
We now recover some results regarding quartic reciprocity that can be
found in [L, \S5] or on wikipedia at
http://en.wikipedia.org/wiki/Quartic$\_\,$reciprocity
\subsection*{Formulas for $\leg mp_4$}
By the special case 2 if $p\ev 1\pmod 4$ is a prime and $m=\pm
q_1\cdots q_k$ with $\leg{q_i}p=1$ then $(m,p,p)\in\mathcal D$ and
$$f_2(m,p,p)=\leg mp_4.$$
By Theorem 2.5(i) and (ii) we have $\leg
mp_4=f_2(m,p,p)=f_1(p,p,m)$. We have $p=a^2+b^2$ with $2\mid b$. Then
$p^2-b^2p=a^2p$ so in the definition of $f_1(p,p,m)$ we may take
$(x,y,z)=(p,b,a)$. We have $f_1(p,p,m)=\a_\j\prod_{q\mid 2m}\a_q$. Now
$x=p>0$ so $\a_\j =1$ and if $q\mid m$, $q>2$ then $q\nmid p=\( A$ so
$\a_q=\leg{x+y\k{\( A}}q=\leg{p+b\k p}q$.
Assume first that $m=2$. Then $\leg 2p=1$ so $p\ev 1\pmod 8$,
i.e. $4\mid b$. We have $\leg 2p_4=f_1(p,p,2)=\a_\j\a_2=\a_2$. Since
$\( A=p\ev 1\pmod 8$ we have $\a_2=\leg{x+y\k{\( A}}2=\leg{p+b\k
p,2}2$. Now $\k p$ is an odd integer in $\ZZ_2$ and so $4\mid b$ so
$b\k p\ev b\pmod 8$ so $p+b\k p\ev p+b\ev 1+b\pmod 8$. If $b\ev 0\pmod
8$ then $p+b\k p$ is $\ev 1\pmod 8$ so it is a square in $\ZZ_2^\ti$
so $\leg{p+b\k p,2}2=1$ so $\leg 2p_4=1$. If $b\ev 4\pmod 8$ then
$p+b\k p\ev 1+4=5\pmod 8$ so $\leg{p+b\k p,2}2=\leg{5,2}2=-1$ so $\leg
2p_4=-1$. In conclusion $\leg 2p_4=1$ iff $8\mid b$, which is one of
Euler's conjectures, proved by Gauss in 1828.
Take now $m=q^*:=(-1)^{\h{q-1}2}q$, where $q>2$ is prime, $\leg
qp=1$. Then $\leg{q^*}p_4=f_1(p,p,q^*)=\a_\j\a_q\a_2=\a_q\a_2$. We
have $\a_q=\leg{p+a\k p}q$ and since $\( A=p$ and $\( C=q^*$ and $p\ev
q^*\ev 1\pmod 4$ we have three cases for $\a_2$. If $q^*\ev 1\pmod 8$
then $\a_2=1$. If $p\ev 1\pmod 8$ then $\a_2=\leg{x+y\k{\(
A},C}2=\leg{p+b\k p,q^*}2$. But $p$ is odd and $b$ is even so $p+b\k p$
is an odd $2$-adic integer. Since also $q^*\ev 1\pmod 4$ we get again
that $\a_2=1$. If $p\ev q^*\ev 5\pmod 8$ then
$\a_2=\leg{x,5}2=\leg{p,5}2=1$. So $\a_2=1$ and
$\leg{q^*}p=\a_q=\leg{p+b\k p}q$. This result belongs to Lehmer. (See
[L, \S5.4, p. 167].)
\subsection*{Burde}
Assume that $p\ev q\ev 1\pmod 4$ and $\leg pq =1$ and write
$p=a^2+b^2$, $q=c^2+d^2$ with $b$ and $d$ even. Then $pq=e^2+f^2$,
where $e=ac-bd$, $f=ad+bc$ and $e$ is odd and $f$ even.
Now $p\odot p\odot q+p\odot q\odot q=p\odot pq\odot q$ so by Theorem
2.5(iii)
$$\leg pq_4\leg qp_4=f(p,p,q)f(q,q,p)=f(p,pq,q).$$
Now $f(p,pq,q)=f_2(pq,p,q)$. In the definition of $f_2$ we have $A=p$,
$B=pq$ and $C=q$ so $ABC=1$. Since $e^2-pq=-f^2$ we may take
$(x,y,z)=(e,1,f)$. We have $C=q>0$ so $\b_\j =1$ and the only primes
dividing $2\( C=2q$ are $2,q$ so $f_2(pq,p,q)=\b_q\b_2$. Now $q\nmid
p=\( A$ so $\b_q=\leg xq=\leg eq$. We have $\( A=pq$, $\( C=q$ so $\(
A\ev\( C\ev 1\pmod 4$. We have theree cases. If $\( C\ev 1\pmod 8$
then $\b_2=1$. If $\( C\ev 1\pmod 8$ then $\b_2=\leg{x,C}2=1$. ($x=e$
is odd and $\( C\ev 1\pmod 4$.) If $\( A\ev\( C\ev 5\pmod 8$ then
$\b_2=\leg{y+z\k{\h{\({ABC}}{\( B}},5}2=\leg{y\pm z,5}2$. (See 2.3.)
Here the $\pm$ sign is taken such that $\ord_2 (y\pm z)$ is
minimum. But for both choices of the $\pm$ sign $y\pm z=1\pm f$ is an
odd integer so again $\b_2=1$. Hence $\leg pq_4\leg qp_4=\b_q\b_2=\leg
eq\cdot 1=\leg{ac-bd}q$, which is Burde's law. (See [L, \S5.4,
p. 167].)
We now prove the other formulas for $\leg pq_4\leg qp_4$ from [L,
Theorem 5.7].
We write $\leg pq_4\leg qp_4=f(p,pq,q)=f_2(p,pq,q)$. We have $\(
A=pq$, $\( B=p$, and $\( C=q$ so $\({ABC}=1$. Since $a^2-p=-b^2$ we
can take $(x,y,z)=(a,1,b)$. Again $f_2(p,pq,q)=\b_q\b_2$. Since
$q\nmid\( B=p$ we have $\b_q=\leg{2(x+y\k{\( B})}q=\leg{\h 12(a+\k
p)}q$ and, by the same proof as in the previous case, $\b_2=1$. Hence
$\leg pq_4\leg qp_4=\leg{\h 12(a+\k p)}q$. To obtain $\leg pq_4\leg
qp_4=f_2(p,pq,q)=\leg{b+\k p}q$ we write $b^2-p=-a^2$ and we have
$(x,y,z)=(b,1,a)$. Same as before, $\b_q=\leg{\h 12(b+\k p)}q$, but
this time $\b_2=\leg 2q$, i.e. $\b_2=1$ if $\( C=q\ev 1\pmod 8$ and
$\b_2=-1$ if $\( C\ev 5\pmod 8$. Alternatively we use the fact that in
$\ZZ_q$ $2(a+\k p)(b+\k p)=(a+b+\k p)^2$ so $\leg{\h 12(a+\k
p)}q=\leg{b+\k p}q$. Similarly $\leg pq_4\leg qp_4=\leg{\h 12(c+\k
q)}p=\leg{d+\k p}q$.
In order to obtain $\leg pq_4\leg
qp_4=f(p,pq,q)=\leg{a+b\k{-1}}q=\leg{c+d\k{-1}}p$ we note that $p\odot
pq\odot q=p\odot (-pq)\odot q+p\odot (-1)\odot q$ so
$$\leg pq_4\leg qp_4=f(p,pq,q)=f(p,-pq,q)f(p,-1,q).$$
Now $f(p,-pq,q)=f_2(-pq,p,q)$ and since $A=p$, $B=-pq$ and $C=q$ we
have $ABC=-1$ so we can use the special case 3. We have $n=\h{\(
C}{(\( C,2\( B)}=1$ and, since $\( B=-pq\not\ev 1\pmod 4$,
$\b_2=1$. It follows that $f_2(-pq,p,q)=(-1)^{\h{n^2-1}8}\b_2=1$. So
$$f(p,pq,q)=f(p,-1,q)=f_1(p,-1,q).$$
We have $A=-1$, $B=p$ and $C=q$. Since $a^2+b^2=p$ we may take
$(x,y,z)=(a,b,p)$. We have $C=q>0$ so $\a_\j =1$ so
$f(p,-1,q)=\a_q\a_2$. Since $q\nmid -1=\( A$ we have
$\a_q=\leg{a+b\k{-1}}q$. We have $\( B=p$ an $\( C=q$ ao $\(
B\ev\( C\ev 1\pmod 4$. There are three cases. If $\( C\ev 1\pmod 8$
then $\a_2=1$. If $\( B\ev 1\pmod 8$ then $\a_2=\leg{2(x+z\k{\(
B}),C}2=\leg{2(a+\k p),C}2$. But $\( C\ev 1\pmod 4$ so by 2.3 we have
$\a_2=\leg{2(a\pm 1),C}2$, where the $\pm$ sign is chosen such that
$\ord_2(a\pm 1)$ is minimum. But $a$ is odd so we must have $a\pm 1\ev
2\pmod 4$ (so that $\ord_2(a\pm 1)=1$). Then $\h{a\pm 1}2$ is odd and
since $\( C\ev 1\pmod 4$ we have $\a_2=\leg{2(a\pm 1),C}2=\leg{\h{a\pm
1}2,C}2=1$. If $\( B\ev\( C\ev 5\pmod 8$ then
$\a_2=-\leg{y,5}2=-\leg{b,5}2$. But $a^2+b^2=p=\( C\ev 5\pmod 8$ and
$b$ is even so $b\ev 2\pmod 4$. It follows that $\leg{b,5}2=-1$ and
$\a_2=1$. Since $\a_2=1$ in all cases
$f_1(p,-1,q)=\a_q=\leg{a+b\k{-1}}q$. Note that since $q\nmid p=\( B$
we may also take $\a_q=\leg{2(a+\k p)}q$ and thus we recover the
equality $\leg pq_4\leg qp_4=\leg{\h 12(a+\k p)}q$ from
above. (Alernatively we may use the relation $2(a+b\k{-1})(a+\k
p)=(a+b\k{-1}+\k p)^2$ from Remark 2.4, which implies that $\leg{a+b\k
{-1}}q=\leg{2(a+\k p)}q$ so the two statements are equivalent.)
To prove [L, Ex. 5.5, p. 176] we note that we also may write
$$\leg pq_4\leg qp_4=f(p,-1,q)=f_1(p,q,-1).$$
We have $A=q$, $B=p$ and $C=-1$ so $f_1(p,q,-1)=\a_\j\a_2$. If
$e^2=pf^2+qg^2$ then $e^2-qg^2=pf^2$ so we may take
$(x,y,z)=(e,g,f)$. We want to prove that
$f_1(p,q,-1)=(-1)^{\h{fg}2}\leg{-1}e$. By permuting, if necessary $p$
and $q$ we may assume that if $p\ev q\ev 5\pmod 8$ then $f$ is even
and $g$ is odd and if $p\not\ev q\pmod 8$ then $p\ev 5,q\ev 1\pmod
8$. We have $\( A=q$, $\( B=p$ so $\( A\ev\( B\ev 1\pmod 4$ and $\(
C=-1$. Since $C<0$ we have $\a_\j =sgn(e)$. There are two cases:
$q\ev 1\pmod 8$ and $q\ev p\ev 5\pmod 8$. In the second case by our
assumption $g$ is even. In the first case $\( A\ev 1\pmod 8$ so
$\a_2=\leg{x+y\k{\( A},C}2=\leg{e+g\k q,-1}2$. If $g$ is even then
$\a_2=\leg{e,-1}2\leg{1+\h ge\k q,-1}2$. Now $\h 1e\k q$ is an odd
$2$-adic integer and $g$ is even. If $2\| g$ then $1+\h ge\k q\ev
3\pmod 4$ so $\leg{1+\h ge\k q,-1}2=-1$. If $4\mid g$ then $1+\h ge\k
q\ev 3\pmod 4$ so $\leg{1+\h ge\k q,-1}2=1$. Hence $\leg{1+\h ge\k
q,-1}2=(-1)^{\h g2}=(-1)^{\h{fg}2}$ and so
$\a_2=(-1)^{\h{fg}2}\leg{e,-1}2$. If $g$ is odd so $f$ is even note
that $pf^2=e^2-qf^2\ev 1-1=0\pmod 8$ so $4\mid f$. Now $e$ and $g\k q$
are odd $2$-adic integers so by replacing, if necessary $\k q$ by $-\k
q$ we have $e\ev g\k q\pmod 4$. It follows that $2\| e+g\k q$. Since
also $(e-g\k q)(e+g\k q)=e^2-qg^2=pf^2\vdots 16$ we get $e-g\k q\vdots
8$ so $e+g\k q\ev 2e\pmod 8$. It follows that $\h{e+g\k q}{2e}\ev
1\pmod 4$ so $\leg{\h{e+g\k q}{2e},-1}2=1$, which implies that
$\a_2=\leg{e+g\k
q,-1}2=\leg{2e,-1}2=\leg{e,-1}2=(-1)^{\h{fg}2}\leg{e,-1}2$. (Recall,
$4\mid f$.) In the second case $\( A\ev\( B\ev 5\pmod 8$ and $\( C=-1$
so $\a_2=-\leg{3x+z\k{5\(
B},-1}2=-\leg{3e+f\k{5p},-1}2=-\leg{3e,-1}2\leg{1+\h f{3e}\k
q,-1}2$. Now $pf^2=e^2-qg^2\ev 1-5\ev 4\pmod 8$ so $2\| f$. Since also
$\h 1{3e}\k{5p}$ is an odd $2$-adic integer we have $1+\h
f{3e}\k{5p}\ev 3\pmod 4$ so $\leg{1+\h f{3e}\k q,-1}2=-1$. Since also
$-\leg{3e,-1}2=\leg{e,-1}2$ we get
$\a_2=-\leg{e,-1}2=(-1)^{\h{fg}2}\leg{e,-1}2$. In conclusion $\leg
pq_4\leg qp_4=\a_\j\a_2=sgn(e)\cdot
(-1)^{\h{fg}2}\leg{e,-1}2=(-1)^{\h{fg}2}\leg{-1}e$.
Next we prove that if $p=r^2+qs^2$ then $\leg pq_4\leg qp_4=\leg
2q^s$ (see [L, Ex. 5.6, p. 176]). This time we use
$$\leg pq_4\leg qp_4=f(p,-1,q)=f_2(p,q,-1).$$
We have $A=q$, $B=p$ and $C=-1$ so $ABC=-pq$. Since $p^2-pr^2=pqs^2$
we may take $(x,y,z)=(p,r,s)$. We have $f_2(p,q,-1)=\b_\j\b_2$. Since
$x=p>0$ we have $\b_\j =1$ so we have to prove that $\b_2=\leg
2q^s$. Since $C=-1$ there are three cases for $\b_2$. If $q=\( A\ev
1\pmod 8$ then $\b_2=\leg{x,C}2=\leg{p,-1}2=1=\leg 2q^s$. If $q\ev
5\pmod 8$ and $p=\( B\ev 1\pmod 8$ then $\leg 2q^s=(-1)^s$ and
$\b_2=\leg{2(x+y\k{\( B}),C}2=\leg{2(p+r\k p),-1}2=\leg{p+r\k
p,-1}2$. If $s$ is odd and $r$ is even then $r^2=qs^2-p\ev 1-5\ev
4\pmod 8$ so $2\| r$. We have $\b_2=\leg{p,-1}2\leg{1+r\k{\h
1p},-1}2$. But $2\| r$ and $\k{\h 1p}$ is an odd integer so $1+r\k{\h
1p}\ev 3\pmod 4$ so $\leg{1+r\k{\h 1p},-1}2=-1$. It follows that
$\b_2=\leg{p,-1}2\cdot (-1)=-1=\leg 2q^s$. (Recall that $s$ is odd and
$q\ev 5\pmod 8$.) If $s$ is even and $r$ odd then $qs^2=p-r^2\ev
1-1=0\pmod 8$ so $4\mid s$. Since $p$, $r\k p$ are $2$-adic odd
integers, by multiplying $\k p$ with $\pm 1$ we may assume that $p+r\k
p\ev 2\pmod 4$, i.e. $2\| p+r\k p$. Together with
$p^2-pr^2=pqs^2\vdots 16$ we get that $8\mid p-r\k p$ so $p+r\k p\ev
2p\pmod 8$ so $\h 1{2p}(p+r\k p)\ev 1\pmod 4$ so $\leg{\h 1{2p}(p+r\k
p),-1}2=1$, which implies that $\b_2=\leg{2p,-1}2=1=\leg 2s^s$ ($s$ is
even). Now assume that $p\ev q\ev 5\pmod 8$, i.e. $\( A\ev\( B\ev
5\pmod 8$, so $\b_2=\leg{-2(5x+y\k{5\(
B}),C}2=\leg{-2(5p+r\k{5p}),-1}2=-\leg{5p+r\k{5p},-1}2$. If $s$ is
odd and $r$ even then $r^2=qs^2-p\ev 5-5=0\pmod 8$ so $4\mid r$ so
$5p+r\k{5p}\ev 5p\ev 1\pmod 4$ so $\b_2=-\leg{5p+r\k{5p},-1}2=-1=\leg
2q^s$. (Recall that $s$ is odd and $q\ev 5\pmod 8$.) If
$s$ is even and $r$ is odd then $qs^2=p-r^2\ev 5-1=4\pmod 8$ so $2\|
s$. By multiplying $\k{5p}$ with $\pm 1$ we may assume that $2\|
5p+r\k{5p}$. We also have $25p^2-5pr^2=20p^2+5pqs^2\ev 4+4\ev
8\pmod{16}$. ($p^2\ev 1\pmod 4$ so $20p^2\ev 20\ev 4\pmod{16}$ and
$5pq(\h s2)^2\ev 1\pmod 4$ so $5pqs^2\ev 4\pmod{16}$.) Hence
$5p-r\k{5p}\ev 4\pmod 8$ so $5p+r\k{5p}\ev 10p+4\ev 50+4\ev 6\pmod 8$,
which implies that $\leg{5p+r\k{5p},-1}2=-1$ so $\b_2=1=\leg 2q^s$
($s$ is even).
\subsection*{Scholz}
Assume that $p\ev q\ev 1\pmod 4$ are primes such that $\leg pq=1$ and
assume that $\e$ is a unit of $\QQ (\k p)$ of norm $-1$. Then Scholz's
law states that $\leg\e q=\leg pq_4\leg qp_4$. (See [L, p. 167].) We
could prove Scholz's law directly if $\e\in\ZZ [\leg{1+\k
p}2]\setminus\ZZ [\k p]$. However in order to overcome the $2$-adic
complications we replace $\e$ by $\e^3$ and we may assume that $\e
=t+u\k p\in\ZZ [\k p]$. We have
$$\leg pq_4\leg qp_4=f(p,-1,q)=f_1(-1,p,q)$$
so $A=p$, $B=-1$, $C=q$. Since $t^2-pu^2=-1$ we may take
$(x,y,z)=(t,u,1)$. We have $f_1(-1,p,q)=\a_\j\a_q\a_2$. But $C>0$ so
$\a_\j =1$ and $q\nmid p=\( A$ so $\a_p=\leg{x+y\k{\( A}}q=\leg{t+u\k
p}p=\leg\e p$. So we have to prove that $\a_2=1$. If $\( C=q\ev 1\pmod
8$ then $\a_2=1$ by definition. If $\( C=q\ev 5\pmod 8$ and $\( A
=p\ev 1\pmod 8$ then $\a_2=\leg{x+y\k {\( A},C}2=\leg{t+u\k
p,5}2$. But $t$ and $u$ have opposite parities so $t+u\k p$ is an odd
$2$-adic integer, which implies that $\a_2=1$. If $p\ev q\ev 5\pmod
8$, i.e. $\( A\ev\( C\ev 5\pmod 8$ then
$\a_2=\leg{z,5}2=\leg{1,5}2=1$.
\vskip 3mm
A similar result holds if $q=2$. Namely, if $p\ev 1\pmod 8$ then
$\leg{1+\k 2}p=\leg 2p_4\leg p2_4$. (See [L, p. 169].) We repeat the
reasoning used to prove that $\leg pq_4\leg qp_4=f(p,-1,q)$. By the
special case 2 we have $f_2(2,p,p)=\leg 2p_2$ and $f_2(p,2,2)=\leg
p2_4$ so $\leg 2p_4\leg p2_4=f(2,2p,p)$. We have
$f(2,2p,p)=f(2,-2p,p)f(2,-1,p)$. But $f(2,-2p,p)=f_2(2,-2p,p)$ can be
calculated using the special case 3 with $B=2$, $C=p$. We have
$n=\leg{\( C}{\( C,2\( B}=p$ and $\( B=2\not\ev 1\pmod 4$ so
$\b_2=1$. Thus $f_2(2,-2p,p)=(-1)^{\h{p^2-1}8}\cdot 1=1$ so $\leg
2p_4\leg p2_4=f(2,2p,p)=f(2,-1,p)=f_1(-1,2,p)$. We have $A=2$, $B=-1$,
$C=p$ so $f_1(-1,2,p)=\a_\j\a_p\a_2$. We have $1^2-2=-1$ so we may
take $(x,y,z)=(1,1,1)$. Since $C=p>0$ we have $\a_\j =1$ and since
$p\nmid 2=\( A$ we may take $\a_p=\leg{x+y\k{\( A}}p=\leg{1
+\k 2}p$ so we have to prove that $\a_2=1$. But this follows from $\(
C=p\ev 1\pmod 8$.
\section*{References}
[L] Lemmermeyer, Franz, {\it Reciprocity Laws: from Euler to
Eisenstein}, Berlin: Springer-Verlag, 2000
\end{document}
|
\section*{Introduction}
In many concrete situations the statistician observes a finite path $X_1, \ldots, X_n$ of a real temporal phenomena which can be modeled as realizations of a stationary process ${\mathbf{X}}:=(X_t)_{t\in\mathbb{Z}}$ (we refer, for example, to \cite{davis}, \cite{Ste} and references therein).
Here we consider a second order weakly stationary process, which implies that its mean is constant and that $\mathbb{E}(X_tX_s)$ only depends on the distance between $t$ and $s$. In the sequel, we will assume that the process is Gaussian, which implies that it is also strongly stationary, in the sense that, for any $t,n \in \mathbb{Z}$,
$$(X_1,\cdots,X_n)
\stackrel{\mathcal{L}}{=}
(X_{t+1},\cdots,X_{t+n}),\;(t\in\mathbb{Z},n\in\mathbb{N}).$$
Our aim is to predict this series when only a finite number of past values are observed. Moreover, we want a sharp control of the prediction error. For this, recall that, for Gaussian processes, the best predictor of $X_t, t \geq 0$, when observing $X_{-N}, \cdots, X_{-1}$, is obtained by a suitable linear combination of the $(X_i)_{i = -N, \cdots, -1}$. This predictor, which converges to the predictor onto the infinite past, depends on the unknown covariance of the time series. Thus, this covariance has to be estimated. Here, we are facing a blind filtering problem, which is a major difficulty with regards to the usual prediction framework.
Kriging methods often impose a parametric model for the covariance (see \cite{krig2}, \cite{AzDa}, \cite{Ste}). This kind of spatial prediction is close to our work. Nonparametric estimation may be done in a functional way (see \cite{Lili10}, \cite{pred2}, \cite{pred4}). This approach is not efficient in the blind framework. Here, the blind problem is bypassed using an idea of Bickel \cite{bickel} for the estimation of the inverse of the covariance. He shows that the inverse of the empirical estimate of the covariance is a good choice when many samples are at hand.
We propose in this paper a new methodology, when only a path of the process is observed. For this, following Comte \cite{comte}, we build an accurate estimate of the projection operator. Finally this estimated projector is used to build a predictor for the future values of the process. Asymptotic properties of these estimators are studied.
The paper falls into the following parts. In Section \ref{s:notations}, definitions and technical properties of time series are given. Section \ref{s:frame} is devoted to the construction of the empirical projection operator whose asymptotic behavior is stated in Section \ref{s:rate}. Finally, we build a prediction of the future values of the process in Section \ref{section_schur}. All the proofs are gathered in Section \ref{s:append}.
\section{Notations and preliminary definitions}
\label{s:notations}
In this section, we present our general frame, and recall some basic properties about time series, focusing on their predictions.
Let $\mathbf{X} = (X_k)_{k\in\mathbb{Z}}$ be a zero-mean Gaussian stationary process. Observing a finite past $X_{-N}, \cdots, X_{-1}$ ($N\geq 1$) of the process, we aim at predicting the present value $X_0$ without any knowledge on the covariance operator.
Since $X$ is stationary, let $r_{i-j } := \operatorname{Cov}(X_i,X_j), (i,j \in \mathbb{Z}) $ be the covariance between $X_i$ and $X_j$.
Here we will consider short range dependent processes, and thus we assume that
$$\sum_{k \in \mathbb{Z}} r_k^2 < + \infty,$$
So that there exists a measurable function
$f^\star \in \mathbb{L}_2\left(\left[0,2\pi\right)\right)$ defined by
\begin{equation*}
f^\star(t) : = \sum_{k= - \infty}^{\infty} r_{k}e^{ikt}, (a.e.)
\end{equation*}
This function is the so-called spectral density of the time series. It is real, even and non negative. As $\mathbf{X}$ is Gaussian, the spectral density conveys all the information on the process distribution.\\
Define the covariance operator $\Gamma$ of the process $\mathbf{X}$, by setting
$$\forall i,j \in \mathbb{Z},\Gamma_{ij} = \operatorname{Cov}(X_i,X_j) .$$
Note that $\Gamma$ is the Toeplitz operator associated to $f^\star$. It is usually denoted by $T(f^\star)$ (for a thorough overview on the subject, we refer to \cite{bottcher}).
This Hilbertian operator acts on $l^2(\mathbb{Z})$ as follows
$$\forall u \in l^2(\mathbb{Z}), i \in \mathbb{Z}, (\Gamma u)_i := \sum_{j \in \mathbb{Z}} \Gamma_{ij}u_j = \sum_{j \in \mathbb{Z}} r_{i-j}u_j = (T(f^\star)u)_i .$$
For sake of simplicity, we shall from now denote Hilbertian operators as infinite matrices.
Recall that for any bounded Hilbertian operator $A$, the spectrum $\operatorname{Sp}(A)$ is defined as the set of complex numbers $\lambda$ such that
$\lambda \operatorname{Id} - A$ is not invertible (here $\operatorname{Id}$ stands for the identity on $l^2(\mathbb{Z})$).
The spectrum of any Toeplitz operator, associated with a bounded function, satisfies the following property (see, for instance \cite{davis}):
$$\forall f \in \mathbb{L}_\infty \left(\left[0,2\pi\right)\right), \operatorname{Sp}(T(f)) \subset \left[\min(f), \max(f) \right] .$$
Now consider the main assumption of this paper :
\begin{hyp}\label{a:fbornee}
\begin{equation*}
\exists m,m' >0, \forall t \in \left[0,2\pi\right), m < f^\star(t) < m'.
\end{equation*}
\end{hyp}
This assumption ensures the invertibility of the covariance operator, since $f^\star$ is bounded away from zero.
As a positive definite operator, we can define its square-root $\Gamma^{\frac{1}{2}}$.
Let
$Q$ be any linear operator acting on $l^2(\mathbb{Z})$, consider the operator norm $
\left\| Q \right\|_{2,op} := \sup_{u_\in l_2(\mathbb{Z}), \left\|u\right\|_2 = 1} \left\| Qu\right\|_2,$
and define the warped operator norm as
\begin{equation*}
\left\| Q \right\|_{\Gamma} := \sup_{u_\in l_2(\mathbb{Z}), \left\|\Gamma^{\frac{1}{2}}u\right\|_2 = 1} \left\| \Gamma^{\frac{1}{2}}Qu\right\|_2.
\end{equation*}
Note that, under Assumption \eqref{a:fbornee} $\left\| \Gamma\right\|_{2,op} \leq m'$, hence the warped norm $\left\| . \right\|_\Gamma$ is well
defined and equivalent to the classical one
\begin{equation*}
\frac{m}{m'} \left\| Q \right\|_{2,op} \leq \left\| Q \right\|_{\Gamma} \leq \frac{m'}{m}\left\| Q \right\|_{2,op}.
\end{equation*}
Finally, both the covariance operator and its inverse are continuous with respect to the previous norms.\\
\indent The warped norm is actually the natural inducted norm over the Hilbert space $$H = \left(l_2(\mathbb{Z}),\langle.,. \rangle_\Gamma \right),$$
where
$$\langle x, y \rangle_\Gamma := x^T\Gamma y= \sum_{i,j \in \mathbb{Z}} x_i \Gamma_{ij} y_j.$$
From now on, all the operators are defined on $H$.
Set $$\mathbb{L}^2(\mathbb{P}) := \left\{Y \in \overline{\operatorname{Span}}\left((X_i)_{i \in \mathbb{Z}}\right), \mathbb{E}[Y^2]<+\infty \right\}$$
The following proposition (see for instance \cite{davis}) shows the
particular interest of $H$ :
\begin{prop} The map
\begin{align*}
\Phi : \quad H & \rightarrow \mathbb{L}^2(\mathbb{P}) \\
u & \rightarrow u^T \mathbf{X}= \sum_{i \in \mathbb{Z}} u_i X_i .
\end{align*}
defines a canonical isometry between $H$ and $\mathbb{L}_2(\mathbb{P})$.
\end{prop}
The isometry will enable us to consider, in the proofs, alternatively sequences $u \in H$ or the corresponding random variables $Y \in \mathbb{L}_2(\mathbb{P})$. \vskip .1in
We will use the following notations: recall that $\Gamma$ is the covariance operator and denote, for any $A,B \subset \mathbb{Z}$,
the corresponding minor $(A,B)$ by
$$\Gamma_{AB}:= \left(\Gamma_{ij}\right)_{i \in A, j \in B}.$$
Note that, when $A$ and $B$ are finite, $\Gamma_{AB}$ is the covariance matrix between $(X_i)_{i \in A}$ and $(X_j)_{j \in B}$. Diagonal minors will be simply written $\Gamma_{A}: = \Gamma_{AA}$, for any $A \in \mathbb{Z}$.\\
\indent In our prediction framework, let $O \subset \mathbb{Z}$ and assume that we observe the process $\mathbf{X}$ at times $i \in O $. It is well known that the best
linear prediction of a random variable $Y$ by observed variables
$(X_i)_{i \in O}$ is also the best prediction, defined by $P_O(Y):=\mathbb{E}\left[ Y | (X_i)_{i \in O} \right]$.
Using the isometry, there exist unique $u \in H$ and $v \in H$ with $Y=\Phi(u)$ and $P_O(Y)=\Phi(v)$.
Hence, we can define a projection operator acting on $H$, by setting $p_O(u):=v$. This corresponds to the natural projection in $H$ onto the set $\{ \Phi^{-1}(X_i),\: i \in O \}$.
Note that this projection operator may be written by block
$$p_O u := \left[\begin{matrix}
\Gamma_{O}^{-1}\Gamma_{O\mathbb{Z}}
\\ 0
\end{matrix}\right]
u.$$
The operator $\Gamma_{O}^{-1}$ is well defined since $f^\star \geq m >0$.
Finally, the best prediction observing $(X_i)_{i\in O}$ is $$\mathbb{E}\left[ Y=\Phi(u) | (X_i)_{i \in O} \right]=P_O(\Phi(u)) = \Phi(p_O u).$$
This provides an expression of the projection when the covariance $\Gamma$ is known. Actually, in many practical situations, $\Gamma$ is unknown and need to be estimated from the observations. Recall that we observe $X_{-N},\dots,X_{-1}$. We will estimate the covariance with this sample and use a subset of these observations for the prediction. This last subset will be $\left\{(X_i)_{i \in O_{K(N)}} \right\} $, with $O_{K(N)}:=\left[-K(N), \cdots -1\right]$. Here $\left(K(N)\right)_{N \in \mathbb{N}}$ is a growing suitable sequence.
Hence, the predictor $\hat{Y}$ will be here
$$\hat{Y} = \hat{P}_{O_{K(N)}} Y,$$
where $\hat{P}_{O_{K(N)}}$ denotes some estimator of the projection operator onto $O_{K(N)}$, built with the full sample $(X_i)_{i = -N, \cdots, -1}$.
As usual, we estimate the accuracy of the prediction by the quadratic error
$${\rm MSE}(\hat{Y})=\mathbb{E}\left[ \left(\hat{Y} - Y\right)^2 \right].$$
The bias-variance decomposition gives
$$\mathbb{E}\left[ \left(\hat{Y} - Y\right)^2 \right] =
\mathbb{E}\big[ \left(\hat{P}_{O_{K(N)}} Y - P_{O_{K(N)}} Y \right)^2 \big]
+
\mathbb{E}\big[ \left(P_{O_{K(N)}}Y- P_{\mathbb{Z}^-} Y \right)^2 \big]
+\mathbb{E}\big[ \left(P_{\mathbb{Z}^-} Y - Y \right)^2\big] ,$$
where
$$\hat{P}_{O_{K(N)}} Y =\hat{Y},$$
$$P_{O_{K(N)}} Y =\mathbb{E}\big[Y| (X_i)_{i \in O_{K(N)}} \big],$$
and
$$\hat{P}_{\mathbb{Z}^-} Y =\mathbb{E}\big[Y| (X_i)_{i <0} \big].$$
This error can be divided into three terms
\begin{itemize}
\item The last term $\mathbb{E}\big[ \left(P_{\mathbb{Z}^-} Y - Y \right)^2\big]$ is the prediction with infinite past error. It is induced by the variance of the unknown future values, and may be easily computed using the covariance operator. This variance does not go to zero as $N$
tends to infinity. It can be seen as an additional term that does not depend on the estimation procedure and thus will be omitted in the error term.
\item The second term $\mathbb{E}\big[ \left(P_{O_{K(N)}}Y- P_{\mathbb{Z}^-} Y \right)^2 \big]$ is a bias induced by the temporal threshold on the projector.
\item The first term $\mathbb{E}\big[ \left(\hat{P}_{O_{K(N)}} Y - P_{O_{K(N)}} Y \right)^2 \big]$ is a variance, due to the fluctuations of the estimation, and decreases to zero as soon as the estimator is consistent. Note that to compute this error, we have to handle the dependency between the prediction operator and the variable $Y$ we aim to predict.
\end{itemize}
Finally, the natural risk is obtained by removing the prediction with infinite past error:
\begin{align*}R(\hat{Y} = \hat{P}_{O_{K(N)}} Y ) &: =\mathbb{E}\big[ \left(\hat{P}_{O_{K(N)}} Y - P_{O_{K(N)}} Y \right)^2 \big]
+
\mathbb{E}\big[ \left(P_{O_{K(N)}}Y- P_{\mathbb{Z}^-_*} Y \right)^2 \big]
\\ & =\mathbb{E}\left[ \left( \hat{Y}-\mathbb{E}\left[Y|(X_i)_{i <0}\right] \right)^2 \right].\end{align*}
The global risk will be computed by taking the supremum of $R(\hat{Y})$ among of all random variables $Y$ in a suitable set (growing with $N$). This set
will be defined in the next section.
\section{Construction of the empirical projection operator}
\label{s:frame}
Recall that the expression of the empirical unbiased covariance estimator is given by (see for example \cite{AzDa})
\begin{equation*}
\forall ~0<p< N,~ \hat{r}^{(N)}(p) = \frac{1}{N-p}\sum_{k = -N}^{-p-1}X_k X_{k+p} .
\end{equation*}
Notice that, when $p$ is close to $N$, the estimation is hampered since we only sum $N-p$ terms. Hence, we will not use the complete available data but
rather use a cut-off.
Recall that $O_{K(N)} := [-K(N),-1]$ denotes the indices of the subset used for the prediction step.
We define the empirical spectral density as
\begin{equation} \label{specdens}
\hat{f}_K^{(N)} (t)= \sum_{p = -K(N)}^{K(N)} \hat{r}^{(N)}(p) e^{ipt}.
\end{equation}
We now build an estimator for $p_{O_{K(N)}}$ (see Section \ref{s:notations} for the definition of $p_{O_{K(N)}}$)
First, we divide the index space $\mathbb{Z}$ into $M_K \cup O_K \cup B_K \cup F_K$ where :
\begin{itemize}
\item $M_K = \left\{\cdots, -K-2, -K-1\right\}$ denotes the index of the past data that will not be used for the prediction (missing data)
\item $O_K = -K, \cdots, -1$ the index of the data used for the prediction (observed data)
\item $B_K = 0, \cdots, K -1$ the index of the data we currently want to forecast (blind data)
\item $F_K = K, K+1, \cdots$ the remaining index (future data)
\end{itemize}
In the following, we omit the dependency on $N$ to alleviate the notations.
As discussed in Section \ref{s:notations}, the projection operator $p_{O_K}$ may be written by blocks as:
\begin{equation*}
p_{O_K} = \left[ \begin{matrix} (\Gamma_{O_K})^{-1} \Gamma_{O_K \mathbb{Z}} \\ 0 \end{matrix}\right].
\end{equation*}
Since, we will apply this operator only to sequences with support in $B_K$, we may consider
$$\forall u \in l^2(\mathbb{Z}), \operatorname{Supp}(u) \subset B_K,
p_{O_KB_K}u :=\left[ \begin{matrix} (\Gamma_{O_K})^{-1} \Gamma_{O_K B_K} & 0 \\ 0& 0 \end{matrix} \right]u. $$
The last expression is given using the following block decomposition, if $B_K^C$ denotes the complement of $B_K$ in $\mathbb{Z}$ :
$$\left[ \begin{matrix}O_K B_K & O_K B_K^C \\ O_K^C B_K & O_K^C B_K^C \end{matrix} \right].$$
Hence, the two quantities $\Gamma_{O_K B_K}$ and $(\Gamma_{O_K})^{-1}$ have to be estimated. On the one hand, a natural estimator of the first matrix is given by $\hat{\Gamma}_{O_K B_K}$ defined as
\begin{equation*}
\left( \hat{\Gamma}_{O_KB_K}^{(N)} \right)_{ij} = \hat{r}^{(N)}(\left|j-i\right|), i \in O_K, j \in B_K.
\end{equation*}
On the other hand, a natural way to estimate $(\Gamma_{O_K})^{-1}$
could be to use $(\hat{\Gamma}^{(N)}_{O_K})$ (defined as $\left( \hat{\Gamma}_{O_K}^{(N)} \right)_{ij} = \hat{r}^{(N)}(\left|j-i\right|), i,j \in O_K $) and invert it.
However, it is not sure that this matrix is invertible. So, we will consider an empirical regularized version by setting
\begin{equation*}
\tilde{\Gamma}^{(N)} = \hat{\Gamma}^{(N)}_{O_K} + \hat{\alpha}I_{O_K},
\end{equation*}
for a well chosen $\hat{\alpha}$.
Set
\begin{equation*} \hat{\alpha} = -\min{\hat{f}_K^{(N)}} 1\!\!1_{\min{\hat{f}_K^{(N)}} \leq 0}+\frac{m}{4}1\!\!1_{\min{\hat{f}_K^{(N)}} \leq \frac{m}{4}}. \end{equation*}
so that $\left\|(\tilde{\Gamma}^{(N)}_{O_K})^{-1}\right\|_{2,op} \leq \frac{m}{4}$.
Remark that $\tilde{\Gamma}^{(N)}$ is the Toeplitz matrix associated to the function $\tilde{f}^{(N)} =\hat{f}_K^{(N)} + \hat{\alpha}$,
that has been tailored to ensure that $\tilde{f}^{(N)}$ is always greater than $\frac{m}{4}$, yielding the desired
control to compute $\tilde{\Gamma}^{-1}$.
Other regularization schemes could have been investigated.
Nevertheless, note that adding a translation factor makes computation easier than using, for instance, a threshold on
$\hat{f}^{(N)}_K$. Indeed, with our perturbation, we only modify the diagonal coefficients of the covariance matrix.
\vskip .1in
Finally, we will consider the following estimator, for any $Y\in \mathcal{B}_K:=\operatorname{Span}\left( (X_i)_{i \in B_K }\right)$:
$$\hat{Y}:=\hat{P}_{O_KB_K}^{(N)}(Y) = \Phi\left(\hat{p}_{O_KB_K}^{(N)}\Phi^{-1}(Y)\right),$$
where the estimator $\hat{p}_{O_KB_K}^{(N)}$ of $p_{\mathbb{Z}^-B_K}$, with window $K(N)$, is defined as follows
\begin{equation} \label{est}
\hat{p}_{O_KB_K}^{(N)} = \left(\tilde{\Gamma}^{(N)}_{O_K}\right)^{-1}\hat{\Gamma}^{(N)}_{O_KB_K}.
\end{equation}
\section{Asymptotic behavior of the empirical projection operator} \label{s:rate}
In this section, we give the rate of convergence of the estimator built previously (see Section \ref{s:frame}). We will bound uniformly the bias of prediction error for
random variables in the close future.
First, let us give some conditions on the sequence $(K(N))_{N \in \mathbb{N}})$:
\begin{hyp}\label{a:k}
The sequence $(K(N))_{N \in \mathbb{N}}$ satisfies
\begin{itemize}
\item $\lim K(N) \xrightarrow{N \rightarrow \infty} +\infty. $
\item $\lim \frac{K(N)\log(K(N))}{N} \xrightarrow{N \rightarrow \infty}0 .$
\end{itemize}
\end{hyp}
Recall that the pointwise risk in $Y \in \mathbb{L}^2(\mathbb{P})$ is defined by
$$R(\hat{Y})=\mathbb{E}\left[ \left( \hat{Y}-\mathbb{E}\left[Y|(X_i)_{i <0}\right] \right)^2 \right].$$
The global risk for the window $K(N)$ is defined by taking the supremum of the pointwise risk over all random variables $Y\in \mathcal{B}_K=\operatorname{Span}\left( (X_i)_{i \in B_K }\right)$
$$\mathcal{R}_{K(N)}\left(\hat{P}^N_{O_{K}B_K}\right)= \sup_{Y \in \mathcal{B}_K,\atop \operatorname{Var}(Y)\leq 1} R(\hat{P}^N_{O_{K}}(Y)) . $$
Notice that we could have chosen to evaluate the prediction quality only on $X_0$. Nevertheless the rate of convergence is not modified if we evaluate the prediction quality for all random variables from the close future. Indeed, the
major part of the observations will be used for the estimation, and the conditional expectation is taken only on the most $K(N)$ recent observations.
Our result will be then quite stronger than if we had dealt only with prediction of $X_0$.
To get a control on the bias of the prediction, we need some regularity assumption.
We consider Sobolev's type regularity by setting
\begin{equation*}
\forall s >1, W_s := \left\{g \in \mathbb{L}_2([0,2\pi)), g(t) = \sum_{k \in \mathbb{Z}} a_ke^{ikt}, \sum_{k \in \mathbb{Z}}k^{2s} a_k^2 < \infty \right\}.
\end{equation*}
and define
\begin{equation*}
\forall g \in W_s, g(t) = \sum_{k \in \mathbb{Z}} a_ke^{ikt} \left\|g\right\|_{W_s}: = \inf \left\{M, \sum_{k \in \mathbb{Z}}k^{2s} a_k^2 \leq M \right\}.
\end{equation*}
\begin{hyp} \label{a:sobol}
There exists $s \geq 1$ such that $f^\star \in W_s$.
\end{hyp}
We can now state our results. The following lemmas may be used in other frameworks than the blind problem. More precisely, if the blind prediction problem is very specific,
the control of the loss between prediction with finite and infinite past is more classical, and the following lemmas may be applied for that kind of questions.
The case where independent samples are available may also be tackled with the last estimators, using rates of convergences given in operator norms.
The bias is given by the following lemma
\begin{lem} \label{l:bias}
For $N$ large enough, the following upper bound holds
\begin{equation*}
\left\| p_{O_KB_k} - p_{\mathbb{Z}^-B_K} \right\|_{\Gamma} \leq C_2\frac{1}{K(N)^{\frac{2s-1}{2}}},
\end{equation*}
where $C_2 = \left\|\frac{1}{f^\star}\right\|_{W_{2s}} m'(1+\frac{m'}{m})$.
\end{lem}
In the last lemma, we assume regularity in terms of Sobolev's classes. Nevertheless, the proof may be written with some other kind of regularity.
The proof is given in appendix, and is essentially based on Proposition \ref{p:schur}. This last proposition provides the Schur block inversion of the projection operator.
The control for the variance is given in the following lemma:
\begin{lem}\label{l:var}
\begin{equation*}
\int_0^\infty \mathbb{P}\left(\left\| \hat{p}_{O_KB_K}^N - p_{O_KB_K} \right\|_{\Gamma}^4 > t \right) \mathrm{d} t
\leq C_0^4K(N)^4 (\frac{\log(K(N))}{N})^2 + o(K(N)^4 (\frac{\log(K(N))}{N})^2),
\end{equation*}
where $C_0= 4m'(\frac{6m'}{m^2} +\frac{4}{m}+2)$
\end{lem}
Again, we choose this concentration formulation to deal with the dependency of the blind prediction problem, but this result gives immediately a control of the variance of
the estimator whenever independent samples are observed (one for the estimation, and another one for the prediction).
The proof of this lemma is given in Section \ref{s:proof_lemma}. It is based on a concentration inequality of the estimators
$\hat{r}^{(N)}_p$ (see Comte \cite{comte}).
Integrating this rate of convergence over the blind data, we get our main theorem.
\begin{thm}\label{t:main}
Under Assumptions \ref{a:fbornee}, \ref{a:k} and \ref{a:sobol}, for $N$ large enough, the empirical estimator satisfies
\begin{equation*}
\sqrt{\mathcal{R} (\hat{P}^{(N)}_{O_KB_K} ) }\leq C_1\frac{K(N)^2\sqrt{\log(K(N))}}{\sqrt{N}} +C_2\frac{1}{K(N)^{\frac{2s-1}{2}}},
\end{equation*}
where $C_1$ and $C_2$ are given in Appendix.
\end{thm}
Again, the proof of this result is given in Section \ref{s:proof_main_thm}. It is quite technical. The main difficulty is induced by the blindness. Indeed, in this step, we have to deal with the dependency between the data and the empirical projector.
Obviously, the best rate of convergence is obtained by balancing the variance and the bias and finding the best window $K(N)$.
Indeed, the variance increases with $K(N)$ while the bias decreases.
Define $\hat{P}^{(N)}_\star$ as the projector $\hat{P}^{(N)}_{K^\star(N)}$ associated to the sequence $K^\star(N)$ that minimizes the bound in the last theorem.
We get:
\begin{cor}[Rate of convergence of the prediction estimator]
\label{tmain}
Under Assumptions $1.1$ and $2.1$, for $N$ large enough and choosing $K(N) = \Big{\lfloor} (\frac{N}{\log N})^{\frac{1}{2(2s+3)}} \Big{\rfloor}$, we get
\begin{equation} \label{rate}
\sqrt{\mathcal{R}(\hat{P}^{(N)}_\star)} \leq O\left( \left(\frac{\log N}{N}\right)^\frac{2s-1}{2(2s+3)}\right).
\end{equation}
\end{cor}
Notice that, in real life issues, it would be more natural to balance the risk given in Theorem \ref{t:main}, with the macroscopic term of variance given by
$$\mathbb{E}\big[Y-\mathbb{E}\left[Y|(X_i)_{i<0}\right] \big].$$
This leads to a much greater $K(N)$.
Nevertheless, Corollary \ref{tmain} has a theoretical interest. Indeed, it recovers the classical semi-parametric rate of convergence, and provides a way to get away from dependency.
Notice that, the estimation rate increases with the regularity $s$ of the spectral density $f^\star$.
More precisely, if $s\rightarrow \infty$, we obtain $(\frac{\log N}{N})^{\frac{1}{2}}$. This is, up to the $\log$-term, the optimal speed.
As a matter of fact, in this case, estimating the first coefficients of the covariance matrix is enough. Hence, the bias is very small.
Proving a lower bound on the mean error (that could lead to a minimax result), is a difficult task, since the
tools used to design the estimator are far from the usual estimation methods. \vskip .1in
\section{Projection onto finite observations with known covariance}\label{section_schur}
We aim at providing an exact expression for the projection operator.
For this, we generalize the expression given by Bondon (\cite{bondon}, \cite{bondon2}) for a projector onto infinite past.
Recall that, for any $A \subset \mathbb{Z}$, and if $A^C$ denotes the complement of $A$ in $\mathbb{Z}$, the projector $p_A$ may be written blockwise (see for instance \cite{Ste}) as:
$$p_A = \left[\begin{matrix} Id_A & \Gamma_A^{-1} \Gamma_{AA^C} \\0 & 0 \end{matrix}\right] .$$
Denote also $\Lambda := \Gamma^{-1} = T(\frac{1}{f^\star})$ the inverse of the covariance operator,
the following proposition provides an alternative expression of any projection operators.
\begin{prop}\label{p:schur}
One has
\begin{eqnarray*}
p_A
& = &
=\left[\begin{matrix} \operatorname{Id}_A & - \Lambda_{AA^C}\Lambda_{A^C}^{-1} \\0 & 0 \end{matrix}\right]
\end{eqnarray*}
Furthermore, the prediction error verifies $$\mathbb{E}\left[\left(P_A Y-Y\right)^2\right]=u^T\Lambda_{M}^{-1}u,$$
where $Y = \Phi(u) = u^T X.$
\end{prop}
\noindent The proof of this proposition is given in Appendix. We point out that this proposition is helpful for the computation of the bias.
Indeed, it gives a way to calculate the norm of the difference between two inverses operators.
\section{Appendix} \label{s:append}
\subsection{Proof of Proposition~\ref{p:schur}}
\begin{proof}
For the proof of Proposition \ref{p:schur}, let us choose
$$A \subset \mathbb{Z},$$
and denote the complement of $A$ in $\mathbb{Z}$ by
$$M := A^C $$
First of all, $\Lambda = \Gamma^{-1}$ is a Toeplitz operator over $H$ with eigenvalues in $[\frac{1}{m'} ;\frac{1}{m}]$. $\Lambda_M$
may be inverted as a principal minor of $\Lambda$.
Let us define the Schur complement of $\Lambda$ on sequences with support in $M$ : $S = \Lambda_A - \Lambda_{AM}\Lambda_{M}^{-1}\Lambda_{MA}$.
The next lemma provides an expression of $\Gamma_{A}^{-1}$ (see for instance \cite{schurbook}).
\begin{lem}\label{l:schur}
\begin{eqnarray*}
\Gamma_{A}^{-1} &= &S \\ & = &\Lambda_A - \Lambda_{AM}\Lambda_{M}^{-1}\Lambda_{MA}.
\end{eqnarray*}
\end{lem}
\begin{proof} of Lemma \ref{l:schur}
One can check
\begin{eqnarray*}
& &\left[\begin{matrix}\Lambda_A &\Lambda_{AM} \\ \Lambda_{MA} & \Lambda_M \end{matrix} \right] \left[\begin{matrix}S^{-1}&- S^{-1}\Lambda_{AM}\Lambda_M^{-1} \\ - \Lambda_M^{-1}\Lambda_{MA}S^{-1} & \Lambda_M^{-1} + \Lambda_M^{-1}\Lambda_{MA}S^{-1}\Lambda_{AM}\Lambda_M^{-1} \end{matrix}\right] \\
& = & \left[\begin{matrix}\Lambda_AS^{-1} -\Lambda_{AM}\Lambda_M^{-1}\Lambda_{MA}S^{-1} & -\Lambda_AS^{-1}\Lambda_{AM}\Lambda_M^{-1} + \Lambda_{AM}(\Lambda_M^{-1} + \Lambda_M^{-1}\Lambda_{MA}S^{-1}\Lambda_{AM}\Lambda_M^{-1}) \\ \Lambda_{MA}S^{-1} - \Lambda_M\Lambda_M^{-1}\Lambda_{MA}S^{-1} & - \Lambda_{MA}S^{-1}\Lambda_{AM}\Lambda_M^{-1} + \Lambda_M (\Lambda_M^{-1} + \Lambda_M^{-1}\Lambda_{MA}S^{-1}\Lambda_{AM}\Lambda_M^{-1} ) \end{matrix}\right]
\\ & = & \left[\begin{matrix}SS^{-1} & (\Lambda_{AM} \Lambda_M^{-1}\Lambda_{MA}S^{-1}+I_A-\Lambda_AS^{-1})\Lambda_{AM}\Lambda_M^{-1} \\ \Lambda_{MA}S^{-1}- \Lambda_{MA}S^{-1} & - \Lambda_{MA} S^{-1}\Lambda_{AM}\Lambda_M^{-1} + I_M +\Lambda_{MA}S^{-1}\Lambda_{AM}\Lambda_M^{-1} \end{matrix}\right]
\\ &=&\left[\begin{matrix}I_A& 0 \\ 0 & I_M \end{matrix}\right].
\end{eqnarray*}
Since the matrix are symmetric, we can transpose the last equality. We obtain that
\begin{eqnarray*}
\left[\begin{matrix}S^{-1}&- S^{-1}\Lambda_{AM}\Lambda_M^{-1} \\ - \Lambda_M^{-1}\Lambda_{MA}S^{-1} & \Lambda_M^{-1} + \Lambda_M^{-1}\Lambda_{MA}S^{-1}\Lambda_{AM}\Lambda_M^{-1} \end{matrix}\right] & = & \Lambda^{-1}
\\ &= &\Gamma.
\end{eqnarray*}
So that $\Gamma_A = S^{-1} $.
\end{proof}
We now compute the projection operator:
\begin{eqnarray*}
p_A &= &
\left[\begin{matrix} Id_A & \Gamma_A^{-1}\Gamma_{AM} \\0 & 0 \end{matrix}\right] \\
& = & \left[\begin{matrix} Id_A & S \Gamma_{AM} \\0 & 0 \end{matrix}\right]\\
& = & \left[\begin{matrix} Id_A & (\Lambda_A - \Lambda_{AM}\Lambda_{M}^{-1}\Lambda_{MA})\Gamma_{AM} \\0 & 0 \end{matrix}\right]\\
& = & \left[\begin{matrix} Id_A & \Lambda_A \Gamma_{AM} - \Lambda_{AM}\Lambda_{M}^{-1}(Id_M - \Lambda_M \Gamma_M) \\0 & 0 \end{matrix}\right]\\
& = & \left[\begin{matrix} Id_A & \Lambda_A \Gamma_{AM} - \Lambda_{AM}\Lambda_{M}^{-1} + \Lambda_{AM} \Gamma_M \\0 & 0 \end{matrix}\right]\\
& = & \left[\begin{matrix} Id_A & - \Lambda_{AM}\Lambda_{M}^{-1} \\0 & 0 \end{matrix}\right].
\end{eqnarray*}
Where we have used $\Lambda \Gamma = Id$ in the last two lines.
Now consider $Q$ the quadratic error operator. It is defined as
$$\forall u \in l^2(\mathbb{Z}), u^TQu := \left\|(p_Au-u)^2 \right\|_{\Gamma} = \mathbb{E}\left[(\Phi(u)- P_A \Phi(u) )^2\right]. $$
This operator $Q$ can be obtained by a direct computation (writing the product right above), but it is easier to use the expression of the variance of a projector in the Gaussian case given for instance by \cite{Ste}.
\begin{equation*}
Q = \Gamma_M-\Gamma_{MA}\Gamma_{A}^{-1} \Gamma_{AM}
\end{equation*}
Again, notice that $Q$ is the Schur complement of $\Gamma$ on sequences with support in $A$, and thanks to Lemma~\ref{l:schur} applied to $\Lambda$ instead of $\Gamma$,
we get
\begin{equation*}
Q = \Lambda_M^{-1}.
\end{equation*}
This ends the proof of Proposition \ref{p:schur}.
\end{proof}
\subsection{Proof of Theorem \ref{t:main}}
\label{s:proof_main_thm}
\begin{proof} of Theorem \ref{t:main}
Recall that we aim at providing a bound on $\sqrt{\mathcal{R}(\hat{P}^{(N)}_{O_KB_K})}$.
Notice first that we have
$$\sqrt{\mathcal{R}(\hat{P}^{(N)}_{O_KB_K})} \leq \sqrt{\sup_{Y \in \mathcal{B}_K,\atop \operatorname{Var}(Y)\leq 1}
\mathbb{E}\left[ (\hat{P}_{O_KB_K}^{(N)}(Y) - P_{O_KB_K}(Y))^2\right] } + \sqrt{\mathcal{R}(P_{O_KB_K}))} . $$
Using Lemma \ref{l:bias} for a sequence $(K(N))_{N \in \mathbb{N}}$ and a centered random variable $Y \in \operatorname{Span}\left((X_i)_{i \in B_K}\right)$ such that $\mathbb{E}\left[Y^2\right] = 1$, we have
\begin{eqnarray*}
\sqrt{\mathcal{R}(P_{O_KB_K}))} & \leq & \left\| p_{O_KB_K} - p_{\mathbb{Z}^-B_K} \right\|_{\Gamma} \sqrt{\mathbb{E}\left[Y^2\right] }
\\ & \leq & C_2\frac{1}{K(N)^{\frac{2s-1}{2}}} .
\end{eqnarray*}
For the variance, we first notice that $Y = \Phi(u) = u^T\mathbf{X}$,
\begin{equation*} 1 = \mathbb{E}\left[Y^2\right] = u^T\Gamma_{B_K}u \geq m u^Tu = m\sum_{i=0}^{K(N)-1}u_i^2,\end{equation*}
Denote $A=\hat{p}_{O_KB_K}^{(N)} - p_{O_KB_K}$. We can write, by applying twice Cauchy-Schwarz's inequality,
\begin{eqnarray*}
\mathbb{E}\left[\Big(\hat{P}_{O_KB_K}^{(N)}Y - P_{O_KB_K}Y\Big)^2\right] & = & \int_\omega \Big( \sum_{ i = -K(N)}^{-1} \sum_{j = 0}^{K(N)-1} A_{ij}(\omega)u_jX_i(\omega)\Big)^2\mathrm{d} \mathbb{P}(\omega)
\\ & \leq & \int_\omega \sum_{i = -K(N)}^{-1} (\sum_{j = 0}^{K(N)-1} A_{ij}(\omega)u_j)^2 \sum_{i = -K(N)}^{-1} X_i^2(\omega)\mathrm{d} \mathbb{P}(\omega)
\\ & \leq & \int_\omega \sum_{ i = -K(N)}^{-1} \sum_{j = 0}^{K(N)-1} A_{ij}^2(\omega) \sum_{j = 0}^{K(N)-1} u_j^2 \sum_{i = -K(N)}^{-1} X_i^2(\omega)\mathrm{d}
\mathbb{P}(\omega),
\end{eqnarray*}
So that,
\begin{eqnarray*}
\mathbb{E}\left[\Big(\hat{P}_{O_KB_K}^{(N)}Y - P_{O_KB_K}Y\Big)^2\right] & \leq & \int_\omega \sum_{ i = -K(N)}^{-1}\sum_{j = 0}^{K(N)-1} A_{ij}^2(\omega) \frac{1}{m} \sum_{i = n_0+1}^{K(N)+n_0} X_i^2\mathrm{d} \mathbb{P}(\omega).
\end{eqnarray*}
Using the following equivalence between two norms for finite matrices with size $(n,m)$ (see for instance \cite{matrixhandbook}),
\begin{equation*}
\sqrt{\sum_{i = 1}^{n} \sum_{ j = 1}^{m} A_{ij}^2} \leq \sqrt{n} \left\| A \right\|_{2,op},
\end{equation*}
we obtain
\begin{eqnarray*}
\mathbb{E}\left[\Big(\hat{P}_{O_KB_K}^{(N)}Y - P_{O_KB_K}Y\Big)^2\right] & \leq & \frac{K(N)}{m} \int_\omega \left\| A (\omega)\right\|_{2,op}^2 \sum_{i = n_0+1}^{K(N)+n_0} X_i^2(\omega)\mathrm{d} \mathbb{P}(\omega).
\end{eqnarray*}
Further,
\begin{eqnarray*}
\mathbb{E}\left[\Big(\hat{P}_{O_KB_K}^{(N)}Y - P_{O_KB_K}Y\Big)^2\right]
& \leq & \frac{K(N)}{m} \int_\omega \left\| A(\omega) \right\|_{2,op}^2 \sum_{j = n_0+1}^{K(N)+n_0} X_j^2(\omega)\mathrm{d} \mathbb{P}(\omega)
\\ & \leq & \frac{K(N)}{m} \sqrt{\int_\omega \left\| A (\omega)\right\|_{2,op}^4\mathrm{d} \mathbb{P}(\omega)} \sqrt{\int_\omega\left(\sum_{j = n_0+1}^{K(N)+n_0} X_j^2(\omega)\right)^2\mathrm{d} \mathbb{P}(\omega)}
\\ & \leq & \frac{K(N)}{m} \sqrt{\int_{\mathbb{R}^+} \mathbb{P}\left( \left\| A \right\|_{2,op}^4>t \right)\mathrm{d} t} \sqrt{K(N)^2\int_\omega\left( X_j^4\right)\mathrm{d} \mathbb{P}(\omega)},
\end{eqnarray*}
We have used here again Cauchy-Schwarz's inequality and the fact that, for all nonnegative random variable $Y$,
\begin{equation*}
\mathbb{E}\left[Y\right] = \int_{\mathbb{R}^+} \mathbb{P}\left( Y>t\right) \mathrm{d} t.
\end{equation*}
Since $X_0$ is Gaussian, its moment of order four $r_4$ is finite. Then Lemma \ref{l:var} yields that, for $N$ large enough,
\begin{equation*}
\mathbb{E}\left[\Big(\hat{P}_{O_KB_K}^{(N)}Y - P_{O_KB_K}Y\Big)^2\right] \leq \frac{C_0^2\sqrt{r_4}K(N)^4\log(K(N)))}{mN}.
\end{equation*}
So that,
\begin{equation*}
\sqrt{\sup_{Y \in \mathcal{B}_K,\atop \operatorname{Var}(Y)\leq 1}
\mathbb{E}\left[ (\hat{P}_{O_K}^{(N)}(Y) - P_{O_K}(Y))^2\right] } \leq \frac{C_1K(N)^2\sqrt{\log(K(N))}}{\sqrt{N}},
\end{equation*}
with $C_1 = \frac{C_0\sqrt[4]{r_4}}{\sqrt{m}}$. This ends the proof of the theorem.
\end{proof}
\subsection{Proofs of concentration and regularity lemmas }
\label{s:proof_lemma}
First, we compute the bias and prove Lemma~\ref{l:bias} :
\begin{proof}{of Lemma \ref{l:bias}}
Recall that we aim to obtain a bound on $\left\| p_{O_KB_K} - p_{\mathbb{Z}^-B_K} \right\|_{\Gamma}$.
Using Proposition~\ref{p:schur}, we can write
\begin{eqnarray*}
\left\| p_{O_KB_K} - p_{\mathbb{Z}^-B_K} \right\|_{\Gamma} &\leq& \left\| p_{O_K\mathbb{Z}^+} - p_{\mathbb{Z}^-\mathbb{Z}^+} \right\|_{\Gamma}
\\ & \leq & \left\| \left[\begin{matrix} (\Gamma_{O_K})^{-1}\Gamma_{O_K\mathbb{Z}^+} \\ 0 \end{matrix}\right] - \left[\begin{matrix} - \Lambda_{O_K\mathbb{Z}^+}(\Lambda_{\mathbb{Z}^+})^{-1} \\
-\Lambda_{M_K^-\mathbb{Z}^+}(\Lambda_{\mathbb{Z}^+})^{-1} \end{matrix}\right] \right\|_{\Gamma}.
\end{eqnarray*}
So that, using the norms equivalence,
\begin{eqnarray*}
\left\| p_{O_KB_K} - p_{\mathbb{Z}^-B_K} \right\|_{\Gamma}& \leq & \frac{m'}{m} \left\| \left[\begin{matrix} (\Gamma_{O_K})^{-1}\Gamma_{O_K\mathbb{Z}^+} \\ 0 \end{matrix}\right] - \left[\begin{matrix} - \Lambda_{O_K\mathbb{Z}^+}(\Lambda_{\mathbb{Z}^+})^{-1} \\
-\Lambda_{M_K^-\mathbb{Z}^+}(\Lambda_{\mathbb{Z}^+})^{-1} \end{matrix}\right] \right\|_{2,op} \\
& \leq & \frac{m'}{m} \left\| \left[ \begin{matrix} (\Gamma_{O_K})^{-1}\Gamma_{O_K\mathbb{Z}^+} + \Lambda_{O_K\mathbb{Z}^+}(\Lambda_{\mathbb{Z}^+})^{-1} \\ \Lambda_{M_K^-\mathbb{Z}^+}(\Lambda_{\mathbb{Z}^+})^{-1} \end{matrix}\right] \right\|_{2,op} \\
& \leq & \frac{m'}{m}\left\| \left[ \begin{matrix} (\Gamma_{O_K})^{-1}\Gamma_{O_K\mathbb{Z}^+}\Lambda_{\mathbb{Z}^+} + \Lambda_{O_K\mathbb{Z}^+}\\ \Lambda_{M_K^-\mathbb{Z}^+}\end{matrix}\right] \right\|_{2,op} \left\| (\Lambda_{\mathbb{Z}^+})^{-1} \right\|_{2,op}\\
& \leq & \frac{m'}{m}\left\| (\Lambda_{\mathbb{Z}^+})^{-1} \right\|_{2,op} \left( \left\|(\Gamma_{O_K})^{-1}\Gamma_{O_K\mathbb{Z}^+}\Lambda_{\mathbb{Z}^+} + \Lambda_{O_K\mathbb{Z}^+} \right\|_{2,op} + \left\| \Lambda_{M_K^-\mathbb{Z}^+} \right\|_{2,op} \right).
\end{eqnarray*}
The last step follows from the inequality:
\begin{equation*}\left\| \begin{matrix}A \\ B \end{matrix} \right\|_{2,op} \leq \left\| \begin{matrix}A \\ 0 \end{matrix} \right\|_{2,op} + \left\| \begin{matrix}0 \\ B \end{matrix} \right\|_{2,op} = \left\| A \right\|_{2,op} + \left\| B \right\|_{2,op}.
\end{equation*}
But, since $\Lambda= \Gamma^{-1}$,
\begin{equation*}
\Gamma_{O_K\mathbb{Z}^+}\Lambda_{\mathbb{Z}^+} + \Gamma_{O_K}\Lambda_{O_K\mathbb{Z}^+} = - \Gamma_{O_KM_K^-}\Lambda_{M_K^-\mathbb{Z}^+}.
\end{equation*}
So, we obtain,
\small{\begin{eqnarray*}
\left\| p_{O_KB_K} - p_{\mathbb{Z}^-B_K} \right\|_{\Gamma}
& \leq & \frac{m'}{m}\left\| (\Lambda_{\mathbb{Z}^+})^{-1} \right\|_{2,op} \left( \left\|(\Gamma_{O_K})^{-1}\left(- \Gamma_{O_KM_K^-}\Lambda_{M_K^-\mathbb{Z}^+}\right) \right\|_{2,op} + \left\| \Lambda_{M_K^-\mathbb{Z}^+} \right\|_{2,op} \right)\\
& \leq & \frac{m'}{m} \left\| (\Lambda_{\mathbb{Z}^+})^{-1} \right\|_{2,op} \left( \left\|(\Gamma_{O_K})^{-1} \right\|_{2,op} \left\| - \Gamma_{O_KM_K^-}\Lambda_{M_K^-\mathbb{Z}^+} \right\|_{2,op} + \left\| \Lambda_{M_K^-\mathbb{Z}^+} \right\|_{2,op} \right)\\
& \leq & \frac{m'}{m}\left\| (\Lambda_{\mathbb{Z}^+})^{-1} \right\|_{2,op} \left( \left\|(\Gamma_{O_K})^{-1} \right\|_{2,op} \left\| \Gamma_{O_K M_K^-} \right\|_{2,op} \left\| \Lambda_{M_K^-\mathbb{Z}^+} \right\|_{2,op} + \left\| \Lambda_{M_K^-\mathbb{Z}^+} \right\|_{2,op} \right) \\
& \leq & \frac{m'}{m} \left\| (\Lambda_{\mathbb{Z}^+})^{-1} \right\|_{2,op} \left( \left\|(\Gamma_{O_K})^{-1} \right\|_{2,op} \left\| \Gamma_{O_K M_K^-} \right\|_{2,op} + 1 \right) \left\| \Lambda_{M_K^-\mathbb{Z}^+} \right\|_{2,op} .
\end{eqnarray*}}
But, we have,
\begin{equation*}\left\| (\Lambda_{\mathbb{Z}^+})^{-1} \right\|_{2,op} \leq m', \end{equation*} as the inverse of a principal minor of $\Lambda$.
\begin{equation*}\left\|(\Gamma_{O_K})^{-1} \right\|_{2,op} \leq \frac{1}{m}, \end{equation*} since it is the inverse of a principal minor of $\Gamma$.
\begin{equation*} \left\| \Gamma_{O_KM_K^-} \right\|_{2,op} \leq m' ,\end{equation*} as an extracted operator of $\Gamma$.
Thus, we get
\begin{equation*}
\left\| p_{O_KB_K} - p_{\mathbb{Z}^-B_K} \right\|_{\Gamma} \leq C_4 \left\| \Lambda_{M_K^-\mathbb{Z}^+} \right\|_{2,op},
\end{equation*}
where $C_4 = \frac{m'^2}{m}(1+\frac{m'}{m})$.
Since $f^\star \in H_s$ (Assumption $2.1$), and $f^\star \geq m > 0$, we have also $\frac{1}{f^\star} \in H_s$. If we denote $p(k) = \Lambda_{i,i+k}$ the Fourier coefficient of $\frac{1}{f^\star}$, we get
\begin{eqnarray*}
\left\| \Lambda_{M_K^-\mathbb{Z}^+} \right\|_{2,op} & \leq & \left\|\Lambda_{M_K^-\mathbb{Z}^+} \right\|_2 \\
& \leq & \sqrt{\sum_{i \leq - K(N) ; 0 \leq j} p(j-i)^2}\\
& \leq & \sqrt{\sum_{i = K(N)}^\infty \sum_{j= i}^\infty p(j)^2} \\
& \leq & \sqrt{\sum_{i = K(N)}^\infty \frac{\left\|\frac{1}{f^\star}\right\|_{W_{s}}}{i^{2s}}}\\
& \leq &\sqrt{\left\|\frac{1}{f^\star}\right\|_{H_{s}}\frac{1}{K(N)^{s-1}}}.
\end{eqnarray*}
\end{proof}
So that the lemma is proved and the bias is given by
\begin{equation*}
\left\| p_{O_KB_K} - p_{\mathbb{Z}^-B_K} \right\|_{\Gamma} \leq C_4\sqrt{\left\|\frac{1}{f^\star}\right\|_{W_{s}}}\frac{1}{K(N)^{\frac{2s-1}{2}}}.
\end{equation*}
Actually, the rate of convergence for the bias is given by the regularity of the spectral density, since it depends on the coefficients far away from the principal diagonal.\vskip .1in
Now, we prove Lemma \ref{l:var}, which achieves the proof of the theorem.
\begin{proof}{ of Lemma \ref{l:var}}
Recall that $A=\hat{p}_{O_KB_K}^{(N)} - p_{O_KB_K}$. We aim at proving that
$$\int_0^\infty \mathbb{P}\left(\left\| A \right\|_{\Gamma}^4 > t \right) \mathrm{d} t
\leq C_0^4K(N)^4 (\frac{\log(K(N))}{N})^2 + o(K(N)^4 (\frac{\log(K(N))}{N})^2).$$
First,
\begin{eqnarray*}
\left\| A \right\|_{2,op} & = & \left\| (\tilde{\Gamma}_{O_K})^{-1}\hat{\Gamma}_{O_KB_K} - (\Gamma_{O_K})^{-1}\Gamma_{O_KB_K} \right\|_{2,op}
\\ &\leq & \left\| \Gamma_{O_KB_K}\right\|_{2,op} \left\| (\tilde{\Gamma}_{O_K})^{-1}-(\Gamma_{O_K})^{-1} \right\|_{2,op} + \left\|(\tilde{\Gamma}_{O_K})^{-1}\right\|_{2,op} \left\| \hat{\Gamma}_{O_KB_K} - \Gamma_{O_KB_K} \right\|_{2,op}
\\& \leq & \left\| \Gamma_{O_KB_K} \right\|_{2,op} \left\| (\tilde{\Gamma}_{O_K})^{-1} \right\|_{2,op} \left\|(\Gamma_{O_K})^{-1} \right\|_{2,op} \left\| \tilde{\Gamma}_{O_K} - \Gamma_{O_K} \right\|_{2,op}
\\ & & + \left\|(\tilde{\Gamma}_{O_K})^{-1}\right\|_{2,op} \left\|\hat{\Gamma}_{O_KB_K}- \Gamma_{O_KB_K} \right\|_{2,op} .
\end{eqnarray*}
But, we have,
\begin{equation*}\left\| \Gamma_{O_KB_K}\right\|_{2,op} \leq m', \end{equation*} as an extracted operator of $\Gamma$.
\begin{equation*}\left\|(\Gamma_{O_K})^{-1} \right\|_{2,op} \leq \frac{1}{m}, \end{equation*} as the inverse of a principal minor of $\Gamma$.
\begin{equation*} \left\| (\tilde{\Gamma}_{O_K})^{-1} \right\|_{2,op} \leq \frac{4}{m} , \end{equation*} thanks to the regularization.
Furthermore,
\begin{equation*}\left\| \tilde{\Gamma}_{O_K} - \Gamma_{O_K} \right\|_{2,op} \leq K(N)\sup_{p\leq 2 K(N)} \left\{\left|\hat{r}_N(p)-r(p)\right|\right\}+
\left|\hat{\alpha}\right|. \end{equation*}
So the regularization also gives
\begin{equation*}\left\|\hat{\Gamma}_{O_KB_K}- \Gamma_{O_KB_K} \right\|_{2,op} \leq K(N)\left(\sup_{p\leq 2 K(N)} \left\{ \left|\hat{r}_N(p)-r(p)\right| \right\}\right) .\end{equation*}
\begin{equation*} \left| \hat{\alpha}\right| = \left|-\min{\hat{f}_K^N} 1\!\!1_{\min{\hat{f}_K^N} \leq 0}+\frac{m}{4}1\!\!1_{\min{\hat{f}_K^N} \leq \frac{m}{4}} \right|.
\end{equation*}
So, \begin{equation*} \left| \hat{\alpha}\right| \leq (2K(N)+1)\sup_{p\leq 2K(N)} \left\{ \hat{r}_N(p)-r(p)\right\}+\frac{m}{4}1\!\!1_{\min{\hat{f}_K^N} \leq \frac{m}{4}}.
\end{equation*}
For the last inequality, we used the following lemma, proved in the next section.
\begin{lem}\label{lemspec}
The empirical spectral density is such that, for $N$ large enough
\begin{equation*}
\left\| \hat{f}_{K(N)}^N - f^\star \right\|_\infty \leq (2K(N)+1)\sup_{p\leq 2K(N)} \left\{ \hat{r}_N(p)-r(p)\right\} +\frac{m}{4}.
\end{equation*}
\end{lem}
This implies
\begin{equation*}
\left| \min{\hat{f}_K^N} 1\!\!1_{\min{\hat{f}_K^N} \leq 0} \right| \leq (2K(N)+1)\sup_{p\leq 2K(N)} \left\{ \hat{r}_N(p)-r(p)\right\}.
\end{equation*}
So, we obtain,
\begin{eqnarray*}
\left\| A \right\|_{2,op} & \leq & \frac{4m'}{m^2}\left(K(N)\sup_{p\leq 2 K(N)} \left\{\left|\hat{r}_N(p)-r(p)\right|\right\}+ \left|\hat{\alpha}\right|\right) + \frac{4}{m}K(N)\left(\sup_{p\leq 2 K(N)} \left\{\left|\hat{r}_N(p)-r(p)\right|\right\}\right)
\\ & \leq & \left( \frac{6m'}{m^2} +\frac{4}{m} + 2 + \frac{1}{K(N)} \right)K(N)\left(\sup_{p\leq 2 K(N)} \left\{\left|\hat{r}_N(p)-r(p)\right|\right\}\right) + \frac{m'}{m}1\!\!1_{\min{\hat{f}_K^N \leq \frac{m}{4}}} .\end{eqnarray*}
We will use here some other technical lemmas. Their proofs are also postponed to the last section. The first one gives an uniform concentration result on the estimator $\hat{r}_N(p)$:
\begin{lem}\label{lemconc}
Assume that Assumption \ref{a:k} holds. Then, there exists $N_0$ such that, for all $N \geq N_0$, and $x \geq 0$,
\begin{equation*}
\forall p \leq 2K(N), \left|\hat{r}_N(p)-r(p)\right|> 4m' \left(\sqrt{\frac{(\log(K(N))+x)}{N}} + \frac{x}{N}\right),
\end{equation*}
with probability at least $1-e^{-x}$
\end{lem}
For ease of notations, we set $C_0 = 4m'\left( \frac{6m'}{m^2} +\frac{4}{m} + 2\right)$ and $C_3 = \frac{m'}{m}.$
For the computation of the mean, the interval $[0,+\infty[$ will be divided into three parts, where only the first contribution is significant, thanks to the exponential concentration. We will prove that the two other parts are negligible.
We obtain, for all $x\geq 0$
\begin{equation*}
\left\| A \right\|_{2,op} \leq (C_0 + o(1)) K(N)\left(\sqrt{\frac{\log(K(N)+x}{N}}+\frac{x}{N}\right) + C_31\!\!1_{\min{\hat{f}_K^N \leq \frac{m}{4}}},
\end{equation*}
with probability at least $1-e^{-x}$
$\\$
$\\$
Set $t_1 = \left(C_0K(N)\sqrt{\frac{\log(K(N))}{N}}\right)^4$.
For $t \in [0,t_1]$, we use the inequality \begin{equation*} \mathbb{P}\left(\left\| A \right\|_{2,op}^4>t\right) \leq 1 .\end{equation*}
We obtain the first contribution to the integral. This is also the non negligible part.
\begin{equation*}
\int_0^{t_1} \mathbb{P}\left(\left\| A \right\|_{2,op}^4>t\right)\mathrm{d} t = \left(C_0K(N)\sqrt{\frac{\log(K(N))}{N}}\right)^4.
\end{equation*}
$\\$
$\\$
$\\$
Now, set $t_2 = \left(C_0K(N)\sqrt{\frac{\log(K(N))+N}{N}}+C_3\right)^4$.
For $t \in [t_1,t_2]$, we use
\small{\begin{equation*}
\mathbb{P}\left(\left\| A \right\|_{2,op}^4>\sup \left(C_0^4K(N)^4\left(\frac{\log(K(N))+x}{N}\right)^2,C_0^4K(N)^4\left(\frac{x}{N}\right)^4\right)\right) \leq e^{-x} + \mathbb{P}\left(\min{\hat{f}_K^N \leq \frac{m}{4}}\right).
\end{equation*}}
Notice that the last lemma provides
\begin{equation*}
\mathbb{P}\left(2K(N)\sup_{p\leq 2 K(N)} \left\{\left|\hat{r}_N(p)-r(p)\right|\right\} > \frac{m}{2}\right) \leq e^{-\frac{Nm^2}{(64K(N)m')^2}}).
\end{equation*}
Indeed, set $x_0(N) = \frac{Nm^2}{(64K(N)m')^2}$.
One can compute that with probability at least $1-e^{-x_0(N)}$,
\begin{eqnarray*}
\sup_{p\leq 2 K(N)} \left\{\left|\hat{r}_N(p)-r(p)\right|\right\} &\leq &4m'\left(\sqrt{\frac{\log(K(N)) +x_0(N) }{N}} + \frac{x_0(N)}{N} \right)
\\ & \leq & 4m'\left( \sqrt{\frac{\log(K(N))}{N}+ \frac{m^2}{(64K(N)m')^2}} + \frac{m^2}{(64K(N)m')^2} \right)
\\ & \leq & 4m'\left( \sqrt{\frac{\log(K(N))}{N}} + \sqrt{\frac{m^2}{(64K(N)m')^2}} + \frac{m^2}{(64K(N)m')^2} \right)
\\ & \leq & 4m'\left( \sqrt{\frac{\log(K(N))}{N}} + \frac{m}{(64K(N)m')} + \frac{m^2}{(64K(N)m')^2} \right)
\\ & \leq & \frac{m}{8K(N)},
\end{eqnarray*}
for $N$ large enough.
Hence,
\begin{equation*}
\mathbb{P}\left(\min{\hat{f}^N_{K} \leq \frac{m}{4}}\right) \leq e^{-\frac{Nm^2}{(64K(N)m')^2}}.
\end{equation*}
So, we have
\begin{equation*}
\mathbb{P}\left(\left\| A \right\|_{2,op}^4> \max \left(C_0^4K(N)^4\left(\frac{\log(K(N))+x}{N}\right)^2,C_0^4K(N)^4\left(\frac{x}{N}\right)^4\right) \right) \leq e^{-x}+ e^{-\frac{Nm^2}{(64K(N)m')^2}}.
\end{equation*}
Finally, the following lemma (the proof is again postponed in Appendix) will be useful to transform a probability inequality into an $\mathbb{L}^2$ inequality.
\begin{lem}\label{lemfun}
Let $X$ be a nonnegative random variable such that there exists two one to one maps $f_1$ and $f_2$ and a $C>0$ with
\begin{equation*}
\forall x \geq 0, \mathbb{P}\left( X > \sup(f_1(x),f_2(x))\right) \leq e^{-x} + C,
\end{equation*}
then \begin{equation*}
\mathbb{P}\left( X > t \right) \leq e^{-f_1^{-1}(t)}+e^{-f_2^{-1}(t)} + C.
\end{equation*}
\end{lem}
So, thanks to lemma \ref{lemfun}, we have
\begin{equation*}
\mathbb{P}\left(\left\| A \right\|_{2,op}^4>t\right) \leq e^{-N\sqrt{\frac{t}{C_1^4K(N)^4}+\log(K(N))}} + e^{-N\sqrt[4]{\frac{Nt}{C_1^4K(N)^4}}}
+ e^{-\frac{Nm^2}{(64K(N)m')^2}}.
\end{equation*}
Now, we will prove that each term can be neglected.
Integrating by part, we obtain
\begin{eqnarray*}
\int_{t_1}^{t_2} e^{-\sqrt{\frac{t}{C_0^4K(N)^4}+\log(K(N))}}\mathrm{d} t &\leq& \int_{t_1}^{\infty} e^{-N\sqrt{\frac{t}{C_0^4K(N)^4}+\log(K(N))}}\mathrm{d} t
\\ & \leq & \left[\frac{-2\sqrt{t}C_0^2K(N)^2}{N}e^{-N\sqrt{\frac{t}{C_0^4K(N)^4}+\log(K(N))}} \right]_{t_1}^{\infty}
\\ & & + \int_{t_1}^{\infty}\frac{C_0^2K(N)^2}{N\sqrt{t}} e^{-N\sqrt{\frac{t}{C_0^4K(N)^4}+\log(K(N))}} \mathrm{d} t
\\ & \leq & \frac{2\log(K(N))C_0^4K(N)^4}{N^2} + \frac{2C_0^4K(N)^4}{N^2}
\\ & = & o\left( \left(C_0K(N)\sqrt{\frac{\log(K(N))}{N}}\right)^4 \right).
\end{eqnarray*}
Then,
\begin{eqnarray*}
\int_{t_1}^{t_2} e^{-\sqrt[4]{\frac{Nt}{C_0^4K(N)^4}}}\mathrm{d} t &\leq& t_2 e^{-N\sqrt[4]{\frac{t_1}{C_0^4K(N)^4}}}
\\ & \leq & t_2 e^{-\sqrt{N\log(K(N))}}
\\ & = & o\left( \left(C_0K(N)\sqrt{\frac{\log(K(N))}{N}}\right)^4 \right).
\end{eqnarray*}
So that,
\begin{eqnarray*}
\int_{t_1}^{t_2} e^{-x_0(N)}\mathrm{d} t &\leq& t_2 e^{-\frac{Nm^2}{(64K(N)m')^2}}
\\ & = & o\left( \left(C_0K(N)\sqrt{\frac{\log(K(N))}{N}}\right)^4 \right).
\end{eqnarray*}
Leading to
\begin{equation*}
\int_{t_1}^{t_2} \mathbb{P}\left(\left\| A \right\|_{2,op}^4>t\right) \mathrm{d} t= o\left( \left(C_0K(N)\sqrt{\frac{\log(K(N))}{N}}\right)^4 \right),
\end{equation*}
$\\$
$\\$
$\\$
Finally, for $t \in [t_2,+\infty[$, we use
\begin{equation*}
\mathbb{P}\left(\left\| A \right\|_{2,op}^4>\max \left(\left(C_0K(N)\sqrt{\frac{\log(K(N))+x}{N}}+C_3\right)^4,\left(C_0K(N)\frac{x}{N}+C_3\right)^4\right)\right) \leq e^{-x}.
\end{equation*}
Thanks to lemma \ref{lemfun}, we get
\begin{equation*}
\mathbb{P}\left(\left\| A \right\|_{2,op}^4>t\right) \leq e^{-N\left(\frac{\sqrt[4]{t}-C_3}{C_0K(N)}\right)^2+\log(K(N))} + e^{-N\frac{(\sqrt[4]{t}-C_3)}{C_0K(N)}}.
\end{equation*}
So, integrating by part once more, we obtain
\begin{eqnarray*}
\int_{t_2}^{+\infty} e^{-N\left(\frac{\sqrt[4]{t}-C_2}{C_0K(N)}\right)^2+\log(K(N))}\mathrm{d} t &\leq&\int_{\sqrt[4]{t_2}-C_3}^{+\infty} 4(u+C_3)^3 e^{-N\left(\frac{u}{C_0K(N)}\right)^2+\log(K(N))}\mathrm{d} u
\\&\leq & \left[P_1(u,N,K(N)) e^{-N\left(\frac{u}{C_0K(N)}\right)^2+\log(K(N))} \right]_{\sqrt[4]{t_2}-C_3}^{+\infty}
\\ & \leq & P_1(u,N,K(N)) e^{-N}
\\ & \leq & o\left( \left(C_0K(N)\sqrt{\frac{\log(K(N))}{N}}\right)^4 \right).
\end{eqnarray*}
Here, $P_1(u,N,K(N))$ is a polynomial of degree $3$ in $u$ and is rational function in $N$ and $K(n)$.
Furthermore,
\begin{eqnarray*}
\int_{t_2}^{+\infty} e^{-N\frac{(\sqrt[4]{t}-C_3)}{C_0K(N)}} \mathrm{d} t &\leq&\int_{\sqrt[4]{t_2}-C_3}^{+\infty} 4(u+C_3)^3 e^{-N\frac{u}{C_0K(N)}}\mathrm{d} u
\\&\leq & \left[P_2(u,N,K(N)) e^{-N\frac{u}{C_0K(N)}} \right]_{\sqrt[4]{t_2}-C_3}^{+\infty}
\\ & \leq & P_2(u,N,K(N))e^{-\sqrt{N(\log(K(N))+N)}}
\\ & \leq & o\left( \left(C_0K(N)\sqrt{\frac{\log(K(N))}{N}}\right)^4 \right),
\end{eqnarray*}
where $P_2(u,N,K(N))$ is a polynomial of degree $3$ in $u$ and is rational function in $N$ and $K(n)$.
We proved here
\begin{equation*}
\int_0^\infty \mathbb{P}\left(\left\| A \right\|_{2,op}^4 > t \right) \leq C_0^4K(N)^4 (\frac{\log(K(N))}{N})^2 + o(K(N)^4 (\frac{\log(K(N))}{N})^2).
\end{equation*}
This ends the proof.
\end{proof}
\subsection{Technical lemmas}
We prove now the technical lemmas:
\begin{proof}{of Lemma~\ref{lemconc}}
Notice that $\hat{r}^{(N)}(p) = X^T T_N(g_p) X$ with $g_p(t) = \frac{N}{N-p}\cos(pt)$.
We use the following proposition from Comte \cite{comte}. Let $X_1,\cdots,X_n$ be a centered Gaussian stationary sequence and $g$ a bounded function such that $T_n(g)$ is a symmetric non negative matrix. Then the following concentration inequality holds for $Z_n(g) = \frac{1}{n}\left(X^TT_n(g)X - \mathbb{E}[X^TT_n(g)X] \right)$:
\begin{equation*}
\mathbb{P}\left(Z_n(g) \geq 2\left\|f\right\|_\infty\left(\left\|g\right\|_2\sqrt{x} + \left\|u\right\|_\infty x\right)\right)\leq e^{-nx}.
\end{equation*}
By applying this result respectively with $g_p$ and $-g_p$ and we obtain
\begin{equation*}
\mathbb{P}\left( \left|\hat{r}^{(N)}(p) - r(p)\right| > 2m'\frac{N}{N-p}(\sqrt{x}+x)\right) \leq 2e^{-Nx}.
\end{equation*}
or, equivalently,
\begin{equation*}
\left|\hat{r}^{(N)}(p) - r(p)\right| > 2m'\frac{N}{N-p}(\sqrt{\frac{x+\log(K(N))+2\log(2)}{N}}+\frac{x+\log(K(N))+2\log(2)}{N}),
\end{equation*}
with probability lower than $\frac{e^{-x}}{2K(N)}$.
By taking an equivalent, we obtain that there exists $N_0$ such that, for all $N \geq N_0$, for all $p \leq 2K(N)$
\begin{equation*}
\mathbb{P}\left( \left|\hat{r}^{(N)}(p) - r(p)\right| > 4m'\sqrt{\frac{x+\log(K(N))}{N}} + \frac{x}{N}\right) \leq \frac{e^{-x}}{2K(N)}.
\end{equation*}
\end{proof}
\begin{proof}{of Lemma~\ref{lemfun}}
We set $t = \sup(f_1(x),f_2(x))$
If $t = f_1(x)$ then
\begin{equation*}
\mathbb{P}\left( X > t \right) \leq e^{-f_1^{-1}(t)} + C \leq e^{-f_1^{-1}(t)}+e^{-f_2^{-1}(t)} + C.
\end{equation*}
Symmetrically, if $t = f_2(x)$ we have
\begin{equation*}
\mathbb{P}\left( X > t \right) \leq e^{-f_1^{-1}(t)}+e^{-f_2^{-1}(t)} + C.
\end{equation*}
\end{proof}
\begin{proof}{of Lemma~\ref{lemspec}}
It is sufficient to ensure that the bias is small enough. Choose $N_0$ such that
\begin{equation*}
2 \left\|f^\star\right\|_{H_s}K(N)^{-s+1} \leq \frac{m}{4}.
\end{equation*}
Then we use
\begin{eqnarray*}
\left\|\hat{f}^N_{K(N)} - f^\star \right\|_\infty & \leq & \sum_{p = -K(N)}^{K(N)} \left|\hat{r}^N(p) - r(p)\right| + 2\sum_{p > K(N)}\left|r(p)\right|
\\& \leq & (2K(N)+1)\sup_{p\leq 2K(N)} \left\{ \hat{r}_N(p)-r(p)\right\} + 2 \left\|f^\star\right\|_{H_s}K(N)^{-s+1}
\\& \leq & (2K(N)+1)\sup_{p\leq 2K(N)} \left\{ \hat{r}_N(p)-r(p)\right\} +\frac{m}{4}.
\end{eqnarray*}
This ends the proof of the last lemma.
\end{proof}
\textbf{Acknowledgement :}
\normalsize
The authors
wish to thank the anonymous referee for a careful reading of the manuscript and
for providing very useful comments.
\bibliographystyle{plain}
\def$'${$'$}
|
\section{Introduction}
{\bf Invariants}\\[1pt]
Let $\mathcal{O} (V_n)^{\SLtwo}$ denote the algebra of invariants of binary
forms (forms in two variables) of degree $n$ with complex coefficients.
This algebra was extensively studied in the nineteenth century,
and for $n \leq 6$ the structure was clear and a finite basis
(minimal set of generators) was known.
While Cayley (1856)\footnote{See references at the end of this note.}
states that for $n = 7$ there is no such finite basis,
Gordan (1868) proved that $\mathcal{O} (V_n)^{\SLtwo}$ has a finite basis for all $n$.
After initial work by von Gall (1880, 1888),
the degrees of the basic invariants in the cases $n=7$ and $n=8$ were
found by Dixmier \& Lazard (1986) and Shioda (1967), respectively.
Bedratyuk (2007) gave an explicit basis in the case $n=7$.
Here we consider the case $n=9$, and settle a 130-year-old question
by showing that $\mathcal{O} (V_9)^{\SLtwo}$ is generated by 92 basic invariants.
The degrees are given in Proposition \ref{i92}.
The rather large computation needed is discussed in
Section \ref{computation} below.
Earlier work on the case $n=9$ was done by Sylvester \& Franklin (1879)
and by Cr\"{o}ni (2002).
\medskip\noindent
{\bf Systems of parameters}\\
A (homogeneous) {\it system of parameters} for a graded algebra $A$
is an algebraically independent set $S$ of homogeneous elements of $A$
such that $A$ is module-finite over the subalgebra generated by the set $S$.
Hilbert (1893)
showed the existence of a system of parameters
for algebras of invariants, cf.~Proposition \ref{hilbert} below.
In the case $\mathcal{O} (V_9)^{\SLtwo}$ considered here, Dixmier (1985)
proved the following.
\begin{proposition}\label{dixmierparams}
$\mathcal{O}(V_9)^{\SLtwo}$ has a homogeneous system of parameters
of degrees $4$, $8$, $10$, $12$, $12$, $14$, $16$.
\end{proposition}
\noindent
Dixmier was unable to give an explicit such system.
Here we find an explicit system of parameters for $\mathcal{O}(V_9)^{\SLtwo}$
with these degrees (Theorem \ref{explhsop}),
and show the existence of systems of parameters for certain
further sequences of degrees (Proposition \ref{fivehsops}).
\medskip\noindent
{\bf Contents}\\[1pt]
Section 2 gives the Poincar\'e series of the invariant ring.
Its coefficients are the dimensions of the graded parts,
and tell us how many independent invariants we need in each degree.
Section 3 gives the (degrees of) the basic invariants, the main result
of this paper. This result follows by a large computation based on
the knowledge of (the degrees of) a system of parameters.
An explicit such system is given in Section 4, and the proof that
this indeed is a system of parameters follows in Section 5.
Other possible sets of degrees for a system of parameters
are discussed in Section 6, and all such sets for the nonic
are determined in Section 7.
\medskip
\noindent
{\bf Acknowledgements}\\[1pt]
The second author thanks Hanspeter Kraft for the many inspiring
and supporting discussions on the topic of this article.
\section{Invariants and Poincar\'e series}
Let $V_n=\mathbb{C} [x,y]_n$ be the $\SLtwo$-module of binary forms
(homogeneous polynomials in $x$ and $y$) of degree $n$,
on which $\SLtwo$ acts via
$$
g\cdot f(v)=f(g^{-1}v),
$$
for $g\in \SLtwo$, $f\in\mathbb{C} [x,y]$ and $v\in\mathbb{C}^2$. The coordinate ring of
$V_n$, denoted by $\mathcal{O}(V_n)$, is isomorphic to the polynomial ring
$\mathbb{C}[a_0,\ldots,a_n]$.
The group $\SLtwo$ acts on the coordinate ring $\mathcal{O}(V_n)$ via the action
$$
g\cdot j(f)=j(g^{-1}\cdot f),
$$
for $g\in \SLtwo $, $j\in \mathcal{O}(V_n)$ and $f\in V_n$. An {\it invariant\/} of
$V_n$ is an element $j\in \mathcal{O}(V_n)$ such that $g\cdot j=j$ for all
$g \in\SLtwo$. The set of elements of $\mathcal{O}(V_n)$ invariant under the action of
$\SLtwo$ forms the {\it ring of invariants\/} $I := \mathcal{O} (V_n)^{\SLtwo}$.
This ring of invariants $I$ is graded by degree, so that
$I = \oplus_m I_m$, where $I_m$ is the subspace of $I$
consisting of the invariants that are homogeneous of degree $m$.
The Poincar\'e series (or Hilbert series) of $I$ is the series
$P(t) = \sum_m \dimc{I_m} t^m$.
Already Cayley and Sylvester (\cite{Cay,Sy}) knew how to compute
this Poincar\'e series.
For a modern account, see, e.g., Springer \cite{Spr}.
In our case $(n=9)$ the series is given by
$$P(t) = \frac{a(t)}{(1 - t^4)(1 - t^8)(1 - t^{10})(1 - t^{12})^2
(1 - t^{14})(1 - t^{16})}$$
with
\begin{align*}
a(t) = \phantom{} & 1+t^4+5t^8+4t^{10}+17t^{12}+20t^{14}+47t^{16}+61t^{18}+97t^{20}+\\
& 120t^{22}+165t^{24}+189t^{26}+223t^{28}+241t^{30}+254t^{32}+254t^{34}+\\
& 241t^{36}+223t^{38}+189t^{40}+165t^{42}+120t^{44}+97t^{46}+61t^{48}+\\
& 47t^{50}+20t^{52}+17t^{54}+4t^{56}+5t^{58}+t^{62}+t^{66},
\end{align*}
so that
\begin{align*}
P(t) = \phantom{} & 1+2t^4+8t^8+5t^{10}+28t^{12}+27t^{14}+84t^{16}+99t^{18}+217t^{20}+\\
& 273t^{22}+506t^{24}+647t^{26}+1066t^{28}+1367t^{30}+2082t^{32}+2649t^{34}+\\
& 3811t^{36}+4796t^{38}+6612t^{40}+8228t^{42}+10960t^{44}+13483t^{46}+\\
& 17487t^{48}+21274t^{50}+26979t^{52}+32490t^{54}+40443t^{56}+48242t^{58}+\\
& 59107t^{60}+69885t^{62}+84470t^{64}+99074t^{66}+...
\end{align*}
\section{The basic invariants}
A minimal set of homogeneous generators for the algebra $I$ is called
a set of `basic invariants' or basis. Such a set is not unique,
but whenever there is a reference to a basic invariant we mean a member
of such a set, fixed in that context.
Let $J_m$ be the subspace of $I_m$ generated by products of invariants
of smaller degree, that is, in $\bigcup_{j<m} I_j$.
The number of basic invariants of degree $m$ is
$d_m := \dimc{I_m/J_m}$.
\begin{proposition}\label{i92}
The algebra $I$ of invariants for the binary nonic (form of degree 9)
is generated by $92$ invariants. The nonzero numbers $d_m$ of
basic invariants of degree $m$ are
\medskip
\begin{center}\begin{tabular}{c|ccccccccc}
$m$ & $4$ & $8$ & $10$ & $12$ & $14$ & $16$ & $18$ & $20$ & $22$ \\
\hline
$d_m$ & $2$ & $5$ & $5$ & $14$ & $17$ & $21$ & $25$ & $2$ & $1$ \\
\end{tabular}\end{center}
\end{proposition}
Finding a basis for the invariants is a simple but boring procedure:
For each degree $m$, multiply invariants of lower degrees
to see what part of $I_m$ is known already. The Poincar\'e series
tells us how large $I_m$ is, and if the known invariants do not yet span it,
one finds in some way some more invariants, until they do span.
This procedure terminates. Gordan \cite{Go} shows that the algebra $I$
is generated by finitely many of its elements.
Better, we know when to stop.
By Proposition \ref{dixmierparams},
$I$ has a system of parameters of degrees 4, 8, 10, 12, 12, 14, 16.
Let $H$ be the ideal in $I$ generated by such a system of parameters.
Now the Poincar\'e series tells us that if
$a(t) = \sum a_i t^i$ then $\dimc{I_i / (I_i \cap H)} = a_i$,
and, in particular, that $I_i \subseteq H$ for $i > 66$.
This means that $d_m = 0$ for $m > 66$.
We followed this procedure, and found the stated values for $d_m$.
These values agree with those given in \cite{Cr} for $m \le 20$.
The existence of a basic invariant of degree 22 was new.
This `finding more invariants in some way' was done by generating
random bracket monomials\footnote{For the classical concept of
bracket monomial, cf.~\cite{Olver}.}. Explicit bracket monomials
for a set of basic invariants are listed in \cite{aeb}.
Checking whether the invariants known span $I_m$
required computing a basis for vector spaces of dimension at most
$\dimc{I_{66}} = 99074$. That is large but doable.
The entire computation can be done in less than a month.
\subsection{Remarks on the computation}\label{computation}
People usually describe invariants in terms of repeated transvectants.
An advantage of working with bracket monomials is that one can simplify
the computations by substituting small constants for a few variables.
This does not work in the approach using transvectants since there one
needs derivatives with respect to the variables.
Given a candidate set for the basic invariants one wants to find
$\dimc{I_m}$ monomials in these basic invariants that span $I_m$.
Since $\dimc{I_m}$ is known, this amounts to the computation of a rank.
The elements involved are far too large to write down. Instead, the
computation is done lazily, and enough coefficients are written down
to find the desired lower bound on the rank.
Also the integer coefficients are far too large, but it suffices to
consider the reduction mod $p$ for some smallish prime $p$,
say with $100 < p < 255$. Now the rank computation of matrices
with sizes like $100000 \times 160000$ just fits within 16 GB of memory.
The generators took a few TB of disk space. Since this problem is still
too large for the standard computer algebra systems, we implemented
our own software (in C, on a Linux system).
Advantage was taken of the presence of multiple CPUs.
This was about the nonics, the case $n = 9$. The difficulty of this
problem grows very quickly with $n$ (and moreover, this computation
cannot be done in a realistic time when the matrices involved are
much larger than main memory). However, the case $n = 2$ (mod 4) is easier,
and $n = 0$ (mod 4) is much easier than the cases of nearby odd $n$.
And indeed, we were able to do the case of decimics ($n = 10$) as well.
For the time being, the case $n = 12$ is still far too large.
\section{A system of parameters for $\mathcal{O}(V_9)^{\SLtwo}$}
Dixmier \cite{Di1} proved that
the invariant ring of $V_9$ has a system of parameters
of degrees 4, 8, 10, 12, 12, 14, and 16. We compute
an explicit system of parameters of $\mathcal{O}(V_9)^{\SLtwo}$
having these degrees.
A {\it covariant of order m and degree d} of $V_n$ is an
$\SLtwo$-equivariant homogeneous polynomial map
$\phi: V_n \rightarrow V_m$ of degree $d$
such that $\phi (g \cdot f)=g\cdot \phi (f)$ for all
$g\in \SLtwo$ and $f \in V_n$. The invariants of $V_n$ are the covariants
of order 0. The identity map is a covariant of order $n$ and degree 1.
Customarily, one indicates such a covariant $\phi$ by giving its
image of a generic element $f \in V_n$. (In particular, the identity map
is noted $f$.)
Let $V_{m,d}$ be the space of covariants of order $m$ and degree $d$.
The simplest examples of covariants are obtained using
{\em transvectants\/}: given $g \in V_m$ and $h \in V_n$
the expression
$$
(g,h) \mapsto (g,h)_p:=\frac{(m-p)!(n-p)!}{m!n!}
\sum_{i=0}^p (-1)^i \binom{p}{i}
\frac{\partial ^p g}{\partial x^{p-i}\partial y^i}
\frac{\partial ^p h}{\partial x^i \partial y^{p-i}}
$$
defines a linear and $\SLtwo$-equivariant map
$V_m\otimes V_n\rightarrow V_{m+n-2p}$, which is classically called
the {\it p-th transvectant} (\"Uberschiebung) (cf.~\cite{Olver}).
We have $(g,h)_0=gh$ and $(g,g)_{2i+1}=0$ for all integers $i\geq 0$.
These maps are the components of the Clebsch-Gordan isomorphism (for $m\geq n$)
\[
V_m \otimes V_n \iso V_{m+n}\oplus V_{m+n-2}\oplus \ldots \oplus V_{m-n}.
\]
These maps induce maps $V_{m,d} \otimes V_{n,e} \rightarrow V_{m+n-2p,d+e}$.
For $f \in V_9$, consider the following covariants
\begin{alignat*}{2}
l &= (f,f)_8\in V_{2,2},\quad\quad\quad&
r &= (q,f)_6\in V_{3,3},\\
q &= (f,f)_6\in V_{6,2},&
p &= (f,l)_2\in V_{7,3},\\
u &= (f,f)_2\in V_{14,2},&
k_q &= (q,q)_4\in V_{4,4},
\end{alignat*}
and invariants (the suffix indicates the degree)
\begin{alignat*}{4}
j&_4 &\,=\,& (l,l)_2,&
B&_8 &\,=\,& (q,r^2)_6,\\
j&_{12} &\,=\,& ((k_q,k_q)_2,k_q)_4,\quad&
B&_{12} &\,=\,& ((p,p)_4,l^3)_6,\\
j&_{14} &\,=\,& (q,(r^3,r)_3)_6,&
D&_{10} &\,=\,& ((((u,u)_{10},f)_6,(q,f)_2)_5,q)_6,\\
j&_{16} &\,=\,& ((p,p)_2,l^5)_{10}.
\end{alignat*}
\begin{theorem}\label{explhsop}
The seven invariants $j_4$, $B_8$, $D_{10}$, $B_{12}$, $j_{12}$,
$j_{14}$, $j_{16}$ form a homogeneous system of parameters for
the ring $\mathcal{O}(V_9)^{\SLtwo}$ of invariants of the binary nonic.
\end{theorem}
This is proved below (\S\ref{theproof}) by invoking Hilbert's
characterization of homogeneous systems of parameters as sets
that define the nullcone.
\section{The nullcone}
The {\em nullcone} of $V_n$, denoted $\mathcal{N}(V_n)$, is the set of binary forms
of degree $n$ on which all invariants of positive degree vanish.
It turns out (\cite{Hi2}) that this is precisely the set of binary forms
of degree $n$ with a root of multiplicity $>\frac{n}{2}$.
The elements of $\mathcal{N}(V_n)$ are called {\em nullforms}.
The nullcone $\mathcal{N}(V_n\oplus V_m)$ is
the set of pairs $(g,h)\in V_n\oplus V_m$ such that $g$ and $h$ have a
common root of multiplicity $>\frac{n}{2}$ in $g$ and of multiplicity
$>\frac{m}{2}$ in $h$. (In this note, this result can be taken as the
definition of the symbol $\mathcal{N}(V_n\oplus V_m)$.)
We have the following result, due to Hilbert \cite{Hi2},
formulated for the particular case of binary forms:
\begin{proposition} \label{hilbert}
For $n \ge 3$, consider $i_1,\ldots ,i_{n-2}\in \mathcal{O}(V_n)^{\SLtwo}$
homogeneous non-constant invariants of $V_n$.
The following two conditions are equivalent:
\begin{itemize}
\item[(i)] $\mathcal{N}(V_n)=\mathcal{V}(i_1,\ldots ,i_{n-2})$,
\item[(ii)] $\{i_1,\ldots ,i_{n-2}\}$ is a homogeneous system of parameters
of $\mathcal{O}(V_n)^{\SLtwo}$.
\end{itemize}
\end{proposition}
(Here $\mathcal{V}(J)$ stands for the vanishing locus of $J$.)\\
In other words, if $i_1,\ldots ,i_{n-2}$ are homogeneous invariants
such that $\mathcal{N}(V_n)=\mathcal{V}(i_1,\ldots ,i_{n-2})$, then the ring
$\mathcal{O}(V_n)^{\SLtwo}$ is a finitely generated module over
$\mathbb{C}[i_1,\ldots ,i_{n-2}]$. But invariant rings of binary forms are
Cohen-Macaulay (\cite{HoRo}), which implies that $\mathcal{O}(V_n)^{\SLtwo}$
is a free $\mathbb{C}[i_1,\ldots ,i_{n-2}]$-module. Hence the description of
the algebra of invariants of $V_n$ is partly reduced to finding a
system of parameters of $\mathcal{O}(V_n)^{\SLtwo}$.
We prove Theorem \ref{explhsop} by first finding a defining set
for the nullcone that is still too large, and then showing
that some elements are superfluous.
\medskip
\medskip
We need information on the invariants of $V_n$ for $n = 2,\,3,\,6,\,7$:
\begin{lemma}\label{hsops}
The following are systems of parameters of
$\mathcal{O}(V_n)^{\SLtwo}$ for $n = 2,\,3,\,6,\,7$.
\begin{itemize}
\item[(i)] If $n=2$: $(f,f)_2$ of degree $2$.
\item[(ii)] If $n=3$: $((f,f)_2,(f,f)_2)_2$ of degree $4$.
\item[(iii)] If $n=6$: $(f,f)_6$, $(k,k)_4$, $((k,k)_2,k)_4$,
and $(m^2,(k,k)_2)_4$ of degrees $2$, $4$, $6$, and $10$,
where $k = (f,f)_4$ and $m = (f,k)_4$.
\item[(iv)] If $n=7$: $(l,l)_2$, $((p,p)_4,l)_2$,
$((k_q,k_q)_2,k_q)_4$, $((p,p)_2,l^3)_6$,
$(m_q\supr{2},(k_q,k_q)_2)_4$
of degrees $4$, $8$, $12$, $12$, and $20$,
where $l = (f,f)_6$,
$p = (f,l)_2$,
$q = (f,f)_4$,
$k_q = (q,q)_4$,
$m_q = (q,k_q)_4$.
\end{itemize}
\end{lemma}
\begin{proof}
This is classical for $n=2$, $3$, $6$, see, e.g., \cite{Clebsch,GrYo,Schur}.
Systems of parameters for $n=7$ were given by Dixmier \cite{Di0}
and Bedratyuk \cite{Be}.
The above system was constructed by the second author (unpublished).
That it is a system of parameters
can be easily verified using the methods of this section.
\end{proof}
\begin{lemma}\label{jerzy} {\rm (Weyman \cite{We1})}
Let $f\in V_d$. If $d > 4k-4$ and all $(f,f)_{2k}$, $(f,f)_{2k+2}$, ...
vanish, then $f$ has a root of multiplicity $d-k+1$. If $d = 4k-4$ and
$((f,f)_{2k-2},f)_d$, $(f,f)_{2k}$, $(f,f)_{2k+2}$, ... vanish,
then $f$ has a root of multiplicity $d-k+1$. \qed
\end{lemma}
\begin{lemma} \label{nullform}
Let $f \in V_9$ and consider its covariants $l=(f,f)_8$, $q=(f,f)_6$,
$p=(f,l)_2$, and $r=(f,q)_6$.
\begin{itemize}
\item[(i)] If $l \neq 0$ and $(l,p)\in \mathcal{N}(V_2\oplus V_7)$, then $f$ has
a root of multiplicity $5$.
\item[(ii)] If $l=0$, $q\neq 0$ and $(q,r)\in \mathcal{N}(V_6\oplus V_3)$ then $f$ has
a root of multiplicity $6$.
\item[(iii)] If $l=q=0$, then $f$ has a root of multiplicity $7$.
\end{itemize}
\end{lemma}
\begin{proof}
Let $f = \sum_{i=0}^9 \binom{9}{i} a_i x^{9-i}y^i$.
\medskip\noindent
(i). From $(l,p)\in \mathcal{N}(V_2\oplus V_7)$ it follows that both $l$ and $p$
are nullforms and have a common root of multiplicity 2 in $l$ and
4 in $p$. Without loss of generality we suppose $l=x^2$. Then:
$$
p=(f,x^2)_2=\frac{1}{72}\sum_{i=2}^{9}\binom{9}{i}i(i-1) a_ix^{9-i}y^{i-2},
$$
and $x^4$ must divide $p$, which implies $a_6=a_7=a_8=a_9=0$.
Now
$$
l=(f,f)_8=70 a_5\supr{2} y^2+ 28 a_4 a_5 xy + (70 a_4\supr{2} - 112 a_3 a_5)x^2,
$$
and as we suppose $l=x^2$ we also obtain $a_5=0$ and then it follows that
$x^5 \mid f$, so $f$ will have a root of multiplicity 5.
\medskip\noindent
(ii). From $(q,r)\in \mathcal{N}(V_6\oplus V_3)$ it follows that both $q$ and $r$
are nullforms and have a common root of multiplicity 4 in $q$ and 2 in $r$.
Without loss of generality we consider the
following 3 cases: $q=x^6$, $q=x^5y$, and $q=x^4y(x+y)$.
\noindent
Case 1: $q=x^6$. Then
$$
r=(f,x^6)_6=a_9y^3+ 3 a_8xy^2+3 a_7x^2y+ a_6x^3,
$$
and $x^2$ must divide $r$.
We obtain $a_9=a_8=0$ and substitute that in $q$ and $l$:
\begin{align*}
q=(f,f)_6=\, &(-20 a_6\supr{2} + 30 a_5 a_7)y^6 + (-30 a_5 a_6 + 54 a_4 a_7)xy^5+\\
& (-90 a_5\supr{2} + 114 a_4 a_6 - 12 a_3 a_7)x^2y^4+\\
& (-72 a_4 a_5 + 124 a_3 a_6 - 60 a_2 a_7)x^3y^3+\\
& (-90 a_4\supr{2} + 114 a_3a_5 - 12 a_2 a_6 - 18 a_1 a_7)x^4y^2+\\
& (-30 a_3 a_4 + 54 a_2 a_5 - 30 a_1 a_6 + 6 a_0 a_7)x^5y+\\
&(-20 a_3\supr{2} + 30 a_2 a_4 - 12 a_1 a_5 + 2 a_0 a_6)x^6,\\
l=(f,f)_8=\,&(70 a_5\supr{2} - 112 a_4 a_6 + 56 a_3 a_7)y^2+\\
& (28 a_4 a_5 - 56 a_3 a_6 + 40 a_2 a_7)xy+\\
& (70 a_4\supr{2} - 112 a_3 a_5 + 56 a_2 a_6 - 16 a_1 a_7)x^2.
\end{align*}
Since we suppose $q=x^6$ and $l=0$, the coefficients of $x^i y^{6-i}$ in $q$
and of $x^j y^{2-j}$ in $l$ are 0 for $0 \le i \le 5$ and $0 \le j \le 2$.
If $a_7=0$ then it follows that $a_6=a_5=a_4=0$ and then $x^6 \mid f$,
so $f$ will have a root of multiplicity 6.
If $a_7 \neq 0$ then
\begin{align*}
& a_5=\frac{2 a_6\supr{2}}{3 a_7}, ~~
a_4=\frac{10 a_6\supr{3}}{27 a_7\supr{2}}, ~~
a_3=\frac{5 a_6\supr{4}}{27 a_7\supr{3}},\\
& a_2=\frac{7 a_6\supr{5}}{81 a_7\supr{4}}, ~~
a_1=\frac{28 a_6\supr{6}}{729 a_7\supr{5}}, ~~
a_0=\frac{4 a_6\supr{7}}{243 a_7\supr{6}},
\end{align*}
but then we have $q=0$, contrary to the assumption.
\medskip\noindent
Case 2: $q=x^5y$. Then
\[r=(f,x^5y)_6=-a_8y^3-3 a_7 xy^2-3 a_6 x^2y -a_5 x^3\]
and $x^2$ must divide $r$. We obtain $a_8=a_7=0$ and substitute
this in $q$ and $l$:
\begin{align*}
q=(f,f)_6=\,&(-20 a_6\supr{2} + 2 a_3 a_9)y^6 + (-30 a_5 a_6 + 6 a_2 a_9)xy^5+\\
& (-90 a_5\supr{2} + 114 a_4 a_6 + 6 a_1 a_9)x^2y^4+ \cdots + \\
& (-90 a_4\supr{2} + 114 a_3a_5 - 12 a_2 a_6)x^4y^2+\\
& (-30 a_3 a_4 + 54 a_2 a_5 - 30 a_1 a_6 )x^5y+ \cdots \\
l=(f,f)_8=\,&(70 a_5\supr{2} - 112 a_4 a_6 + 2 a_1 a_9)y^2+ \cdots
\end{align*}
Since we supposed $q=x^5y$ and $l=0$, the coefficient $c$ of $y^2$ in $l$,
and the coefficients $d_i$ of $x^i y^{6-i}$ in $q$ vanish for
$0 \le i \le 4$, while $d_5 \ne 0$. Now
$$5d_5a_9 = -75a_4d_0 + 45a_5d_1 - a_6(9c + 22d_2) = 0$$
so that $a_9 = 0$, and then also $a_6=a_5=a_4=0$, $d_5 = 0$,
contradicting $d_5 \ne 0$.
\medskip\noindent
Case 3: $q=x^4y(x+y)$. Then:
$$
r=(f,x^4y(x+y))_6=(a_7 - a_8)y^3+3(a_6 -a_7) xy^2+
3(a_5 - a_6)x^2y+(a_4 - a_5)x^3
$$
and $x^2$ must divide $r$. We obtain $a_8=a_7=a_6$ which we
replace in $q$ and $l$:
\begin{align*}
q=(f,f)_6=\,&-2 (6 a_4 a_6 - 15 a_5 a_6 + 10 a_6\supr{2} - a_3 a_9)y^6-\\
& -6 (5 a_3 a_6 - 9 a_4 a_6 + 5 a_5 a_6 - a_2 a_9)xy^5-\\
& -6 (15 a_5\supr{2} + 3 a_2 a_6 + 2 a_3 a_6 - 19 a_4 a_6 - a_1 a_9)x^2y^4-\\
& -2 (36 a_4 a_5 - 3 a_1 a_6 + 30 a_2 a_6 - 62 a_3 a_6 - a_0 a_9)x^3y^3-\\
& -6 (15 a_4\supr{2} -19 a_3 a_5 - a_0 a_6 + 3 a_1 a_6 +2 a_2 a_6)x^4y^2-\\
& -6 (5 a_3 a_4 - 9 a_2 a_5 - a_0 a_6 + 5 a_1 a_6)x^5y-\\
& -2 (10 a_3\supr{2} - 15 a_2 a_4 + 6 a_1 a_5 -a_0 a_6)x^6,\\
l=(f,f)_8=\,& 2 (35 a_5\supr{2} - 8 a_2 a_6 + 28 a_3 a_6 - 56 a_4 a_6 + a_1 a_9)y^2+\\
& 2 (14 a_4 a_5 - 7 a_1 a_6 + 20 a_2 a_6 - 28 a_3 a_6 + a_0 a_9)xy+\\
& 2 (35 a_4\supr{2} - 56 a_3 a_5 +a_0 a_6 - 8 a_1 a_6 + 28 a_2 a_6)x^2.
\end{align*}
As we supposed $q=x^4y(x+y)$ and $l=0$, the coefficients of $y^6$, $xy^5$,
$x^2y^4$, $x^3y^3$, $x^6$ in $q$ and all coefficients of $l$ must vanish.
We denote by $I$ the ideal generated by these coefficients.
Also, we denote by $p_1$, $p_2$ the coefficients of $x^4y^2$ and $x^5y$ in $q$:
\begin{align*}
& p_1=15 a_4\supr{2} -19 a_3 a_5 - a_0 a_6 + 3 a_1 a_6 +2 a_2 a_6,\\
& p_2=5 a_3 a_4 - 9 a_2 a_5 - a_0 a_6 + 5 a_1 a_6.
\end{align*}
A Gr\"obner basis computation shows that $p_1^4$, $p_2\supr{2} \in I$
so that $p_1$ and $p_2$ vanish, contradicting the assumption $q=x^4y(x+y)$.
\medskip\noindent
(iii). This is a consequence of Lemma \ref{jerzy}.
\end{proof}
\begin{lemma} \label{lpv9}
Let $g \in V_2$ and $h \in V_7$ be two non-zero binary forms.
If both $g$ and $h$ are nullforms and if
\[((h,h)_6,g)_2=((h,h)_4,g^3)_6=((h,h)_2,g^5)_{10}=(h^2,g^7)_{14}=0,\]
then $(g,h)\in \mathcal{N}(V_2\oplus V_7)$.
\end{lemma}
\begin{proof}
Suppose that $(g,h) \notin \mathcal{N}(V_2\oplus V_7)$. This means that
$g$ and $h$ have no common root which has multiplicity 2 in $g$
and multiplicity 4 in $h$. Without loss of generality we suppose
\begin{align*}
& g=x^2,\\
& h=y^4(b_1 x^3+b_2x^2 y+b_3x y^2+b_4y^3).
\end{align*}
We have then
\begin{align*}
0&=((h,h)_6,g)_2=-\frac{4}{245} b_1\supr{2} ,\\
0&=((h,h)_4,g^3)_6=\frac{2}{735}(5 b_2\supr{2} - 12 b_1 b_3),\\
0&=((h,h)_2,g^5)_{10}=-\frac{2}{147} (3 b_3\supr{2} - 7 b_2 b_4),\\
0&=(h^2,g^7)_{14}=b_4\supr{2}
\end{align*}
and it follows that $b_1=b_2=b_3=b_4=0$, which implies $h=0$.
This contradicts the assumption that $h \ne 0$.
\end{proof}
\begin{lemma} \label{qrv9} Let $g \in V_6$, $h\in V_3$ be two non-zero
binary forms. If both $g$ and $h$ are nullforms and if
$$
((g^2,g)_6,h^2)_6=((\!(g,g)_2,g)_1,h^4)_{12}=(g,h^2)_6=
(g,(h,h)_2\supr{3})_6=(g,(h^3,h)_3)_6\!=0
$$
then $(g,h)\in \mathcal{N}(V_6\oplus V_3)$.
\end{lemma}
\begin{proof}
Suppose that $(g,h) \notin \mathcal{N}(V_6\oplus V_3)$. This means that
$g$ and $h$ have no common root which has multiplicity 4 in $g$
and multiplicity 2 in $h$.
Without loss of generality we consider two cases:
\begin{align*}
& g=x^4(b_1 x^2+ b_2xy+b_3 y^2),\\
& h=y^3
\end{align*}
and
\begin{align*}
& g=x^4(b_1 x^2+ b_2xy+b_3 y^2),\\
& h=xy^2.
\end{align*}
\noindent
Case 1: $h=y^3$. Then we have:
\begin{align*}
0&=((g^2,g)_6,h^2)_6=\frac{1}{495}b_3\supr{3},\\
0&=(((g,g)_2,g)_1,h^4)_{12}=-\frac{1}{540}b_2(5 b_2\supr{2} - 18 b_1 b_3),\\
0&=(g,h^2)_6=b_1
\end{align*}
and it follows that $b_1=b_2=b_3=0$, which implies $g=0$,
contradicting the assumption $g \ne 0$.
\noindent
Case 2: $h=x y^2$. Then we have:
\begin{align*}
0&=(g,h^2)_6=\frac{1}{15}b_3,\\
0&=(g,(h,h)_2\supr{3})_6=-\frac{8}{729}b_1,\\
0&=(g,(h^3,h)_3)_6=\frac{1}{84} b_2
\end{align*}
and it follows that $b_1=b_2=b_3=0$, which implies $g=0$,
contradicting the assumption $g \ne 0$.
\end{proof}
\subsection{Proof of Theorem \ref{explhsop}}\label{theproof}
We consider the following covariants of $V_9$:
{\begin{alignat*}{6}
l&_p&\,=\,&(p,p)_6&\,&\in V_{2,6},\quad&
q&_p&\,=\,&(p,p)_4&&\in V_{6,6},\\
p&_p&\,=\,&(p,l_p)_2&&\in V_{5,9},&
k&_{qp}&\,=\,&(q_p,q_p)_4&&\in V_{4,12},\\
k&_q&\,=\,&(q,q)_4&&\in V_{4,4},&
m&_{qp}&\,=\,&(q_p,k_{qp})_4&&\in V_{2,18},\\
m&_q&\,=\,&(q,k_q)_4&&\in V_{2,6},
\end{alignat*}}
and the following invariants of $V_9$:
\begin{alignat*}{4}
j&_4&\,=\,&(l,l)_2,&
A&_4&\,=\,&(q,q)_6,\\
j&_8&\,=\,&(k_q,k_q)_4,&
A&_8&\,=\,&((p,p)_6,l)_2,\\
j&_{12}&\,=\,&((k_q,k_q)_2,k_q)_4,&
A&_{12}&\,=\,&(l_p,l_p)_2,\\
j&_{14}&\,=\,&(q,(r^3,r)_3)_6,&
A&_{20}&\,=\,&(p^2,l^7)_{14},\\
j&_{16}&\,=\,&((p,p)_2,l^5)_{10},&
A&_{36}&\,=\,&((p_p,p_p)_2,l_p\supr{3})_6,\\
j&_{18}&\,=\,&(((q,q)_2,q)_1,r^4)_{12},\quad\quad&
B&_8&\,=\,&(q,r^2)_6,\\
j&_{20}&\,=\,&(m_q\supr{2},(k_q,k_q)_2)_4,&
B&_{12}&\,=\,&((p,p)_4,l^3)_6,\\
j&_{24}&\,=\,&((p_p,p_p)_4,l_p)_2,&
B&_{20}&\,=\,&(q,(r,r)_2\supr{3})_6,\\
j&_{36}&\,=\,&((k_{qp},k_{qp})_2,k_{qp})_4,&
C&_{12}&\,=\,&((r,r)_2,(r,r)_2)_2,\\
j&_{60}&\,=\,&(m_{qp}\supr{2},(k_{qp},k_{qp})_2)_4,&
D&_{12}&\,=\,&((q^2,q)_6,r^2)_6.
\end{alignat*}
Apply Lemma \ref{hsops} to $l \in V_2$, $r \in V_3$, $q \in V_6$
and $p \in V_7$.
It follows that if $j_4=0$ then $l$ is a nullform,
if $C_{12}=0$ then $r$ is a nullform,
if $A_4=j_8=j_{12}=j_{20}=0$ then $q$ is a nullform, and
if $A_{12}=j_{24}=j_{36}=A_{36}=j_{60}=0$, then $p$ is a nullform.
If we combine this information with Lemma \ref{nullform}, Lemma \ref{lpv9}
and Lemma \ref{qrv9} we obtain that
\begin{align*}
\mathcal{N}(V_9)&=\mathcal{V}(j_4,A_4,j_8,A_8,B_8,j_{12},A_{12},B_{12},C_{12},D_{12},j_{14},j_{16},j_{18},j_{20},A_{20},B_{20},\\ &j_{24},j_{36},A_{36},j_{60}).
\end{align*}
This can be improved to the following result:
\begin{proposition} \label{nullsmall}
The nullcone $\mathcal{N}(V_9)$ is the zero set of the following invariants:
\[\mathcal{N}(V_9)=\mathcal{V} (j_4,A_4,j_8,A_8,j_{12},B_{12},j_{14},j_{16},j_{20},A_{20}). \]
\end{proposition}
\begin{proof}
If $j_4=0$ then $l$ is a nullform.
\medskip\noindent
Case 1: $l=0$.
\noindent
If $A_4=j_8=j_{12}=j_{20}=0$ then $q$ is a nullform. Without loss of
generality we suppose $x^4 \mid q$. Modulo the ideal generated by the
coefficients of $l$ and the coefficients of $x^3y^3,x^2y^4,xy^5,y^6$
in $q$ we have
\[B_8=C_{12}=D_{12}=j_{18}=B_{20}=0.\]
(This was an easy computation in Mathematica.)
From Lemma \ref{nullform} it follows then that if $l=0$ and
\[A_4=j_8=j_{12}=j_{14}=j_{20}=0,\]
then $f$ is a nullform.
\medskip\noindent
Case 2: $l=x^2$ (without loss of generality).
\noindent
Here we have:
\begin{align*}
A_{20}&=a_9\supr{2},\\
j_{16}&=-2 (a_8\supr{2} - a_7 a_9),\\
B_{12}&=2 (3 a_7\supr{2} - 4 a_6 a_8 + a_5 a_9),\\
A_8&=-2 (10 a_6\supr{2} - 15 a_5 a_7 + 6 a_4 a_8 - a_3 a_9).
\end{align*}
Hence if $A_{20}=j_{16}=B_{12}=A_8=0$,
then $a_9=a_8=a_7=a_6=0$,
and if we combine this with $l=x^2$ we get $a_5=0$ too,
hence $f$ is a nullform.
\end{proof}
But we are still not in the position to apply Proposition \ref{hilbert}.
For that we have to refine our result even more.
We introduce the covariant $s = (f,f)_4 \in V_{10,2}$ and
the following invariants:
\begin{alignat*}{2}
C&_8&\,=\,\,&((q,q)_4,l^2)_4,\\
D&_8&\,=\,\,&((q,q)_4,(q,s)_6)_4,\\
j&_{10}&\,=\,\,&((p,(f,q)_6)_3,(q,q)_4)_4,\\
A&_{10}&\,=\,\,&((p,(f,q)_6)_3,l^2)_4,\\
B&_{10}&\,=\,\,&(((f,q)_6,(f,s)_6)_3,(s,s)_8)_4,\\
C&_{10}&\,=\,\,&((((s,s)_6,f)_6,(l,f)_2)_3,q)_6,\\
D&_{10}&\,=\,\,&((((u,u)_{10},f)_6,(q,f)_2)_5,q)_6.
\end{alignat*}
The invariants $j_8$, $A_8$, $B_8$, $C_8$, and $D_8$ are linearly independent
and together with $j_4\supr{2}$, $A_4\supr{2}$, $A_4j_4$ generate the vector
space of invariants of degree 8 which is of dimension 8.
(This can be seen, e.g., by a small computation in Mathematica.)
In a similar way it can be seen that the vector space of invariants
of degree 10 is generated by $j_{10}$, $A_{10}$, $B_{10}$, $C_{10}$,
and $D_{10}$.
Using invariants of degree $\leq 16$ we built a list of 219 monomials
of degree 20, each of them dividing one of the invariants $j_4$,
$A_4$, $j_8$, $A_8$, $B_8$, $C_8$, $D_8$, $C_{10}$ or $D_{10}$, to
which we added
\begin{align*}
B&_{20}=((r,r)_2\supr{3},q)_6 ,\\
C&_{20}=(((r^3,r)_3,q)_4,((f,u)_8,(f,s)_8)_3)_4 .
\end{align*}
Let $I$ be the ring of invariants, and $I_i$ its $i$-th graded part.
We evaluated the monomials at $\dimc{I_{20}} = 217$ random points in $V_9$,
giving as result a matrix of (full) rank 217.
Adding $j_{20}$, $A_{20}$, $j_{10}^2$, $A_{10}^2$,
and $B_{10}^2$ to the list of monomials and repeating
the evaluation step gave (of course) again matrices of rank 217.
From the nullspaces of these matrices we obtained the relations
$$
j_{20},A_{20},j_{10}^2,A_{10}^2,B_{10}^2 \in
(j_4,A_4,j_8,A_8,B_8,C_8,D_8,C_{10},D_{10})
$$
(that is, $B_{20}$ and $C_{20}$ are not needed to span the elements
mentioned).
Using invariants of degree $\leq 20$ we built a list of 3561 monomials
of degree 32, each of them dividing one of the invariants $j_4$,
$B_8$, $D_8$, $C_{10}$, $D_{10}$, $j_{12}$, $B_{12}$, $j_{14}$, or
$j_{16}$. We evaluated the monomials at $\dimc{I_{32}} = 2082$
random points in $V_9$, and this resulted in a matrix of rank 2082.
The rank computations were made modulo 32003, but as we obtained
the maximal rank, these monomials must generate $I_{32}$. It follows that
$$
j_8,A_8,C_8,A_4 \in
\sqrt {(j_4,B_8,D_8,C_{10},D_{10},j_{12},B_{12},j_{14},j_{16})},
$$
and then, combining it with Proposition \ref{nullsmall}, we get
\[\mathcal{N}(V_9)=\mathcal{V} (j_4,B_8,D_8,C_{10},D_{10},j_{12},B_{12},j_{14},j_{16}).\]
In the same way one can show that
\[\mathcal{N}(V_9)=\mathcal{V} (A_4,B_8,D_8,C_{10},D_{10},j_{12},B_{12},j_{14},j_{16}).\]
It remains to remove two elements from one of these two sets of generators.
Since this did not seem easy to do by hand, we reverted to the boring approach,
as follows.
Let $H = (j_4,B_8,D_{10},j_{12},B_{12},j_{14},j_{16})$.
We computed $\dimc{I_i \cap H}$ for $i \le 60$
and found $\dimc{I_{60} \cap H} = 59107 = \dimc{I_{60}}$,
so that $I_{60} \subseteq H$.
But then $H$ contains powers of all invariants of degrees 4, 10, 20,
so that in particular $A_4, C_{10} \in \sqrt {H}$.
Now let $H' = (j_4,A_4,B_8,D_{10},j_{12},B_{12},j_{14},j_{16})$.
We computed $\dimc{I_i \cap H'}$ for $i \le 40$ and found
$\dimc{I_{40} \cap H'} = 6612 = \dimc{I_{40}}$, so that
$I_{40} \subseteq H'$.
But then $H'$ contains powers of all invariants of degree 8,
so that in particular $D_8 \in \sqrt {H'}$.
But then $\sqrt {H} = \sqrt {H'} = I$.
Thus,
\[\mathcal{N}(V_9)=\mathcal{V} (j_4,B_8,D_{10},j_{12},B_{12},j_{14},j_{16}),\]
and from Proposition \ref{hilbert} it follows that
$\{j_4,B_8,D_{10},j_{12},B_{12},j_{14},j_{16}\}$
is a homogeneous system of parameters of $I$. \qed
\medskip\noindent {\bf Remark}
As a consequence of this result, the proof of Proposition \ref{i92}
no longer requires Proposition \ref{dixmierparams}.
On the other hand, since the end of the proof of the theorem needs
computer work anyway, one can avoid all discussion of the nullcone
following Proposition \ref{hilbert} and show directly that
$\sqrt {H} = I$. From Proposition \ref{i92} we learn that $I$ is
generated by invariants of degrees 4, 8, 10, 12, 14, 16, 18, 20, 22.
Now one can verify that $I_m \subseteq H'$ for $36 \le m \le 44$ and $m = 48$,
hence $\sqrt {H} = \sqrt {H'} = I$.
Thus, Theorem \ref{explhsop} also follows from Dixmier \cite{Di1}
and computer work.
\section{The degrees in a system of parameters}
We give some restrictions on the set of degrees
for the forms in a homogeneous system of parameters (hsop).
Assume $n \ge 3$.
\begin{lemma} Fix integers $j$, $t$ with $t > 0$.
If an invariant of degree $d$ is nonzero on a form $\sum a_i x^{n-i} y^i$
with the property that all nonzero $a_i$ have $i \equiv j$ (mod $t$),
then $d(n-2j)/2 \equiv 0$ (mod $t$).
\end{lemma}
\begin{proof}
For an invariant of degree $d$ with nonzero term $\prod a_i^{m_i}$ we have
$\sum m_i = d$ and $\sum i m_i = nd/2$.
If $i \equiv j$ (mod $t$) when $a_i \ne 0$, then
$nd/2 = \sum i m_i \equiv j \sum m_i = jd$ (mod $t$).
\end{proof}
\begin{lemma} Fix integers $j$, $t$ with $t > 1$ and $0 \le j \le n$.
Among the degrees $d$ of a hsop, at least
$\lfloor (n-j)/t \rfloor$ satisfy $d(n-2j)/2 \equiv 0$ (mod $t$).
\end{lemma}
\begin{proof}
We may suppose $0 \le j < t$.
There are $1 + \lfloor (n-j)/t \rfloor$ coefficients $a_i$
with $i \equiv j$ (mod $t$), so that the subvariety of $V_n$
defined by $a_i = 0$ for $i \not\equiv j$ (mod $t$)
has dimension at least $\lfloor (n-j)/t \rfloor$.
If this is zero, there is nothing to prove. Otherwise,
adding the conditions that the elements of a hsop vanish
reduces this subvariety to a subset of the nullcone.
But the part of this subvariety defined by
$a_i \ne 0$ for $i \equiv j$ (mod $t$) is disjoint from the nullcone.
Indeed, consider the form $a_j x^{n-j}y^j + \cdots + a_{n-k} x^ky^{n-k}$,
where $0 \le j < t$ and $0 \le k < t$ and $j+k \le n-t$ and $a_j$,
$a_{n-k}$ are nonzero but $a_i = 0$ when $i \not\equiv j$ (mod $t$).
The nullcone consists of the forms with a zero of multiplicity
more than $n/2$, but $x=0$ and $y=0$ are zeros of multiplicity $j$ and $k$,
respectively, and if e.g. $j > n/2$, then $k \le n-t-j < n-2j < 0$,
impossible. This means that a zero of multiplicity more than $n/2$
also is a zero of $a_jx^{n-j-k} + \cdots + a_{n-k}$, but this is a
polynomial in $x^t$ and has no roots of multiplicity more than $n/t$.
\end{proof}
\begin{proposition}
Let $t$ be an integer with $t > 1$.
(i) If $n$ is odd, and $j$ is minimal such that $0 \le j \le n$ and
$(n-2j,t) = 1$, then among the degrees of any hsop at least
$\lfloor (n-j)/t \rfloor$ are divisible by $2t$.
(ii) If $n$ is even, and $j$ is minimal with $0 \le j \le \frac{1}{2}n$ and
$(\frac{1}{2}n-j,t) = 1$, then among the degrees of any hsop at least
$\lfloor (n-j)/t \rfloor$ are divisible by $t$. \qed
\end{proposition}
\begin{corollary}
Let $t = p^e$ be a power of a prime $p$, where $e > 0$.
(i) Suppose $p=2$. If $n$ is odd, then among the degrees of any hsop
at least $\lfloor n/t \rfloor$ are divisible by $2t$.
If $n/2$ is odd, then at least $\lfloor n/t \rfloor$ degrees
are divisible by $t$. If $4|n$, then at least $\lfloor (n-2)/t \rfloor$
degrees are divisible by~$t$.
(ii) Suppose $p>2$. Among the degrees of any hsop at least
$\lfloor (n-1)/t \rfloor$ are divisible by $t$. \qed
\end{corollary}
For example, there exist homogeneous systems of parameters
with degree sequences 4 $(n=3)$; 2, 3 $(n=4)$; 4, 8, 12 $(n=5)$;
2, 4, 6, 10 $(n=6)$; 4, 8, 12, 12, 20 and 4, 8, 8, 12, 30 $(n=7)$;
2, 3, 4, 5, 6, 7 $(n=8)$.
\section{\'Ecritures minimales}
Dixmier \cite{Di0} defines an {\em \'ecriture minimale}
of the Poincar\'e series as an expression $P(t) = a(t) / \prod (t^{d_i}-1)$
with minimal $a(1)$ (or, equivalently, with minimal $\prod d_i$; indeed,
$\underset{t\rightarrow 1}{\lim} (t-1)^{n-2} P(t)= a(1) / \prod d_i$).
He gives the example of $V_7$ where
$P(t) = a(t)/\prod (t^{d_i}-1) = b(t)/\prod (t^{e_i}-1)$
with $d_i = 4,8,12,12,20$ and $e_i = 4,8,8,12,30$,
and there exist systems of parameters of degrees
4, 8, 12, 12, 20 and of degrees 4, 8, 8, 12, 30.
In our case $n = 9$, in view of the restrictions given in the
previous section, the Poincar\'e series can be written
in precisely five minimal ways:
\medskip
\begin{tabular}{cl}
degree $a(t)$ & degrees of factors in denominator \\
\hline
66 & 4, 8, 10, 12, 12, 14, 16 \\
74 & 4, 4, 10, 12, 14, 16, 24 \\
78 & 4, 4, 8, 12, 14, 16, 30 \\
86 & 4, 4, 8, 10, 12, 16, 42 \\
90 & 4, 4, 8, 10, 12, 14, 48 \\
\end{tabular}
\medskip\noindent
and we saw that the first corresponds to a system of parameters.
In fact all five do, as one can show by following the approach
of Dixmier \cite{Di1}.
\begin{proposition} {\rm (Dixmier \cite{Di1})}
Let $G$ be a reductive group over $\mathbb{C}$, with a rational representation
in a vector space $R$ of finite dimension over $\mathbb{C}$. Let $\mathbb{C}[R]$ be the
algebra of complex polynomials on $R$, $\mathbb{C}[R]^G$ the subalgebra of
$G$-invariants, and $\mathbb{C}[R]^G_d$ the subset of homogeneous polynomials
of degree $d$ in $\mathbb{C}[R]^G$. Let $V$ be the affine variety such that
$\mathbb{C}[V] = \mathbb{C}[R]^G$. Let $\delta = \dim V$. Let $(q_1,\ldots,q_\delta)$
be a sequence of positive integers. Assume that for each subsequence
$(j_1,\ldots,j_p)$ of $(q_1,\ldots,q_\delta)$ the subset
of points of $V$ where all elements of all $\mathbb{C}[R]^G_j$ with
$j \in \{j_1,\ldots,j_p\}$ vanish has codimension not less than $p$
in $V$. Then $\mathbb{C}[R]^G$ has a system of parameters of degrees
$q_1,\ldots,q_\delta$.
\end{proposition}
Dixmier gives the covariant $l := (f,f)_8$ and invariants $q_j$ of degree $j$
($j=4,8,10,12,14,16$) such that if $l=0$ and all $q_j$ vanish then
$f$ belongs to the nullcone. It follows that the set of elements in $V$
where $l=0$ and $p$ of the invariants $q_j$ vanish has codimension
not less than $p+1$.
Note that when all invariants of degree $3j$ vanish then also all
invariants of degree $j$ vanish. Therefore, each of the above
five sequences has the property that a subsequence $\sigma$ of length $p+1$
contains at least $p$ distinct elements, and the set of elements in $V$
where $l=0$ and all invariants of the degrees in $\sigma$ vanish has
codimension not less than $p+1$.
Let $[j_1,\ldots,j_p]'$ be the codimension in $V$ of the set of elements
where $l \ne 0$ and all invariants of degrees in $\{j_1,\ldots,j_p\}$
vanish. In order to show that each of the five sequences above is
the sequence of degrees of a system of parameters it suffices to show
that $[4,14]' \ge 3$, $[4,10,14]' \ge 4$, $[4,8,10,14]' \ge 5$,
$[4,8,14,16,30]' \ge 6$, $[4,8,10,16,42]' \ge 6$, given that Dixmier
already proved the requirements of the proposition for the first sequence.
We did this, using instead of `all invariants of degree $j$'
the invariants $p_4,q_4,p_8,p_{10},p_{12},p_{14},p_{16}$ defined
by Dixmier, and moreover $p_{30}$ and $p_{42}$ found by putting
$\tau_1 := (\psi_8,\psi_{10})_0 \in V_{6,10}$,
$\tau_2 := (\psi_8,\psi_{10})_1 \in V_{4,10}$,
$\tau_3 := (\psi_9,\psi_{10})_0 \in V_{6,14}$,
$\tau_4 := (\psi_9,\psi_{10})_1 \in V_{4,14}$,
$p_{30} := ((\tau_1,\tau_1)_4,\tau_2)_4$,
$p_{42} := ((\tau_3,\tau_3)_4,\tau_4)_4$.
The details are very similar to the computation made by Dixmier.
The only less trivial part was to show that $[4,10,14]' \ge 4$,
which was done using the computer algebra system Singular. Thus:
\begin{proposition}\label{fivehsops}
The ring of invariants of $V_9$ has systems of parameters with
each of the five sequences of degrees
$4$, $8$, $10$, $12$, $12$, $14$, $16$ and
$4$, $4$, $10$, $12$, $14$, $16$, $24$ and
$4$, $4$, $8$, $12$, $14$, $16$, $30$ and
$4$, $4$, $8$, $10$, $12$, $16$, $42$ and
$4$, $4$, $8$, $10$, $12$, $14$, $48$. \qed
\end{proposition}
|
\section{Cosmic Ray Observations using Atmospheric Calorimetry}\label{sec:atmocal}
\subsection{The Air Fluorescence Technique}
The charged secondary particles in extensive air showers produce copious
amounts of ultraviolet light -- of order $10^{10}$ photons per meter near the
peak of a $10^{19}$~eV shower. Some of this light is due to nitrogen
fluorescence, in which molecular nitrogen excited by a passing shower emits
photons isotropically into several dozen spectral bands between $300$ and
$420$~nm. A much larger fraction of the shower light is emitted as Cherenkov
photons, which are strongly beamed along the shower axis. With square-meter
scale telescopes and sensitive photodetectors, the UV emission from the
highest energy air showers can be observed at distances in excess of $30$~km
from the shower axis.
The flux of fluorescence photons from a given point on an air shower track is
proportional to $dE/dX$, the energy loss of the shower per unit slant depth
$X$ of traversed atmosphere~\cite{Bunner:1967,Arqueros:2008cx}. The emitted
light can be used to make a calorimetric estimate of the energy of the
primary cosmic ray~\cite{Baltrusaitis:1985mx,Unger:2008uq}, after a small
correction for the ``missing energy'' not contained in the electromagnetic
component of the shower. Note that a large fraction of the light received
from a shower may be contaminated by Cherenkov photons. However, if the
Cherenkov fraction is carefully estimated, it can also be used to measure the
longitudinal development of a shower~\cite{Unger:2008uq}.
The fluorescence technique can also be used to determine cosmic ray
composition. The slant depth at which the energy deposition rate, $dE/dX$,
reaches its maximum value, denoted $X_\text{max}$\xspace, is correlated with the mass of the
primary particle~\cite{Linsley:1977,McComb:1982}. Showers generated by light
nuclei will, on average, penetrate more deeply into the atmosphere than
showers initiated by heavy particles of the same energy, although the exact
behavior is dependent on details of hadronic interactions and must be
inferred from Monte Carlo simulations. By observing the UV light from air
showers, it is possible to estimate the energies of individual cosmic rays,
as well as the average mass of a cosmic ray data set.
\subsection{Challenges of Atmospheric Calorimetry}
The atmosphere is responsible for producing light from air showers. Its
properties are also important for the transmission efficiency of light from
the shower to the air fluorescence detector. The atmosphere is variable, and
so measurements performed with the air fluorescence technique must be
corrected for changing conditions, which affect both light production and
transmission.
For example, extensive balloon measurements conducted at the
Pierre Auger Observatory\xspace~\cite{Keilhauer:2005ja} and a study using radiosonde data from various
geographic locations~\cite{BADC:website} have shown that the altitude profile
of the atmospheric depth, $X(h)$, typically varies by $\sim5$~$\text{g~cm}^{-2}$\xspace from one
night to the next. In extreme cases, the depth can change by $20$~$\text{g~cm}^{-2}$\xspace on
successive nights, which is similar to the differences in depth between the
seasons~\cite{Wilczynska:2006wt}. The largest variations are comparable to
the $X_\text{max}$\xspace resolution of the Auger air fluorescence detector, and could
introduce significant biases into the determination of $X_\text{max}$\xspace if not properly
measured. Moreover, changes in the bulk properties of the atmosphere such as
air pressure $p$, temperature $T$, and humidity $u$ can have a significant
effect on the rate of nitrogen fluorescence emission~\cite{Keilhauer:2005nk},
as well as light transmission.
In the lowest $15$ km of the atmosphere where air shower measurements occur,
sub-$\mu$m to mm-sized aerosols also play an important role in modifying the
light transmission. Most aerosols are concentrated in a boundary layer that
extends about $1$~km above the ground, and throughout most of the
troposphere, the ultraviolet extinction due to aerosols is typically several
times smaller than the extinction due to
molecules~\cite{Tegen:1996,Baltensperger:2003,Kinne:2006}. However, the
variations in aerosol conditions have a greater effect on air shower
measurements than variations in $p$, $T$, and $u$, and during nights with
significant haze, the light flux from distant showers can be reduced by
factors of $3$ or more due to aerosol attenuation. The vertical density
profile of aerosols, as well as their size, shape, and composition, vary
quite strongly with location and in time, and depending on local particle
sources (dust, smoke, etc.) and sinks (wind and rain), the density of
aerosols can change substantially from hour to hour. If not properly
measured, such dynamic conditions can bias shower reconstructions.
\subsection{The Pierre Auger Observatory}
The Pierre Auger Observatory contains two cosmic ray detectors. The first is
a Surface Detector (SD) comprising 1600 water Cherenkov stations to observe
air shower particles that reach the ground~\cite{Allekotte:2007sf}. The
stations are arranged on a triangular grid of $1.5$~km spacing, and the full
SD covers an area of 3,000~km$^2$. The SD has a duty cycle of nearly 100\%,
allowing it to accumulate high-energy statistics at a much higher rate than
was possible at previous observatories.
Operating in concert with the SD is a Fluorescence Detector (FD) of 24 UV
telescopes~\cite{fdpaper}. The telescopes are arranged to overlook the SD
from four buildings around the edge of the ground array. Each of the four FD
buildings contains six telescopes, and the total field of view at each site
is $180^\circ$ in azimuth and $1.8^\circ-29.4^\circ$ in elevation. The main
component of a telescope is a spherical mirror of area $11$~m$^2$ that
directs collected light onto a camera of 440 hexagonal photomultipliers
(PMTs). One photomultiplier ``pixel'' views approximately
$1.5^\circ\times1.5^\circ$ of the sky, and its output is digitized at
$10$~MHz. Hence, every PMT camera can record the development of air showers
with $100$~ns time resolution.
The FD is only operated during dark and clear conditions, when the shower UV
signal is not overwhelmed by moonlight or blocked by low clouds or rain.
These limitations restrict the FD duty cycle to $\sim10\%-15$\%, but unlike
the SD, the FD data provide calorimetric estimates of shower energies.
Simultaneous SD and FD measurements of air showers, known as hybrid
observations, are used to calibrate the absolute energy scale of the SD,
reducing the need to calibrate the SD with shower simulations. The hybrid
operation also dramatically improves the geometrical and longitudinal profile
reconstruction of showers measured by the FD, compared to showers observed by
the FD
alone~\cite{Sommers:1995dm,AbuZayyad:2000ay,Mostafa:2006id,Dawson:2007di}.
This high-quality hybrid data set is used for all physics analyses based on
the FD.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=0.6\textwidth,clip]{auger-detector}
\caption{\label{fig:auger_detector} \slshape The Surface Detector
stations and Fluorescence Detector sites of the Pierre Auger Observatory\xspace. Also
shown are the locations of Malarg\"{u}e\xspace and the atmospheric monitoring
instruments operating at the Observatory (see text for details).}
\end{center}
\end{figure}
To remove the effect of atmospheric fluctuations that would otherwise impact
FD measurements, an extensive atmospheric monitoring program is carried out
at the Pierre Auger Observatory\xspace. A list of monitors and their locations relative to the FD
buildings and SD array are shown in fig.~\ref{fig:auger_detector}.
Atmospheric conditions at ground level are measured by a network of weather
stations at each FD site and in the center of the SD; these provide updates
on ground-level conditions every five minutes. In addition, regular
meteorological radiosonde flights (one or two per week) are used to measure
the altitude profiles of atmospheric pressure, temperature, and other bulk
properties of the air. The weather station monitoring and radiosonde flights
are performed day or night, independent of the FD data acquisition.
During the dark periods suitable for FD data-taking, hourly measurements
of aerosols are made using the FD telescopes, which record vertical UV laser
tracks produced by a Central Laser Facility (CLF) deployed on site since
2003~\cite{Fick:2006yn}. These measurements are augmented by data from
lidar stations located near each FD building~\cite{BenZvi:2006xb}, a
Raman lidar at one FD site, and the eXtreme Laser Facility (or XLF, named for
its remote location) deployed in November 2008. Two Aerosol Phase Function
Monitors (APFs) are used to determine the aerosol scattering properties of
the atmosphere using collimated horizontal light beams produced by Xenon
flashers~\cite{BenZvi:2007px}. Two optical telescopes --- the Horizontal
Attenuation Monitor (HAM) and the (F/ph)otometric Robotic Telescope for
Atmospheric Monitoring (FRAM) --- record data used to determine the
wavelength dependence of the aerosol
attenuation~\cite{BenZvi:2007it,BenZvi:2007uj}. Finally, clouds are measured
hourly by the lidar stations, and infrared cameras on the roof of
each FD building are used to record the cloud coverage in the FD field of
view every five minutes~\cite{Valore:2009}.
\section{Introduction}\label{sec:introduction}
The Pierre Auger Observatory\xspace in Malarg\"{u}e\xspace, Argentina (69$^{\circ}$ W, 35$^{\circ}$ S, 1400~m a.s.l.) is
a facility for the study of ultra-high energy cosmic rays. These are primarily
protons and nuclei with energies above $10^{18}$~eV. Due to the extremely low
flux of high-energy cosmic rays at Earth, the direct detection of such
particles is impractical; but when cosmic rays enter the atmosphere, they
produce extensive air showers of secondary particles. Using the atmosphere as
the detector volume, the air showers can be recorded and used to reconstruct
the energies, arrival directions, and nuclear mass composition of primary
cosmic ray particles. However, the constantly changing properties of the
atmosphere pose unique challenges for cosmic ray measurements.
In this paper, we describe the atmospheric monitoring data recorded at the Pierre Auger Observatory\xspace
and their effect on the reconstruction of air showers. The paper is organized
as follows: Section~\ref{sec:atmocal} contains a review of the observation of
air showers by their ultraviolet light emission, and includes a description of
the Pierre Auger Observatory\xspace and the issues of light production and transmission that arise when
using the atmosphere to make cosmic ray measurements. The specifics of light
attenuation by aerosols and molecules are described in
Section~\ref{sec:prod_and_trans}. An overview of local molecular measurements
is given in Section~\ref{sec:molecular_effects}, and in
Section~\ref{sec:measurements} we discuss cloud-free aerosol measurements
performed at the Observatory. The impact of these atmospheric measurements on
the reconstruction of air showers is explored in
Section~\ref{sec:aerosol_effects}. Cloud measurements with infrared cameras
and backscatter lidars are briefly described in Section~\ref{sec:future}.
Conclusions are given in Section~\ref{sec:conclusion}.
\input atmocal
\input atmotrans
\input measmol
\input measaer
\input hybridres
\input futuredevel
\section{Conclusions}\label{sec:conclusion}
A large collection of atmospheric monitors is operated at the Pierre Auger Observatory\xspace to provide
frequent observations of molecular and aerosol conditions across the
detector. These data are used to estimate light scattering losses between
air showers and the FD telescopes, to correct air shower light production for
various weather effects, and to prevent cloud-obscured data from distorting
estimates of the shower energies, shower maxima, and the detector aperture.
In this paper, we have described the various light production and
transmission effects due to molecules and aerosols. These effects have been
converted into uncertainties in the hybrid reconstruction. Most of the
reported uncertainties are systematic, not only due to the use of local
empirical models to describe the atmosphere --- such as the monthly molecular
profiles --- but also because of the nature of the atmospheric uncertainties
--- such as the systematics-dominated and highly correlated aerosol optical
depth profiles.
Molecular measurements are vital for the proper determination of light
production in air showers, and molecular scattering is the dominant term in
the description of atmospheric light propagation. However, the time
variations in molecular scattering conditions are small relative to
variations in the aerosol component. The inherent variability in aerosol
conditions can have a significant impact on the data if aerosol measurements
are not incorporated into the reconstruction. Because the highest energy air
showers are viewed at low elevation angles and through long distances in the
aerosol boundary layer, aerosol effects become increasingly important at high
energies.
Efforts are currently underway to reduce the systematic uncertainties due to
the atmosphere, with particularly close attention paid to the uncertainties
in energy and $X_\text{max}$\xspace. The shoot-the-shower program will improve the time
resolution of atmospheric measurements, and increase the identification of
atmospheric inhomogeneities that can affect observations of showers with the
FD telescopes.
\section{Acknowledgments}\label{sec:acknowledgments}
\input acknowledgments
\section{The Production of Light by the Shower and its Transmission through the
Atmosphere}\label{sec:prod_and_trans}
Atmospheric conditions impact on both the production and transmission of UV
shower light recorded by the FD. The physical conditions of the molecular
atmosphere have several effects on fluorescence light production, which we
summarize in Section~\ref{subsec:weath_light_prod}. We treat light
transmission, outlined in Section~\ref{subsec:weath_trans}, primarily as a
single-scattering process characterized by the atmospheric optical depth
(Sections~\ref{subsubsec:mol_optical_depth} and
\ref{subsubsec:aero_optical_depth}) and scattering angular dependence
(Section~\ref{subsubsec:mol_aero_scattering}). Multiple scattering
corrections to atmospheric transmission are discussed in
Section~\ref{subsubsec:mult_scat}.
\subsection{The Effect of Weather on Light Production}\label{subsec:weath_light_prod}
The yields of light from the Cherenkov and fluorescence emission processes
depend on the physical conditions of the gaseous mixture of molecules in the
atmosphere. The production of Cherenkov light is the simpler of the two
cases, since the number of photons emitted per charged particle per meter per
wavelength interval depends only on the refractive index of the atmosphere
$n(\lambda,p,T)$. The dependence of this quantity on pressure, temperature,
and wavelength $\lambda$ can be estimated analytically, and so the effect of
weather on the light yield from the Cherenkov process are relatively simple
to incorporate into air shower reconstructions.
The case of fluorescence light is more complex, not only because it is
necessary to consider additional weather effects on the light yield, but also
due to the fact that several of these effects can be determined only by
difficult experimental measurements (see
\cite{Kakimoto:1995pr,Nagano:2004am,Abbasi:2008zz,Bohacova:2008vg} and
references in \cite{Arqueros:2008en}).
One well-known effect of the weather on light production is the collisional
quenching of fluorescence emission, in which the radiative transitions of
excited nitrogen molecules are suppressed by molecular collisions. The rate
of collisions depends on pressure and temperature, and the form of this
dependence can be predicted by kinetic gas
theory~\cite{Bunner:1967,Nagano:2004am}. However, the cross section for
collisions is itself a function of temperature, which introduces an
additional term into the $p$ and $T$ dependence of the yield. The
temperature dependence of the cross section cannot be predicted \textsl{a
priori}, and must be determined with laboratory
measurements~\cite{Ave:2007xh}.
Water vapor in the atmosphere also contributes to collisional quenching, and
so the fluorescence yield has an additional dependence on the absolute
humidity of the atmosphere. This dependence must also be determined
experimentally, and its use as a correction in shower reconstructions using
the fluorescence technique requires regular measurements of the altitude
profile of humidity. A full discussion of these effects is beyond the scope
of this paper, but detailed descriptions are available
in~\cite{Arqueros:2008cx,Keilhauer:2005nk,Keilhauer:2008sy}. We will
summarize the estimates of their effect on shower energy and $X_\text{max}$\xspace in
Section~\ref{subsec:mol_uncertainties}.
\subsection{The Effect of Weather on Light Transmission}\label{subsec:weath_trans}
The attenuation of light along a path through the atmosphere between a light
source and an observer can be expressed as a transmission coefficient $\mathcal{T}$\xspace,
which gives the fraction of light not absorbed or scattered along the path.
If the optical thickness (or optical depth) of the path is $\tau$, then
$\mathcal{T}$\xspace is estimated using the Beer-Lambert-Bouguer law:
\begin{linenomath}
\begin{equation}\label{eq:trans_od}
\mathcal{T}=e^{-\tau}.
\end{equation}
\end{linenomath}
The optical depth of the air is affected by the density and composition of
molecules and aerosols, and can be treated as the sum of molecular and
aerosol components: $\tau=\tau_m+\tau_a$. The optical depth is a function of
wavelength and the orientation of a path within the atmosphere. However, if
the atmospheric region of interest is composed of horizontally uniform
layers, then the full spatial dependence of $\tau$ reduces to an altitude
dependence, such that $\tau\equiv\tau(h,\lambda)$. For a slant path elevated
at an angle $\varphi$ above the horizon, the light transmission along the
path between the ground and height $h$ is
\begin{linenomath}
\begin{equation}\label{eq:aerotrans}
\mathcal{T}(h,\lambda,\varphi) =
e^{-\tau(h,\lambda)/\sin{\varphi}}.
\end{equation}
\end{linenomath}
In an air fluorescence detector, a telescope recording isotropic
fluorescence emission of intensity $I_0$ from a source of light along a
shower track will observe an intensity
\begin{linenomath}
\begin{equation}\label{eq:intensity}
I=I_0\cdot
\mathcal{T}_m\cdot\mathcal{T}_a\cdot(1+H.O.)\cdot\frac{\Delta\Omega}{4\pi}\text{,}
\end{equation}
\end{linenomath}
where $\Delta\Omega$ is the solid angle subtended by the telescope diaphragm
as seen from the light source. The molecular and aerosol transmission factor
$\mathcal{T}_m$\xspace$\cdot$$\mathcal{T}_a$\xspace primarily represents single-scattering of photons out
of the field of view of the telescope. In the ultraviolet range used for air
fluorescence measurements, the absorption of light is much less important
than scattering~\cite{Tegen:1996,Seinfeld:2006}, although there are some
exceptions discussed in Section~\ref{subsubsec:mol_optical_depth}. The term
$H.O.$ is a higher-order correction to the Beer-Lambert-Bouguer law that
accounts for the single and multiple scattering of Cherenkov and fluorescence
photons into the field of view.
To estimate the transmission factors and scattering corrections needed in
eq.~\eqref{eq:intensity}, it is necessary to measure the vertical height
profile and wavelength dependence of the optical depth $\tau(h,\lambda)$, as
well as the angular distribution of light scattered from atmospheric
particles, also known as the phase function $P(\theta)$. For these
quantities, the contributions due to molecules and aerosols are considered
separately.
\subsubsection{The Optical Depth of Molecules}\label{subsubsec:mol_optical_depth}
The probability per unit length that a photon will be scattered or absorbed
as it moves through the atmosphere is given by the total volume extinction
coefficient
\begin{linenomath}
\begin{equation}\label{eq:ext_coeff}
\alpha_\text{ext}(h,\lambda) = \alpha_\text{abs}(h,\lambda)
+ \beta(h,\lambda)\text{,}
\end{equation}
\end{linenomath}
where $\alpha_\text{abs}$\xspace and $\beta$ are the coefficients of absorption and
scattering, respectively. The vertical optical depth between a telescope
at ground level and altitude $h$ is the integral of the atmospheric
extinction along the path:
\begin{linenomath}
\begin{equation}\label{eq:optical_depth_definition}
\tau(h,\lambda)=\int_{h_\text{gnd}}^h\alpha_\text{ext}(h',\lambda)dh'.
\end{equation}
\end{linenomath}
Molecular extinction in the near UV is primarily an elastic scattering
process, since the Rayleigh scattering of light by molecular nitrogen
(N$_2$\xspace) and oxygen (O$_2$\xspace) dominates inelastic scattering and
absorption~\cite{Killinger:1987}. For example, the Raman scattering cross
sections of N$_2$\xspace and O$_2$\xspace are approximately $10^{-30}$~cm$^{-2}$
between $300-420$~nm \cite{Burris:1992}, much smaller than the Rayleigh
scattering cross section of air ($\sim10^{-27}$~cm$^{-2}$) at these
wavelengths~\cite{Bucholtz:1995}. Moreover, while O$_2$\xspace is an important
absorber in the deep UV, its absorption cross section is effectively zero
for wavelengths above $240$~nm~\cite{Seinfeld:2006}. Ozone (O$_3$\xspace)
molecules absorb light in the UV and visible bands, but O$_3$\xspace is mainly
concentrated in a high-altitude layer above the atmospheric volume used for
air fluorescence measurements~\cite{Seinfeld:2006}.
Therefore, for the purpose of air fluorescence detection, the total
molecular extinction $\alpha_\text{ext}^m(h,\lambda)$ simply reduces to the
scattering coefficient $\beta_m(h,\lambda)$. At standard temperature and
pressure, molecular scattering can be defined analytically in terms of the
Rayleigh scattering cross section \cite{Bucholtz:1995,Naus:2000}:
\begin{linenomath}
\begin{equation}\label{eq:molecular_scattering_coeff}
\beta^\text{STP}_m(h,\lambda)\equiv\beta_s(\lambda) =N_s\sigma_\text{R}(\lambda)
= \frac{24\pi^3}{N_s\lambda^4}
\left(\frac{n_s^2(\lambda)-1}{n_s^2(\lambda)+2}\right)^2
\frac{6+3\rho(\lambda)}{6-7\rho(\lambda)}.
\end{equation}
\end{linenomath}
In this expression, $N_s$ is the molecular number density under standard
conditions and $n_s(\lambda)$ is the index of refraction of air. The
depolarization ratio of air, $\rho(\lambda)$, is determined by the
asymmetry of N$_2$\xspace and O$_2$\xspace molecules, and its value is
approximately $0.03$ in the near UV~\cite{Bucholtz:1995}. The wavelength
dependence of these quantities means that between $300$~nm and $420$~nm,
the wavelength dependence of molecular scattering shifts from the classical
$\lambda^{-4}$ behavior to an effective value of $\lambda^{-4.2}$.
Since the atmosphere is an ideal gas, the altitude dependence of the
scattering coefficient can be expressed in terms of the vertical
temperature and pressure profiles $T(h)$ and $p(h)$,
\begin{linenomath}
\begin{equation}\label{eq:scatter_profile}
\alpha_\text{ext}^m(h,\lambda)\equiv\beta_m(h,\lambda) = \beta_s(\lambda)
\frac{p(h)}{p_s} \frac{T_s}{T(h)}\text{,}
\end{equation}
\end{linenomath}
where $T_s$ and $p_s$ are standard temperature and
pressure~\cite{Bucholtz:1995}. Given the profiles $T(h)$ and $p(h)$
obtained from balloon measurements or local climate models, the vertical
molecular optical depth is estimated via numerical integration of
equations~\eqref{eq:optical_depth_definition} and
\eqref{eq:scatter_profile}.
\subsubsection{The Optical Depth of Aerosols}\label{subsubsec:aero_optical_depth}
The picture is more complex for aerosols than for molecules because in
general it is not possible to calculate the total aerosol extinction
coefficient analytically. The particulate scattering theory of Mie, for
example, depends on the simplifying assumption of spherical
scatterers~\cite{Mie:1908}, a condition which often does not hold in the
field\footnote{Note that in spite of this, aerosol scattering is often
referred to as ``Mie scattering.''}. Moreover, aerosol scattering depends
on particle composition, which can change quite rapidly depending on the
wind and weather conditions.
Therefore, knowledge of the aerosol transmission factor $\mathcal{T}_a$\xspace depends on
frequent field measurements of the vertical aerosol optical depth $\tau_{a}(h,\lambda)$\xspace.
Like other aerosol properties, the altitude profile of $\tau_{a}(h,\lambda)$\xspace can change
dramatically during the course of a night. However, in general $\tau_{a}(h,\lambda)$\xspace
increases rapidly with $h$ only in the first few kilometers above ground
level, due to the presence of mixed aerosols in the planetary boundary
layer.
In the lower atmosphere, the majority of aerosols are concentrated in the
mixing layer. The thickness of the mixing layer is measured from the
prevailing ground level in the region, and its height roughly follows the
local terrain (excluding small hills and escarpments). This gives the
altitude profile of $\tau_{a}(h,\lambda)$\xspace a characteristic shape: a nearly linear
increase at the lowest heights, followed by a flattening as the aerosol
density rapidly decreases with altitude.
Figure~\ref{fig:vaod_transmission} depicts an optical depth profile
inferred using vertical laser shots from the CLF at $355$~nm viewed from
the FD site at Los Leones. The profile, corresponding to a moderately
clear atmosphere, can be considered typical of this location. Also shown
is the aerosol transmission coefficient between points along the vertical
laser beam and the viewing FD, corresponding to a ground distance of
$26$~km.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=\textwidth,clip]{transmittance}
\caption{\label{fig:vaod_transmission} \slshape
Left: a vertical aerosol optical depth profile
$\tau_a(h,355~\text{nm})$ measured using the FD at Los Leones
with vertical laser shots from the CLF ($26$~km distance).
The uncertainties are dominated by systematic effects and are
highly correlated. Also shown is the monthly average
molecular optical depth $\tau_m(h,355~\text{nm})$. Right:
molecular and aerosol light transmission factors for the
atmosphere between the vertical CLF laser beam and the Los
Leones FD. The dashed line at $1$~km indicates the lower
edge of the FD field of view at this distance (see
Section~\ref{subsubsec:central_laser_facility} for details).}
\end{center}
\end{figure}
The wavelength dependence of $\tau_{a}(h,\lambda)$\xspace depends on the wavelength of the
incident light and the size of the scattering aerosols. A conventional
parameterization for the dependence is a power law due to
{\AA}ngstr{\o}m\xspace~\cite{Angstrom:1929},
\begin{linenomath}
\begin{equation}\label{eq:wavelength_dependence}
\tau_a(h,\lambda) =
\tau(h,\lambda_0)\cdot\left(\frac{\lambda_0}{\lambda}\right)^\gamma,
\end{equation}
\end{linenomath}
where $\gamma$ is known as the {\AA}ngstr{\o}m\xspace exponent. The exponent is also
measured in the field, and the measurements are normalized to the value of
the optical depth at a reference wavelength $\lambda_0$. The normalization
point used at the Auger Observatory is the wavelength of the Central Laser
Facility, $\lambda_0=355$~nm, approximately in the center of the nitrogen
fluorescence spectrum.
The {\AA}ngstr{\o}m\xspace exponent is determined by the size distribution of scattering
aerosols, such that smaller particles have a larger exponent --- eventually
reaching the molecular limit of $\gamma\approx4$ --- while larger particles
give rise to a smaller $\gamma$ and thus a more ``wavelength-neutral''
attenuation~\cite{McCartney:1976,Schuster:2006}. For example, in a review
of the literature by Eck et al.~\cite{Eck:1999}, aerosols emitted from
burning vegetation and urban and industrial areas are observed to have a
relatively large {\AA}ngstr{\o}m\xspace coefficient ($\gamma=1.41\pm0.35$). These
environments are dominated by fine ($<1~\mu$m) ``accumulation mode''
particles, or aerodynamically stable aerosols that do not coalesce or
settle out of the atmosphere. In desert environments, where coarse
($>1~\mu$m) particles dominate, the wavelength dependence is almost
negligible~\cite{Eck:1999,Whittet:1987}.
\subsubsection{Angular Dependence of Molecular and Aerosol Scattering}\label{subsubsec:mol_aero_scattering}
Only a small fraction of the photons emitted from an air shower arrive at a
fluorescence detector without scattering. The amount of scattering must be
estimated during the reconstruction of the shower, and so the scattering
properties of the atmosphere need to be well understood.
For both molecules and aerosols, the angular dependence of scattering is
described by normalized angular scattering cross sections, which give the
probability per unit solid angle $P(\theta)=\sigma^{-1}d\sigma/d\Omega$
that light will scatter out of the beam path through an angle $\theta$.
Following the convention of the atmospheric literature, this work will
refer to the normalized cross sections as the molecular and aerosol phase
functions.
The molecular phase function $P_m(\theta)$ can be estimated analytically,
with its key feature being the symmetry in the forward and backward
directions. It is proportional to the $(1+\cos^2{\theta})$ factor of the
Rayleigh scattering theory, but in air there is a small correction factor
$\delta\approx1\%$ due to the anisotropy of the N$_2$ and O$_2$ molecules
\cite{Bucholtz:1995}:
\begin{linenomath}
\begin{equation}\label{eq:mol_phase_function}
P_m(\theta) = \frac{3}{16\pi(1+2\delta)}
\left(1+3\delta + (1-\delta)\cos^2{\theta}\right).
\end{equation}
\end{linenomath}
The aerosol phase function $P_a(\theta)$, much like the aerosol optical
depth, does not have a general analytical solution, and in fact its
behavior as a function of $\theta$ is quite complex. Therefore, one is
often limited to characterizing the gross features of the light scattering
probability distribution, which is sufficient for the purposes of air
fluorescence detection. In general, the angular distribution of light
scattered by aerosols is very strongly peaked in the forward direction,
reaches a minimum near $90^{\circ}$, and has a small backscattering
component. It is reasonably approximated by the
parameterization~\cite{BenZvi:2007px,Fishburne:1976,Riewe:1978}
\begin{linenomath}
\begin{equation}\label{eq:aero_phase_function}
P_a(\theta) = \frac{1-g^2}{4\pi}\cdot
\left(\frac{1}{(1+g^2-2g\cos{\theta})^{3/2}}+
f\frac{3\cos^2{\theta}-1}{2(1+g^2)^{3/2}}
\right).
\end{equation}
\end{linenomath}
The first term, a Henyey-Greenstein scattering function~\cite{Henyey:1941},
corresponds to forward scattering; and the second term --- a second-order
Legendre polynomial, chosen so that it does not affect the normalization of
$P_a(\theta)$ --- accounts for the peak at large $\theta$ typically found
in the angular distribution of aerosol-scattered light. The quantity
$g=\langle\cos{\theta}\rangle$ measures the asymmetry of scattering, and
$f$ determines the relative strength of the forward and backward scattering
peaks. The parameters $f$ and $g$ are observable quantities which depend
on local aerosol characteristics.
\subsubsection{Corrections for Multiple Scattering}\label{subsubsec:mult_scat}
As light propagates from a shower to the FD, molecular and aerosol scattering
can remove photons that would otherwise travel along a direct path toward an
FD telescope. Likewise, some photons with initial paths outside the detector
field of view can be scattered back into the telescope, increasing the
apparent intensity and angular width of the shower track.
During the reconstruction of air showers, it is convenient to consider the
addition and subtraction of scattered photons to the total light flux in
separate stages. The subtraction of light is accounted for in the
transmission coefficients $\mathcal{T}_m$\xspace and $\mathcal{T}_a$\xspace of eq.~\eqref{eq:intensity}.
Given the shower geometry and measurements of atmospheric scattering
conditions, the estimation of $\mathcal{T}_m$\xspace and $\mathcal{T}_a$\xspace is relatively
straightforward. However, the addition of light due to atmospheric
scattering is less simple to calculate, due to the contributions of multiple
scattering. Multiple scattering has no universal analytical description, and
those analytical solutions which do exist are only valid under restrictive
assumptions that do not apply to typical FD viewing
conditions~\cite{Roberts:2005xv}.
A large fraction of the flux of photons from air showers recorded by an FD
telescope can come from multiply-scattered light, particularly within the
first few kilometers above ground level, where the density of scatterers is
highest. In poor viewing conditions, $10\%-15\%$ of the photons arriving
from the lower portion of a shower track may be due to multiple scattering.
Since these contributions cannot be neglected, a number of Monte Carlo
studies have been carried out to quantify the multiply-scattered component of
recorded shower signals under realistic atmospheric
conditions~\cite{Roberts:2005xv,Pekala:2003wu,Giller:2005,Pekala:2009fe}.
The various simulations indicate that multiple scattering grows with optical
depth and distance from the shower. Based on these results,
Roberts~\cite{Roberts:2005xv} and Pekala et al.~\cite{Pekala:2009fe} have
developed parameterizations of the fraction of multiply-scattered photons in
the shower image. Both parameterizations are implemented in the FD event
reconstruction, and their effect on estimates of the shower energy and shower
maximum are described in section~\ref{subsec:ms_correction}.
\section{Additional Developments}\label{sec:future}
We have estimated the uncertainties in shower energy and $X_\text{max}$\xspace due to
atmospheric transmission, but we have not discussed the impact of clouds on
the hybrid reconstruction, which violate the horizontal uniformity assumption
described in section~\ref{subsec:weath_trans}. A full treatment of this
issue will be the subject of future technical publications, but here we
summarize current efforts to understand their effect on the hybrid data.
\subsection{Cloud Measurements}\label{subsec:clouds}
Cloud coverage has a major influence on the reconstruction of air showers,
but this influence can be difficult to quantify. Clouds can block the
transmission of light from air showers, as shown in
Figure~\ref{fig:profile_with_cloud}, or enhance the observed light flux due
to multiple scattering of the intense Cherenkov light beam. They may occur
in optically thin layers near the top of the troposphere, or in thick banks
which block light from large parts of the FD fiducial volume. The
determination of the composition of clouds is nontrivial, making \textsl{a
priori} estimates of their scattering properties unreliable.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=0.55\textwidth,clip]{profile_with_cloud}
\caption{\label{fig:profile_with_cloud} \slshape Shower light profile
with a large gap due to the presence of an intervening cloud.}
\end{center}
\end{figure}
Due to the difficulty of correcting for the transmission of light through
clouds, it is prudent to remove cloudy data using hard cuts on the shower
profiles. But because clouds can reduce the event rate from different parts
of a fluorescence detector, they also have an important effect on the
aperture of the detector as used in the determination of the spectrum from
hybrid data~\cite{Schussler:2009}. Therefore, it is necessary to estimate
the cloud coverage at each FD site as accurately as possible.
Cloud coverage at the Pierre Auger Observatory\xspace is recorded by Raytheon 2000B infrared cloud
cameras located on the roof of each FD building. The cameras have a spectral
range of $7$~$\mu\text{m}$\xspace to $14$~$\mu\text{m}$\xspace, and photograph the field of view of the
six FD telescopes every 5 minutes during normal data acquisition. After the
image data are processed, a coverage ``mask'' is created for each FD pixel,
which can be used to remove covered pixels from the reconstruction. Such a
mask is shown in fig.~\ref{fig:cloud_mask}.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=\textwidth,clip]{cloud_mask}
\caption{\label{fig:cloud_mask} \slshape A mask of grayscale values used
in the cloud database to indicate the cloud coverage of each
pixel in an FD building. Lighter values indicate greater cloud
coverage.}
\end{center}
\end{figure}
While the IR cloud cameras record the coverage in the FD field of view, they
cannot determine cloud heights. The heights must be measured using the lidar
stations, which observe clouds over each FD site during hourly
two-dimensional scans of the atmosphere~\cite{BenZvi:2006xb}. The Central
Laser Facility can also observe laser echoes from clouds, though the
measurements are more limited than the lidar observations. Cloud height
data from the lidar stations are combined with pixel coverage measurements to
improve the accuracy of cloud studies.
\subsection{Shoot-the-Shower}\label{subsec:sts}
When a distant, high-energy air shower is detected by an FD telescope, the
lidars interrupt their hourly sweeps and scan the plane formed by the image
of the shower on the FD camera. This is known as the ``shoot-the-shower''
mode. The shoot-the-shower mode allows the lidar station to probe for local
atmospheric non-uniformities, such as clouds, which may affect light
transmission between the shower and detector. Figure~\ref{fig:fd_sts}
depicts one of the four shoot-the-shower scans for the cloud-obscured event
shown in fig.~\ref{fig:profile_with_cloud}.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=0.7\textwidth,clip]{fd_sts}
\caption{\label{fig:fd_sts} \slshape Lidar sweep of the shower-detector
plane for the cloud-obscured event shown in
fig.~\ref{fig:profile_with_cloud}. The regions of high
backscatter are laser echoes due to optically thick clouds.}
\end{center}
\end{figure}
A preliminary implementation of shoot-the-shower was described
in~\cite{BenZvi:2006xb}. This scheme has been altered recently to use a fast
on-line hybrid reconstruction now operating at the Observatory. The new
scheme allows for more accurate selection of showers of interest. In
addition, the reconstruction output can be used to trigger other atmospheric
monitors and services, such as radiosonde balloon launches, to provide
measurements of molecular conditions shortly after very high energy air
showers are recorded. ``Balloon-the-shower'' radiosonde measurements began
at the Observatory in early 2009~\cite{Keilhauer:2009}.
\section{Impact of the Atmosphere on Accuracy of Reconstruction of Air Shower Parameters}\label{sec:aerosol_effects}
The atmospheric measurements described in
Sections~\ref{sec:molecular_effects} and \ref{sec:measurements} are fully
integrated into the software used to reconstruct hybrid
events~\cite{Allen:2008zz}. The data are stored in multi-gigabyte MySQL
databases and indexed by observation time, so that the atmospheric conditions
corresponding to a given event are automatically retrieved during off-line
reconstruction. The software is driven by XML datacards that provide
``switches'' to study different effects on the
reconstruction~\cite{Allen:2008zz}: for example, aerosol attenuation,
multiple scattering, water vapor quenching, and other effects can be switched
on or off while reconstructing shower profiles. Propagation of
atmospheric uncertainties is also available.
In this section, we estimate the influence of atmospheric effects and the
uncertainties in our knowledge of these effects on the reconstruction of
hybrid events recorded between December 2004 and December 2008. The data
have been subjected to strong quality cuts to remove events contaminated with
clouds\footnote{The presence of clouds distorts the observation of the shower
profile as the UV light is strongly attenuated. Clouds are also responsible
for multiple scattering of the light. Strong cuts on the shower profile
shape can remove observations affected by clouds.}, as well as geometry cuts
to eliminate events poorly viewed by the FD telescopes. These cuts include:
\begin{itemize}
\item Gaisser-Hillas fit of the shower profile with $\chi^2/\text{NDF}<2.5$
\item Gaps in the recorded light profile $<20\%$ of the length of the
profile
\item Shower maximum $X_\text{max}$\xspace observed within the field of view of the FD
telescopes
\item Uncertainty in $X_\text{max}$\xspace (before atmospheric corrections) $<40$~$\text{g~cm}^{-2}$\xspace
\item Relative uncertainty in energy (before atmospheric corrections)
$<20\%$
\end{itemize}
The cuts are the same as those used in studies of the energy
spectrum~\cite{Schussler:2009,DiGiulio:2009} and $X_\text{max}$\xspace
distribution~\cite{Bellido:2009}.
We first describe the effects of the molecular information on
the determinations of energy and $X_\text{max}$\xspace. This is followed by a discussion of
the impact of aerosol information on the measurement of these quantities.
\subsection{Systematic Uncertainties due to the Molecular Atmosphere}\label{subsec:mol_uncertainties}
\subsubsection{Monthly Models}\label{subsubsec:mmm_vary}
The molecular transmission is determined largely by atmospheric pressure
and temperature, as described in eq.~\eqref{eq:scatter_profile}. For the
purpose of reconstruction, these quantities are described by monthly
molecular models generated using local radiosonde data. Pressure and
temperature also affect the fluorescence yield via collisional quenching,
and this effect is included in the hybrid reconstruction.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=\textwidth,clip]{mmm_vs_usstd}
\caption{\label{fig:mmmVsUSStd} Comparison of hybrid events
reconstructed using monthly balloon flights vs. the U.S. Standard
Atmosphere (1976) profile model~\cite{USStdAtm:1976}. Uncertainties
denote the RMS spread.}
\end{center}
\end{figure}
The importance of local atmospheric profile measurements is illustrated in
fig.~\ref{fig:mmmVsUSStd}. The hybrid data have been reconstructed using
the monthly profile models described in Section~\ref{subsec:monthly_models}
and compared to events reconstructed with the U.S. Standard
Atmosphere~\cite{USStdAtm:1976}. The values of $X_\text{max}$\xspace determined using the
molecular atmosphere described by the local monthly models are, on average,
$15$~$\text{g~cm}^{-2}$\xspace larger than the values obtained if the U.S. Standard Atmosphere
is used. This shift is energy-dependent because the average distance
between shower tracks and the FD telescopes increases with energy. It is
clear that the U.S. Standard Atmosphere is not an appropriate climate
model for Malarg\"{u}e\xspace; but even a local annual model would introduce seasonal
shifts into the measurement of $X_\text{max}$\xspace given the monthly variations observed
in the local vertical depth profile (fig.~\ref{fig:monthly_profiles}).
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=\textwidth,clip]{a07_mmmVsBalloon}
\caption{\label{fig:a07_mmmVsBalloon} Comparison of simulated events
reconstructed with monthly average atmospheric profiles vs. profile
measurements from 109 cloud-free balloon flights. The dotted lines
indicate the reference for the 109 balloon flights; the uncertainties
indicate the RMS spread.}
\end{center}
\end{figure}
When the monthly models are used, some systematic uncertainties are
introduced into the reconstruction due to atmospheric variations that occur
on timescales shorter than one month. To investigate this effect, we
compare events reconstructed with monthly models vs. local radio soundings.
A set of 109 cloud-free, night-time balloon profiles was identified using
the cloud camera database. The small number of soundings requires the use
of simulated events, so we simulated an equal number of proton and iron
showers between $10^{17.5}$~eV and $10^{20}$~eV, reconstructed them with
monthly and radiosonde profiles, and applied standard cuts to the simulated
dataset. The radiosonde profiles were weighted in the simulation to
account for seasonal biases in the balloon launch rate.
The difference between monthly models and balloon measurements is indicated
in fig.~\ref{fig:a07_mmmVsBalloon}. The use of the models introduces
rather small shifts into the reconstructed energy and $X_\text{max}$\xspace, though there
is an energy-dependent increase in the RMS of the measured energies from
$0.8\%$ to $2.0\%$ over the simulated energy range. The systematic shift
in $X_\text{max}$\xspace is about $2$~$\text{g~cm}^{-2}$\xspace over the full energy scale, with an RMS of
about $8$~$\text{g~cm}^{-2}$\xspace. We interpret the RMS spread as the decrease in
the resolution of the hybrid detector due to variations in the atmospheric
conditions within each month.
\subsubsection{Combined Effects of Quenching and Atmospheric
Variability}\label{subsubsec:quench_vary}
The simulations described in Section~\ref{subsubsec:mmm_vary} used an air
fluorescence model that does not correct the fluorescence emission for
weather-dependent quenching. Recent estimates of quenching due to
water vapor~\cite{Morozov:2005,Waldenmaier:2007am} and the $T$-dependence
of the N$_2$\xspace-N$_2$\xspace and N$_2$\xspace-O$_2$\xspace collisional cross
sections~\cite{Keilhauer:2008sy} allow for detailed studies of their effect
on the production of fluorescence light.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=\textwidth,clip]{k08Comparison}
\caption{\label{fig:k08Comparison} The effect of water vapor quenching
and the $T$-dependent N$_2$\xspace-N$_2$\xspace and N$_2$\xspace-O$_2$\xspace
collisional cross sections $\sigma_\text{coll}(T)$ on the reconstructed
energy and $X_\text{max}$\xspace of simulated showers. The reference (dotted line)
corresponds to showers reconstructed with the fluorescence model of
Keilhauer et al.~\cite{Keilhauer:2008sy}, which includes $T$-dependent
cross sections, and the vapor quenching model of Morozov et al.
$(\sigma^\text{Moro.}_\text{vapor})$~\cite{Morozov:2005}. The markers
correspond to different combinations of quenching effects and vapor
quenching models. See the text for a detailed explanation.}
\end{center}
\end{figure}
We have applied the two quenching effects to simulated showers in
various combinations using $p$, $T$, and $u$ from the monthly model
profiles (see fig.~\ref{fig:k08Comparison}). As different quenching
effects and models were ``switched on'' in the reconstruction, the showers
were compared to a reference reconstruction that used $T$-dependent
collisional cross sections and the vapor quenching model of Morozov et
al.~\cite{Morozov:2005}. We have considered the following three cases:
\begin{enumerate}
\item In the first case, all quenching corrections were omitted (open blue
squares in fig.~\ref{fig:k08Comparison}). The result is a
$5.5\%$ underestimate in shower energy and a $2$~$\text{g~cm}^{-2}$\xspace overestimate
in $X_\text{max}$\xspace with respect to the reference model.
\item In the second case, temperature corrections to the collisional
cross section were included, but water vapor quenching was not (open
red circles in the figure). Without vapor quenching, the energy
is systematically underestimated by $3\%$, and $X_\text{max}$\xspace is
underestimated by $6-7$~$\text{g~cm}^{-2}$\xspace with respect to the reference model.
\item In the third case, all corrections were included, but the vapor
quenching model of Waldenmaier et al.~\cite{Waldenmaier:2007am} was
used (closed black circles). The resulting systematic differences
are $\Delta E/E$\xspace$=0.5\%$ and $\Delta X_\text{max}$\xspace=$2$~$\text{g~cm}^{-2}$\xspace.
\end{enumerate}
We observe that once water vapor quenching is applied, the particular
choice of quenching model has a minor influence. In addition, there is a
small total shift in $X_\text{max}$\xspace due to the offsetting effects of the
$T$-dependent cross sections, which are important at high altitudes, and
the effect of water vapor, which is important at low altitudes. The
compensation of these two effects leaves the longitudinal profiles of
showers relatively undistorted.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=\textwidth,clip]{k08a07_mmmVsBalloon}
\caption{\label{fig:k08a07_mmmVsBalloon}
Comparison of simulated events to determine the combined effects of
atmospheric variability and quenching corrections. The data were
reconstructed in two sets: using monthly profiles plus a fluorescence
model without quenching corrections~\cite{Ave:2007xh}; and using
local radiosonde profiles plus a fluorescence model with water vapor
quenching and $T$-dependent collisional cross
sections~\cite{Keilhauer:2008sy}.}
\end{center}
\end{figure}
In fig.~\ref{fig:k08a07_mmmVsBalloon}, we plot the combined effects of
atmospheric variability around the monthly averages and the quenching
corrections. Simulated showers were reconstructed with two settings:
monthly average profile models and no quenching corrections; and cloud-free
radiosonde profiles with water vapor quenching and $T$-dependent
collisional cross sections. The reconstructed energy is increased by
$5\%$, on average, and, comparing figs.~\ref{fig:a07_mmmVsBalloon} and
\ref{fig:k08a07_mmmVsBalloon}, we see that the quenching corrections are
dominating systematic uncertainties due to the use of monthly models. For
$X_\text{max}$\xspace, the systematic effects of the monthly models offset the quenching
corrections. The spread of the combined measurements increases with
energy, such that the RMS in energy increases from $1.5\%-3.0\%$, and the
RMS in $X_\text{max}$\xspace increases from $7.2-8.4$~$\text{g~cm}^{-2}$\xspace.
\subsection{Uncertainties due to Aerosols}
For a complete understanding of the effects of aerosols on the
reconstruction, several investigations are of interest:
\begin{enumerate}
\item A test of the effect of aerosols on the reconstruction, compared to
the use of a pure molecular atmosphere.
\item A test of the use of aerosol measurements, compared to a simple
parameterization of average aerosol conditions.
\item The propagation of measurement uncertainties in $\tau_{a}(h)$\xspace, $\gamma$, $f$,
and $g$ in the FD reconstruction, and in particular their effect on
uncertainties in energy and $X_\text{max}$\xspace.
\item A test of the horizontal uniformity of aerosol layers within a
zone.
\end{enumerate}
\subsubsection{The Comparison of Aerosol Measurements with a Pure Molecular Atmosphere}
We have compared the reconstruction of hybrid showers using hourly on-site
aerosol measurements with the same showers reconstructed using a purely
molecular atmosphere (fig.~\ref{fig:aero_vs_mol}). Neglecting the presence
of aerosols causes an $8\%$ underestimate in energy at the lower energies.
This underestimate increases to $25\%$ at the higher energies. Moreover, the
distribution of shifted energies contains a long tail: $20\%$ of all showers
have an energy correction $>20\%$; $7\%$ of showers are corrected by $>30\%$;
and $3\%$ of showers are corrected by $>40\%$. The systematic shift in $X_\text{max}$\xspace
ranges from $-1$~$\text{g~cm}^{-2}$\xspace at low energies to almost $10$~$\text{g~cm}^{-2}$\xspace at high
energies, with an RMS of $10-15$~$\text{g~cm}^{-2}$\xspace.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=\textwidth,clip]{dEdX_nomie_2004_2008}
\caption{\label{fig:aero_vs_mol} \slshape
Comparison of hybrid events reconstructed with hourly CLF aerosol
measurements vs. no aerosol correction (i.e., purely molecular
transmission). Uncertainties indicate the RMS spread for each energy.}
\end{center}
\end{figure}
\subsubsection{The Comparison of Aerosol Measurements with an Average Parameterization}
Aerosols clearly play an important role in the transmission and scattering of
fluorescence light, but it is natural to ask if hourly measurements of
aerosol conditions are necessary, or if a fixed average aerosol
parameterization is sufficient for air shower reconstruction.
We can test the sufficiency of average aerosol models by comparing the
reconstruction of hybrid events using hourly weather data against the
reconstruction using an average profile of the aerosol optical depth in Malarg\"{u}e\xspace.
The average profile was constructed using CLF data, and the differences in
the reconstruction between this average model and the hourly data are shown
in fig.~\ref{fig:aerosol_seasonal}, where $\Delta E/E$ and $\Delta
X_\text{max}$ are grouped by season.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=\textwidth]{dEdX_aerosol_seasonal}
\caption{\label{fig:aerosol_seasonal} \slshape
Top: systematic shifts in the hybrid reconstruction of shower energy and
$X_\text{max}$\xspace caused by the use of average aerosol conditions rather than
hourly measurements (indicated by dotted lines). The mean
$\Delta E/E$ and $\Delta X_\text{max}$ per energy bin, plotted with
uncertainties on the means, are arranged by their occurrence in
austral winter, spring, summer, and autumn. Bottom: distributions of the
differences in energy and $X_\text{max}$\xspace, shown with Gaussian fits.}
\end{center}
\end{figure}
Due to the relatively good viewing conditions in Malarg\"{u}e\xspace during austral winter
and fall, and poorer atmospheric clarity during the spring and summer, the
shifts caused by the use of an average aerosol profile exhibit a strong
seasonal dependence. The shifts also exhibit large tails and are
energy-dependent. For example, $\Delta E/E$ nearly doubles during the fall,
winter, and spring, reaching
$-7\%$ (with an RMS of $15\%$) during the winter. The range of seasonal mean
offsets in $X_\text{max}$\xspace is $+2$~$\text{g~cm}^{-2}$\xspace to $-8$~$\text{g~cm}^{-2}$\xspace (with an RMS of $15$~$\text{g~cm}^{-2}$\xspace),
and the offsets depend strongly on the shower energy.
\subsubsection{Propagation of Uncertainties in Aerosol Measurements}\label{subsubsec:aero_error_prop}
Uncertainties in aerosol properties will cause over- or under-corrections of
recorded shower light profiles, particularly at low altitudes and low
elevation angles. On average, systematic overestimates of the aerosol
optical depth will lead to an over-correction of scattering losses and an
overestimate of the shower light flux from low altitudes; this will increase
the shower energy estimate and push the reconstructed $X_\text{max}$\xspace deeper into the
atmosphere. Systematic underestimates of the aerosol optical depth should
have the opposite effect.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=\textwidth,clip]{dEdXmax_uncertainty}
\caption{\label{fig:aero_uncertainty} \slshape Shifts in the
reconstruction of energy and $X_\text{max}$\xspace when the aerosol optical depth is
varied by its $+1\sigma$ systematic uncertainty (red points) and
$-1\sigma$ systematic uncertainty (blue points). The dotted line
corresponds to the central aerosol optical depth measurement. The
uncertainty bars correspond to the sample RMS in each energy bin.}
\end{center}
\end{figure}
The primary source of uncertainty in aerosol transmission comes from the
aerosol optical depth~\cite{Prouza:2007ur} estimated using vertical
CLF laser shots. The uncertainties in the hourly CLF optical depth profiles
are dominated by systematic detector and calibration effects, and smoothing
of the profiles makes the optical depths at different altitudes highly
correlated. Therefore, a reasonable estimate of the systematic uncertainty
in energy and $X_\text{max}$\xspace can be obtained by shifting the full optical depth
profiles by their uncertainties and estimating the mean change in the
reconstructed energy and $X_\text{max}$\xspace.
This procedure was done using hybrid events recorded by telescopes at Los
Leones, Los Morados, and Coihueco, and results are shown in
fig.~\ref{fig:aero_uncertainty}. The energy dependence of the uncertainties
mainly arises from the distribution of showers with distance: low-energy
showers tend to be observed during clear viewing conditions and within
$10$~km of the FD buildings, reducing the effect of the transmission
uncertainties on the reconstruction; and high-energy showers can be observed
in most aerosol conditions (up to a reasonable limit) and are observed at
larger distances from the FD. The slight asymmetry in the shifts is due to
the asymmetric uncertainties of the optical depth profiles.
By contrast to the corrections for the optical depth of the aerosols, the
uncertainties that arise from the wavelength dependence of the aerosol
scattering and of the phase function are relatively unimportant for the
systematic uncertainties in shower energy and $X_\text{max}$\xspace. By reconstructing
showers with average values of the {\AA}ngstr{\o}m\xspace coefficient and the phase function
measured at the Observatory, and comparing the results to showers
reconstructed with the $\pm1~\sigma$ uncertainties in these measurements, we
find that the wavelength dependence and phase function contribute $0.5\%$ and
$1\%$, respectively, to the uncertainty in the energy, and $\sim2$~$\text{g~cm}^{-2}$\xspace to
the systematic uncertainty in $X_\text{max}$\xspace \cite{Prouza:2007ur}. Moreover, the
uncertainties are largely independent of shower energy and distance.
\subsubsection{Evaluation of the Horizontal Uniformity of the Atmosphere}
The non-uniformity of the molecular atmosphere, discussed in
Section~\ref{subsec:horizontal_molecular}, is very minor and introduces
uncertainties $<1\%$ in shower energies and about $1$~$\text{g~cm}^{-2}$\xspace in $X_\text{max}$\xspace.
Non-uniformities in the horizontal distribution of aerosols may also be
present, and we expect these to have an effect on the reconstruction. For
each FD building, the vertical CLF laser tracks only probe the atmosphere
along one light path, but the reconstruction must use this single aerosol
profile across the azimuth range observed at each site. In general, the
assumption of uniformity within an aerosol zone is reasonable, though the
presence of local inhomogeneities such as clouds, fog banks, and sources of
dust and smoke may render it invalid.
The assumption of uniformity can be partially tested by comparing data
reconstructed with different aerosol zones around each eye: for example,
reconstructing showers observed at Los Leones using aerosol data from the Los
Leones and Los Morados zones.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=\textwidth]{dEdXmax_uniformity}
\caption{\label{fig:aero_uniformity} \slshape Shifts in the estimated
shower energy and $X_\text{max}$\xspace when data from the FD buildings at Los
Morados and Los Leones (dotted line) are reconstructed with
swapped aerosol zones. The values give an approximate estimate
of the systematic uncertainty due to aerosol non-uniformities
across the detector. The uncertainties correspond to the
sample RMS in each energy bin.}
\end{center}
\end{figure}
Data from Los Leones and Los Morados were reconstructed using aerosol
profiles from both zones, and the resulting profiles are compared in
fig.~\ref{fig:aero_uniformity}. The mean shifts $\Delta E/E$\xspace and $\Delta X_\text{max}$\xspace are
relatively constant with energy: $\Delta E/E$\xspace$=0.5\%$, and $\Delta X_\text{max}$\xspace is close to zero.
The distributions of $\Delta E/E$\xspace and $\Delta X_\text{max}$\xspace are affected by long tails, with the
RMS in $\Delta E/E$\xspace growing with energy from $3\%$ to $8\%$. For $\Delta X_\text{max}$\xspace, the RMS
for all energies is about 6~$\text{g~cm}^{-2}$\xspace.
\subsection{Corrections for Multiple Scattering}\label{subsec:ms_correction}
Multiply-scattered light, if not accounted for in the reconstruction, will
lead to a systematic overestimate of shower energy and $X_\text{max}$\xspace. This is
because multiple scattering shifts light into the FD field of view that would
otherwise remain outside the shower image. A na\"{i}ve reconstruction will
incorrectly identify multiply-scattered photons as components of the direct
fluorescence/Cherenkov and singly-scattered Cherenkov signals, leading to an
overestimate of the Cherenkov-fluorescence light production used in the
calculation of the shower profile. The mis-reconstruction of $X_\text{max}$\xspace is
similar to what occurs in the case of overestimated optical depths: not
enough scattered light is removed from the low-altitude tail of the shower
profile, causing an overestimate of $dE/dX$ in the deep part of the profile.
The parameterizations of multiple scattering due to
Roberts~\cite{Roberts:2005xv} and Pekala et al.~\cite{Pekala:2009fe} have
been implemented in the hybrid event reconstruction. The predictions from
both analyses are that the scattered light fraction in the shower image will
increase with optical depth, so that distant high-energy showers will be most
affected by multiple scattering. A comparison of showers reconstructed with
and without multiple scattering (fig.~\ref{fig:ms_corrections}) verifies that
the shift in the estimated energy doubles from $2\%$ to nearly $5\%$ as the
shower energy (and therefore, average shower distance to the FD) increases.
The systematic error in the shower maximum is also consistent with the
overestimate of the light signal that occurs without multiple scattering
corrections.
The multiple scattering corrections due to Roberts and Pekala et al. give
rise to small differences in the reconstructed energy and $X_\text{max}$\xspace. As shown in
fig.~\ref{fig:ms_implement}, the two parameterizations differ in the energy
correction by $<1\%$, and there is a shift of $1$~$\text{g~cm}^{-2}$\xspace in $X_\text{max}$\xspace for all
energies. These values provide an estimate of the systematic uncertainties
due to multiple scattering which remain in the reconstruction.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=\textwidth]{dEdXmax_MS}
\caption{\label{fig:ms_corrections} \slshape Overestimates of shower
energy (left) and $X_\text{max}$\xspace (right) due to lack of multiple
scattering corrections in the hybrid reconstruction. The dotted
lines correspond to a reconstruction with multiple scattering
enabled. The uncertainties correspond to the sample RMS in
each energy bin.}
\end{center}
\end{figure}
\begin{figure}[ht]
\begin{center}
\includegraphics[width=\textwidth]{dEdXmax_PekalaRobertsMS}
\caption{\label{fig:ms_implement} \slshape Systematic differences in
shower energy (left) and $X_\text{max}$\xspace (right) for events
reconstructed using the multiple scattering corrections of
Roberts~\cite{Roberts:2005xv} (dotted lines) and Pekala et
al.~\cite{Pekala:2009fe}.}
\end{center}
\end{figure}
\subsection{Summary}\label{subsec:hrec_unc_summary}
\begin{table}[ht]
\def1.5{1.05}
\begin{center}
\footnotesize{
\begin{tabular}{|l|c|c|c|c|c|}
\hline
\multicolumn{6}{|c|}{
\rule[-3mm]{0mm}{8mm}\bfseries Systematic Uncertainties}\\
\hline
\multirow{2}{*}{Source} & \multirow{2}{*}{$\log{(E/\text{eV})}$}
& $\Delta E/E$\xspace & RMS$(\Delta E/E)$
& $\Delta X_\text{max}$\xspace & RMS$(X_\text{max})$ \\
& & (\%) & (\%) & ($\text{g~cm}^{-2}$\xspace) & ($\text{g~cm}^{-2}$\xspace) \\
\hline
\multicolumn{6}{|c|}{\slshape Molecular Light Transmission and Production} \\
\hline
Horiz. Uniformity & $17.7-20.0$ & 1 & 1 & 1 & 2 \\
\hline
Quenching Effects & $17.7-20.0$ & $+5.5$ & \multirow{2}{*}{$1.5-3.0$}
& $-2.0$ & \multirow{2}{*}{$7.2-8.4$} \\
$p$, $T$, $u$ Variability & $17.7-20.0$ & $-0.5$ & & $+2.0$ & \\
\hline
\multicolumn{6}{|c|}{\slshape Aerosol Light Transmission} \\
\hline
\multirow{3}{*}{Optical Depth}
& $<18.0$ & $+3.6$, $-3.0$ & $1.6\pm1.6$ & $+3.3$, $-1.3$ & $3.0\pm3.0$ \\
& $18.0-19.0$ & $+5.1$, $-4.4$ & $1.8\pm1.8$ & $+4.9$, $-2.8$ & $3.7\pm3.7$ \\
& $19.0-20.0$ & $+7.9$, $-7.0$ & $2.5\pm2.5$ & $+7.3$, $-4.8$ & $4.7\pm4.7$ \\
\hline
$\lambda$-Dependence & $17.7-20.0$ & 0.5 & 2.0 & 0.5 & 2.0 \\
\hline
Phase Function & $17.7-20.0$ & 1.0 & 2.0 & 2.0 & 2.5 \\
\hline
\multirow{3}{*}{Horiz. Uniformity}
& $<18.0$ & 0.3 & 3.6 & 0.1 & 5.7 \\
& $18.0-19.0$ & 0.4 & 5.4 & 0.1 & 7.0 \\
& $19.0-20.0$ & 0.2 & 7.4 & 0.4 & 7.6 \\
\hline
\multicolumn{6}{|c|}{\slshape Scattering Corrections} \\
\hline
\multirow{3}{*}{Mult. Scattering}
& $<18.0$ & 0.4 & 0.6 & 1.0 & 0.8 \\
& $18.0-19.0$ & 0.5 & 0.7 & 1.0 & 0.9 \\
& $19.0-20.0$ & 1.0 & 0.8 & 1.2 & 1.1 \\
\hline
\end{tabular}
}
\caption{\slshape Systematic uncertainties in the hybrid reconstruction
due to atmospheric influences on light transmission or
production.}
\label{table:uncertainties}
\end{center}
\def1.5{1.5}
\end{table}
Table~\ref{table:uncertainties} summarizes our estimate of the impact of the
atmosphere on the energy and $X_\text{max}$\xspace measurements of the hybrid detector of the
Pierre Auger Observatory\xspace. Aside from large quenching effects due to missing quenching
corrections in the reconstruction, the systematic uncertainties are currently
dominated by the aerosol optical depth: $4-8\%$ for shower energy, and about
$4-8$~$\text{g~cm}^{-2}$\xspace for $X_\text{max}$\xspace. This list of uncertainties is similar to that
reported in~\cite{Prouza:2007ur}, but now includes an explicit statement of
the multiple scattering correction\footnote{Note that in previous
publications, this correction has been absorbed into a more general $10\%$
systematic uncertainty due to reconstruction
methods~\cite{Dawson:2007di,Abraham:2008ru}.}.
The RMS values in the table can be interpreted as the spread in measurements
of energy and $X_\text{max}$\xspace due to current limitations in the atmospheric monitoring
program. For example, the uncertainties due to the variability of $p$, $T$,
and $u$ are caused by the use of monthly molecular models in the
reconstruction rather than daily measurements, while uncertainties due to the
horizontal non-uniformity of aerosols are due to limited spatial sampling of
the full atmosphere. Note that the RMS values listed for the aerosol optical
depth are due to a mixture of systematic and statistical uncertainties; we
have estimated these contributions conservatively by expressing the RMS as a
central value with large systematic uncertainties. The combined values from
all atmospheric measurements are, approximately, RMS($\Delta E/E$\xspace)$\approx5\pm1\%$
to $9\pm1\%$ as a function of energy, and RMS($X_\text{max}$\xspace)$\approx11\pm1$~$\text{g~cm}^{-2}$\xspace to
$13\pm1$~$\text{g~cm}^{-2}$\xspace. In principle, the RMS can be reduced by improving the
spatial resolution and timing of the atmospheric monitoring data. Such
efforts are underway, and are described in Section~\ref{sec:future}.
\section{Aerosol Measurements at the Pierre Auger Observatory\xspace}\label{sec:measurements}
Several instruments are deployed at the Pierre Auger Observatory\xspace to observe aerosol scattering
properties. The aerosol optical depth is estimated using UV laser
measurements from the CLF, XLF, and scanning lidars
(Section~\ref{subsec:aod_measurements}); the aerosol phase function is
determined with APF monitors (Section~\ref{subsec:aero_scat_measurements});
and the wavelength dependence of the aerosol optical depth is measured with
data recorded by the HAM and FRAM telescopes
(Section~\ref{subsec:aero_wl_measurements}).
\subsection{Optical Depth Measurements}\label{subsec:aod_measurements}
\subsubsection{The Central Laser Facility}\label{subsubsec:central_laser_facility}
The CLF produces calibrated laser ``test beams'' from its location in the
center of the Auger surface detector~\cite{Fick:2006yn,Wiencke:2006hg}.
Located between $26$ and $39$~km from the FD telescopes, the CLF contains a
pulsed $355$~nm laser that fires a depolarized beam in an quarter-hourly
sequence of vertical and inclined shots. Light is scattered out of the
laser beam, and a small fraction of the scattered light is collected by the
FD telescopes. With a nominal energy of $7$~mJ per pulse, the light
produced is roughly equal to the amount of fluorescence light generated by
a $10^{20}$~eV shower. The CLF-FD geometry is shown in
fig.~\ref{fig:laser_geometry}.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=0.4\textwidth,clip]{laser_geometry}
\caption{\label{fig:laser_geometry} \slshape CLF laser and FD geometry.
Vertical shots ($\varphi_1=90^\circ$) are used for the measurement of
$\tau_{a}(h,\lambda_0)$\xspace, with $\lambda_0=355$~nm.}
\end{center}
\end{figure}
The CLF has been in operation since late 2003. Every quarter-hour during
FD data acquisition, the laser fires a set of 50 vertical shots. The
relative energy of each vertical shot is measured by two ``pick-off''
energy probes, and the light profiles recorded by the FD telescopes are
normalized by the probe measurements to account for shot-by-shot changes in
the laser energy. The normalized profiles are then averaged to obtain
hourly light flux profiles, in units of photons~m$^{-2}$~mJ$^{-1}$ per
100~ns at the FD entrance aperture~\cite{Fick:2006yn}. The hourly profiles
are determined for each FD site, reflecting the fact that aerosol
conditions may not be horizontally uniform across the Observatory during
each measurement period.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=0.85\textwidth,clip]{vaod_200401_200812_ll}
\includegraphics*[width=0.85\textwidth,clip]{vaod_200401_200812_lm}
\includegraphics*[width=0.85\textwidth,clip]{vaod_200401_200812_co}
\caption{\label{fig:clfVaod3km} \slshape
Monthly median CLF measurements of the aerosol optical depth
3~km above the fluorescence telescopes at Los Leones,
Los Morados, and Coihueco (January 2004 -- December 2008).
Measurements from Loma Amarilla are not currently available.
The dark and light contours contain 68\% and 95\% of the
measurements, respectively. Hours with optical depths above 0.1
(dashed lines) are characterized by strong haze, and are cut
from the FD analysis.}
\end{center}
\end{figure}
It is possible to determine the vertical aerosol optical depth $\tau_{a}(h,\lambda_0)$\xspace
between the CLF and an FD site by normalizing the observed light flux with
a ``molecular reference'' light profile. The molecular references are
simply averaged CLF laser profiles that are observed by the FD telescopes
during extremely clear viewing conditions with negligible aerosol
attenuation. The references can be identified by the fact that the laser
light flux measured by the telescopes during clear nights is larger than
the flux on nights with aerosol attenuation (after correction for the
relative calibration of the telescopes). Clear-night candidates can also
be identified by comparing the shape of the recorded light profile against
a laser simulation using only Rayleigh scattering~\cite{Valore:2009}. The
candidate nights are then validated by measurements from the APF monitors
and lidar stations.
A minimum of three consecutive clear hours are used to construct each
reference profile. Once an hourly profile is normalized by a
clear-condition reference, the attenuation of the remaining light is due
primarily to aerosol scattering along the path from the CLF beam to the
telescopes. The optical depth $\tau_{a}(h,\lambda_0)$\xspace can be extracted from the
normalized hourly profiles using the methods described
in~\cite{Abbasi:2005gi}.
Note that the lower elevation limit of the FD telescopes ($1.8^\circ$)
means that the lowest $1$~km of the vertical laser beam is not within
the telescope field of view (see fig.~\ref{fig:vaod_transmission}). While
the CLF can be used to determine the total optical depth between the ground
and $1$~km, the vertical distribution of aerosols in the lowest part of the
atmosphere cannot be observed. Therefore, the optical depth in this region
is constructed using a linear interpolation between ground level, where
$\tau_a$ is zero, and $\tau_a(1~\text{km},\lambda_0)$.
The normalizations used in the determination of $\tau_{a}(h,\lambda_0)$\xspace mean that the
analysis does not depend on the absolute photometric calibration of either
the CLF or the FD, but instead on the accuracy of relative calibrations of
the laser and the FD telescopes.
The sources of uncertainty that contribute to the normalized hourly
profiles include the clear night references ($3\%$)\footnote{The value
$3\%$ contains the statistical and calibration uncertainties in a given
reference profile, but does not describe an uncertainty in the selection of
the reference. This uncertainty will be quantified in a future end-to-end
analysis of CLF data using simulated laser shots.}, uncertainties in the FD
relative calibration ($3\%$), and the accuracy of the laser energy
measurement ($3\%$). Statistical fluctuations in the hourly average light
profiles contribute additional relative uncertainties of $1\%-3\%$ to the
normalized hourly light flux. The uncertainties in $\tau_{a}(h,\lambda_0)$\xspace plotted in
fig.~\ref{fig:vaod_transmission} derive from these sources, and are highly
correlated due to the systematic uncertainties.
Between January 2004 and December 2008, over 6,000 site-hours of optical
depth profiles have been analyzed using measurements of more than one
million CLF shots. Figure~\ref{fig:clfVaod3km} depicts the distribution of
$\tau_{a}(h)$\xspace recorded using the FD telescopes at Los Leones, Los Morados, and
Coihueco. The data $3$~km above ground level are shown, since this
altitude is typically above the aerosol mixing layer. A moderate seasonal
dependence is apparent in the aerosol distributions, with austral summer
marked by more haze than winter. The distributions are asymmetric, with
long tails extending from the relatively clear conditions
($\tau_{a}(3~\text{km})<0.04$) characteristic of most hours to periods of
significant haze ($\tau_{a}(3~\text{km})>0.1$).
Approximately $5\%$ of CLF measurements have optical depths greater than
0.1. To avoid making very large corrections to the expected light flux
from distant showers, these hours are typically not used in the FD
analysis.
\subsubsection{Lidar Observations}
In addition to the CLF, four scanning lidar stations are operated
at the Pierre Auger Observatory\xspace to record $\tau_{a}(h,\lambda_0)$\xspace from every FD site~\cite{BenZvi:2006xb}.
Each station has a steerable frame that holds a pulsed $351$~nm laser,
three parabolic mirrors, and three PMTs. The frame is mounted atop a
shipping container which contains data acquisition electronics. The
station at Los Leones includes a separate, vertically-pointing Raman lidar
test system, which can be used to detect aerosols and the relative
concentration of N$_2$ and O$_2$ in the atmosphere.
During FD data acquisition, the lidar telescopes sweep the sky in a set
hourly pattern, pulsing the laser at 333~Hz and observing the backscattered
light with the optical receivers. By treating the altitude distribution of
aerosols near each lidar station as horizontally uniform, $\tau_{a}(h,\lambda_0)$\xspace can be
estimated from the differences in the backscattered laser signal recorded
at different zenith angles~\cite{Filipcic:2002ba}. When non-uniformities
such as clouds enter the lidar sweep region, the optical depth can still be
determined up to the altitude of the non-uniformity.
Since the lidar hardware and measurement techniques are independent of the
CLF, the two systems have essentially uncorrelated systematic
uncertainties. With the exception of a short hourly burst of horizontal
shots toward the CLF and a shoot-the-shower mode (Section~\ref{subsec:sts})
~\cite{BenZvi:2006xb}, the lidar sweeps occur outside the FD field of view
to avoid triggering the detector with backscattered laser light. Thus, for
many lidar sweeps, the extent to which the lidars and CLF measure similar
aerosol profiles depends on the true horizontal uniformity of aerosol
conditions at the Observatory.
\begin{figure}
\begin{center}
\includegraphics*[width=\textwidth,clip]{lidar_clf_profile_comparison}
\caption{\label{fig:lidar_clf_profile_comparison} \slshape
An hourly aerosol optical depth profile observed by the CLF and
the Coihueco lidar station for relatively dirty conditions in
December 2006. The gray band depicts the systematic uncertainty
in the lidar aerosol profile.}
\end{center}
\end{figure}
Figure~\ref{fig:lidar_clf_profile_comparison} shows a lidar measurement of
$\tau_{a}(h,\lambda_0)$\xspace with vertical shots and the corresponding CLF aerosol profile
during a period of relatively high uniformity and low atmospheric clarity.
The two measurements are in good agreement up to $5$~km, in the region
where aerosol attenuation has the greatest impact on FD observations.
Despite the large differences in the operation, analysis, and viewing
regions of the lidar and CLF, the optical measurements from the two
instruments typically agree within their respective
uncertainties~\cite{BenZvi:2007it}.
\subsubsection{Aerosol Optical Depth Uniformity}
The FD building at Los Leones is located at an altitude of $1420$~m, on a
hill about $15$~m above the surrounding plain, while the Coihueco site is
on a ridge at altitude $1690$~m, a few hundred meters above the valley
floor. Since the distribution of aerosols follows the prevailing ground
level rather than local irregularities, it is reasonable to expect that the
aerosol optical depth between Coihueco and a fixed altitude will be
systematically lower than the aerosol optical depth between Los Leones and
the same altitude. The data in fig.~\ref{fig:aerosol_uniformity} (left
panel) support this expectation, and show that aerosol conditions differ
significantly and systematically between these FD sites. In contrast,
optical depths measured at nearly equal altitudes, such as Los Leones and
Los Morados (1420~m), are quite similar.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=0.49\textwidth,clip]{vaod_co_vs_ll}
\includegraphics*[width=0.49\textwidth,clip]{vaod_lm_vs_ll}
\caption{\label{fig:aerosol_uniformity} \slshape
Comparison of the aerosol optical depths measured with CLF shots at
Los Leones, Los Morados, and Coihueco. The buildings at Los Leones
and Los Morados are located on low hills at similar altitudes, while
the Coihueco FD building is on a large hill $200$~m above the other
sites. The solid lines indicate equal optical depths at two
sites, while the dotted lines show the best linear fits to the
optical depths. The bottom panels show histograms of the differences
between the optical depths.}
\end{center}
\end{figure}
Unlike for the molecular atmosphere, it is not possible to assume a
horizontally uniform distribution of aerosols across the Observatory. To
handle the non-uniformity of aerosols between sites, the FD reconstruction
divides the array into aerosol ``zones'' centered on the midpoints between
the FD buildings and the CLF. Within each zone, the vertical distribution
of aerosols is treated as horizontally uniform by the reconstruction (i.e.,
eq.~\eqref{eq:aerotrans} is applied).
\subsection{Scattering Measurements}\label{subsec:aero_scat_measurements}
Aerosol scattering is described by the phase function $P_a(\theta)$, and
the hybrid reconstruction uses the functional form given in
equation~\eqref{eq:aero_phase_function}. As explained in
Section~\ref{subsubsec:mol_aero_scattering}, the aerosol phase function for
each hour must be determined with direct measurements of scattering in the
atmosphere, which can be used to infer the backscattering and asymmetry
parameters $f$ and $g$ of $P_a(\theta)$.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=0.49\textwidth,clip]{pf_jun_2006}
\includegraphics*[width=0.49\textwidth,clip]{pf_jul_2008}
\caption{\label{fig:apf_fits} \slshape
Light scattering measurements with the APF Xenon flasher at Coihueco.
During a clear night (left), the observed phase function is symmetric
due to the predominance of molecular scattering. An asymmetric phase
function on a different night (right) indicates the presence of
aerosols.}
\end{center}
\end{figure}
At the Auger Observatory, these quantities are measured by two Aerosol
Phase Function monitors, or APFs, located about $1$~km from the FD buildings
at Coihueco and Los Morados~\cite{BenZvi:2007px}. Each APF uses a
collimated Xenon flash lamp to fire an hourly sequence of $350$~nm and
$390$~nm shots horizontally across the FD field of view. The shots are
recorded during FD data acquisition, and provide a measurement of
scattering at angles between $30^\circ$ and $150^\circ$. A fit to the
horizontal track seen by the FD is sufficient to determine $f$ and $g$.
The APF light signal from two different nights is depicted in
fig.~\ref{fig:apf_fits}, showing the total phase function fit and
$P_a(\theta)$ after the molecular component has been subtracted.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=\textwidth,clip]{apfFitChi2G}
\caption{\label{fig:apfFitChi2G} \slshape
Left: distribution of the figure of merit for fits of
equation~\eqref{eq:aero_phase_function} to APF measurements at
Coihueco, June 2006 -- July 2008. Right: distribution of the
asymmetry parameter $g$ measured at $350$~nm and $390$~nm at
Coihueco.}
\end{center}
\end{figure}
The phase function asymmetry parameter $g$ measured at Coihueco between
June 2006 and July 2008 is shown in fig.~\ref{fig:apfFitChi2G}. The value
of $g$ was determined by fitting the modified Henyey-Greenstein function of
eq.~\eqref{eq:aero_phase_function} to the APF data. The reduced-$\chi^2$
distribution for this fit, also shown in the figure, indicates that the
Henyey-Greenstein function describes aerosol scattering in the FD
reasonably well. The measurements at Coihueco yield a site average
$\langle g\rangle=0.56\pm0.10$ for the local asymmetry parameter, excluding
clear nights without aerosol attenuation. On clear (or nearly clear)
nights, we estimate $g=0$ with an uncertainty of $0.2$. The distribution
of $g$ in Malarg\"{u}e\xspace, a desert location with significant levels of sand and
volcanic dust, is comparable to measurements reported in the literature for
similar climates~\cite{2006JGRD..11105S04A}.
Approximately 900 hours of phase function data have been recorded with
both APF monitors since June 2006. The sparse data mean that it is not
possible to use a true measurement of $P_a(\theta)$ for most FD events.
Therefore, the Coihueco site average is currently used as the estimate of
the phase function for all aerosol zones, for all cosmic ray events. The
systematic uncertainty introduced by this assumption will be explored in
Section~\ref{subsubsec:aero_error_prop}.
\subsection{Wavelength Dependence}\label{subsec:aero_wl_measurements}
Measurements of the wavelength dependence of aerosol transmission are used
to determine the {\AA}ngstr{\o}m\xspace exponent $\gamma$ defined in
equation~\eqref{eq:wavelength_dependence}. At the Pierre Auger Observatory\xspace, observations of
$\gamma$ are performed by two instruments: the Horizontal Attenuation
Monitor, or HAM; and the (F/ph)otometric Robotic Telescope for Astronomical
Monitoring, also known as FRAM~\cite{BenZvi:2007it,BenZvi:2007uj}.
The HAM uses a high intensity discharge lamp located at Coihueco to provide
an intense broad band light source for a CCD camera placed at Los Leones,
about $45$~km distant. This configuration allows the HAM to measure the
total horizontal atmospheric attenuation across the Observatory. To
determine the wavelength dependence of the attenuation, the camera uses a
filter wheel to record the source image at five wavelengths between $350$
and $550$~nm. By fitting the observed intensity as a function of
wavelength, subtracting the estimated molecular attenuation, and assuming
an aerosol dependence of the form of
equation~\eqref{eq:wavelength_dependence}, it is possible to determine the
{\AA}ngstr{\o}m\xspace exponent $\gamma$ of aerosol attenuation. During 2006 and 2007, the
average exponent observed by the HAM was $\gamma=0.7$ with an RMS of $0.5$
due to the non-Gaussian distribution of measurements~\cite{BenZvi:2007it}.
The relatively small value of $\gamma$ suggests that Malarg\"{u}e\xspace has a large
component of coarse-mode aerosols. This is consistent with physical
expectations and other measurements in desert-like
environments~\cite{Eck:1999,Kaskaoutis:2006}.
Like the APF monitor data, the HAM and FRAM results are too sparse to use
in the full reconstruction; therefore, during the FD reconstruction, the
HAM site average for $\gamma$ is applied to all FD events in every aerosol
zone. The result of this approximation is described in the next section.
\section{Molecular Measurements at the Pierre Auger Observatory\xspace}\label{sec:molecular_effects}
\subsection{Profile Measurements with Weather Stations and Radiosondes}\label{subsec:profile_measurements}
The vertical profiles of atmospheric parameters (pressure, temperature, etc.)
vary with geographic location and with time so that a global static model of
the atmosphere is not appropriate for precise shower studies. At a given
location, the daily variation of the atmospheric profiles can be as large as
the variation in the seasonal average conditions. Therefore, daily
measurements of atmospheric profiles are desirable.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=0.85\textwidth,clip]{weather_200406_200812_temp}
\includegraphics*[width=0.85\textwidth,clip]{weather_200406_200812_pres}
\includegraphics*[width=0.85\textwidth,clip]{weather_200406_200812_hum}
\caption{\label{fig:weather_ground} \slshape
Monthly median ground temperature, pressure, and water vapor
pressure observed at the CLF weather station ($1.4$~km above
sea level), showing the distributions of 68\% and 95\% of the
measurements as dark and light gray contours, respectively. The
vapor pressure has been calculated using measurements of
the temperature and relative humidity.}
\end{center}
\end{figure}
Several measurements of the molecular component of the atmosphere are
performed at the Pierre Auger Observatory\xspace. Near each FD site and the CLF, ground-based weather
stations are used to record the temperature, pressure, relative humidity, and
wind speed every five minutes. The first weather station was commissioned at
Los Leones in January 2002, followed by stations at the CLF (June 2004), Los
Morados (May 2007), and Loma Amarilla (November 2007). The station at
Coihueco is installed but not currently operational. Data from the CLF
station are shown in fig.~\ref{fig:weather_ground}; the measurements are
accurate to $0.2-0.5^\circ$C in temperature, $0.2-0.5$~hPa in pressure, and
$2\%$ in relative humidity~\cite{Campbell:website}. The pressure and
temperature data from the weather stations are used to monitor the weather
dependence of the shower signal observed by the
SD~\cite{Bleve:2007mb,Bleve:2009}. They can also be used to characterize the
horizontal uniformity of the molecular atmosphere, which is assumed in
eq.~\eqref{eq:aerotrans}.
Of more direct interest to the FD reconstruction are measurements of the
altitude dependence of the pressure and temperature, which can be used in
eq.~\eqref{eq:scatter_profile} to estimate the vertical molecular optical
depth. These measurements are performed with balloon-borne radiosonde
flights, which began in mid-2002 and are currently launched one or two times
per week. The radiosonde measurements include relative humidity and wind
data recorded about every $20$~m up to an average altitude of $25$~km, well
above the fiducial volume of the fluorescence detectors. The accuracy of the
measurements are approximately $0.2^\circ\text{C}$ for temperature,
$0.5-1.0$~hPa for pressure, and $5\%$ for relative
humidity~\cite{GRAW:website}.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=\textwidth,clip]{balloonSeasonalX_BW}
\caption{\label{fig:balloonX} \slshape
Radiosonde measurements of the depth profile above Malarg\"{u}e\xspace recorded
during 261 balloon flights between 2002 and 2009. The data are
plotted as deviations from the average profile of all 261
flights, and are grouped by season. The dark lines indicate the
seasonal averages, and the vertical dashed lines correspond to
the height of Malarg\"{u}e\xspace above sea level.}
\end{center}
\end{figure}
The balloon observations demonstrate that daily variations in the temperature
and pressure profiles depend strongly on the season, with more stable
conditions during the austral summer than in winter~\cite{Keilhauer:2005ja}.
The atmospheric depth profile $X(h)$ exhibits significant altitude-dependent
fluctuations. The largest daily fluctuations are typically $5$~$\text{g~cm}^{-2}$\xspace
observed at ground level, increasing to $10-15$~$\text{g~cm}^{-2}$\xspace between $6$ and
$12$~km altitude. The seasonal differences between summer and winter can be
as large as $20$~$\text{g~cm}^{-2}$\xspace on the ground, increasing to $30$~$\text{g~cm}^{-2}$\xspace at higher
altitudes (fig.~\ref{fig:balloonX}).
\subsection{Monthly Average Models}\label{subsec:monthly_models}
Balloon-borne radiosondes have proven to be a reliable means of measuring the
state variables of the atmosphere, but nightly balloon launches are too
difficult and expensive to carry out with regularity in Malarg\"{u}e\xspace. Therefore, it
is necessary to sacrifice some time resolution in the vertical profile
measurements and use models which quantify the average molecular profile over
limited time intervals.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=\textwidth,clip]{monthly_profiles_nMMM}
\caption{\label{fig:monthly_profiles} \slshape
Left: average profile $X(h)$ above Malarg\"{u}e\xspace, with the altitude of the
site indicated by the vertical dotted line. Right:
deviation of the monthly mean values of $X(h)$ from the yearly
average as a function of month. Data are from the mean monthly
weather models (updated through 2009).}
\end{center}
\end{figure}
Such time-averaged models have been generated for the FD reconstruction using
$261$ local radiosonde measurements conducted between August 2002 and
December 2008. The monthly profiles include average values for the
atmospheric depth, density, pressure, temperature, and humidity as a function
of altitude. Figure~\ref{fig:monthly_profiles} depicts a plot of the annual
mean depth profile $X(h)$ in Malarg\"{u}e\xspace, as well as the deviation of the monthly
model profiles from the annual average. The uncertainties in the monthly
models, not shown in the figure, represent the typical range of conditions
observed during the course of each month. At ground level, the RMS
uncertainties are approximately $3$~$\text{g~cm}^{-2}$\xspace in austral summer and $6$~$\text{g~cm}^{-2}$\xspace
during austral winter; near $10$~km altitude, the uncertainties are
$4$~$\text{g~cm}^{-2}$\xspace in austral summer and $8$~$\text{g~cm}^{-2}$\xspace in austral winter.
The use of monthly averages rather than daily measurements introduces
uncertainties into measurements of shower energies $E$ and shower maxima
$X_\text{max}$\xspace; the magnitudes of the effects are estimated in
Section~\ref{subsec:mol_uncertainties}.
\subsection{Horizontal Uniformity of the Molecular Atmosphere}\label{subsec:horizontal_molecular}
The assumption of horizontally uniform atmospheric layers implied by
equation~\eqref{eq:aerotrans} reduces the estimate of atmospheric
transmission to a simple geometrical calculation, but the deviation of the
atmosphere from true horizontal uniformity introduces some systematic error
into the transmission. An estimate of this deviation is required to
calculate its impact on air shower reconstruction.
\begin{figure}[ht]
\begin{center}
\includegraphics*[width=0.85\textwidth,clip]{diff_200406_200812_temp}
\includegraphics*[width=0.85\textwidth,clip]{diff_200406_200812_pres}
\includegraphics*[width=0.85\textwidth,clip]{diff_200406_200812_hum}
\caption{\label{fig:diff_weather_ground} \slshape
Monthly differences in the ground temperature, pressure, and vapor
pressure observed with the weather stations at Los Leones (LL) and the
CLF. The dark and light gray contours contain 68\% and 95\% of the
measurement differences. Gaps in the comparison during 2007 were
caused by equipment failures in the station at Los Leones.}
\end{center}
\end{figure}
For the molecular component of the atmosphere, the data from different
ground-based weather stations provide a convenient, though limited, check of
weather differences across the Observatory. For example, the differences
between the temperature, pressure, and vapor pressure measured using the
weather stations at Los Leones and the CLF are plotted in
fig.~\ref{fig:diff_weather_ground}. The altitude difference between the
stations is approximately $10$~m, and they are separated by $26$~km, or
roughly half the diameter of the SD. Despite the large horizontal separation
of the sites, the measurements are in close agreement. Note that the
differences in the vapor pressure are larger than the differences in total
pressure, due to the lower accuracy of the relative humidity measurements.
It is quite difficult to check the molecular uniformity at higher altitudes,
with, for example, multiple simultaneous balloon launches. The measurements
from the network of weather stations at the Observatory are currently the
only indications of the long-term uniformity of molecular conditions across
the site. Based on these observations, the molecular atmosphere is treated
as uniform.
|
\section{\label{}}
\section{Introduction}
There are four primary methods for probing the structure and physics of relativistic jets in
blazars on parsec and sub-parsec scales: VLBI imaging in both total and polarized intensity
at millimeter wavelengths, variability of the flux at radio through $\gamma$-ray frequencies,
polarization at radio through optical wavebands, and the spectral energy distribution (SED).
In the past, efforts to study this capricious class of objects have been limited by datasets
with substantial gaps in either time or frequency coverage. Starting in the mid-1990s, this
situation has been greatly alleviated by the availability of instruments such as the Very Long
Baseline Array (VLBA), the {\it Rossi} X-ray Timing Explorer (RXTE), and, most recently, the
{\it Fermi} Gamma-ray Space Telescope. Together with concerted efforts with optical/near-infrared,
millimeter-wave, and radio telescopes, these facilities now allow long-term, comprehensive
multi-waveband monitoring of a number of blazars. Such programs are providing valuable insights
into the processes by which jets form and propagate, as well as the locations and physics of flux
outbursts in blazars \citep[e.g.,][]{Mar08,Chat08,Lar08}.
We are leading a monitoring program that provides the data
needed to combine the techniques listed above in order to locate the sites of high-energy emission,
determine the processes by which X-rays and $\gamma$-rays are produced, and infer the physical
conditions of the jet, including the geometry of its magnetic field. Because of limitations in
sensitivity and available observing time, such comprehensive monitoring can be mounted
only for a relatively small number of objects, 34 in the case of our program. Progress toward an
overall understanding of blazars therefore involves generalization of inferences drawn from the
well-studied specimens into a framework that can be applied to the entire class.
We present here a rich set of observations of the quasar PKS~1510$-$089 ($z$ = 0.361) that
allows us to locate the sites of $\gamma$-ray flares, leading to inferences on the
high-energy emission processes. PKS~1510$-$089 possesses all of the
characteristics of blazars: flat radio spectrum, apparent superluminal motion --- among the fastest
\citep[as high as $45c$ for H$_\circ$=71 km s$^{-1}$Mpc$^{-1}$ and the concordance
cosmology][]{Spergel07} of all blazars observed thus far \citep{J05} --- high-amplitude and rapid
flux variability at all wavebands, strong and variable optical linear polarization, and high
$\gamma$-ray apparent luminosity \citep{Hartman99}. The observations of this blazar from our
program allow us to compare motions, linear polarization, and changing flux of features in the
parsec-scale radio jet with flux variability of the entire source at radio, millimeter-wave, IR,
optical, X-ray, and $\gamma$-ray frequencies, and with optical polarization. The relative timing of
the correlated variations thus found probe the structure and physics of the innermost jet regions
where the flow is accelerated and collimated, and where the emitting electrons are energized.
\section{Observational Results}
Figure \ref{fig1} presents the {\it Fermi}
Large Area Telescope $\gamma$-ray light curve, the R-band optical light curve, and the
optical polarization as a function of time during the first half of 2009, with the data taken
from the above paper. Details of our observations of PKS~1510$-$089 are given in \citet{Mar10},
which includes data from 2006 to 2009. That paper also presents 43 GHz VLBA images showing
that a bright superluminal ($22c$) knot (A) passed through the core on JD 2454959$\pm$4.
Additional VLBA images (see Fig. \ref{fig2}) reveal that another knot (B) with the same apparent
speed was coincident with the core on JD 2455008$\pm$8. We stress that the ``core'' of a blazar
seen on VLBI images lies parsecs downstream of the location of the black hole \citep{Mar02}.
There is mounting evidence that, barring optical depth effects, the core represents emission
from jet plasma that has been compressed and energized by a system of one or more
standing conical shocks owing to pressure mismatches between the jet and the external medium.
One of the main questions that our observational program can answer is whether the high-energy
emission occurs between the central engine and core, within the core, or downstream of the core.
In order to do this, we must first establish an association between a superluminal knot and
high-energy flares, then analyze the timing of the flares relative to the motion of the knot.
In the case of BL~Lac in 2005, we employed this method to
determine that there are multiple sites of X-ray and optical
flares in the jet: upstream of the core where the magnetic field has
a helical geometry, and within the core. Flares occur as a disturbance in the jet flow
(an ``emission feature'') passes through these regions \citep{Mar08}.
Eight major flares are apparent in Figure \ref{fig1}, which covers the first half of 2009. The
figure numbers the flares. (We note that $\gamma$-ray emission was detected at an elevated level
throughout the 50+ days following flare 8, but one cannot isolate distinct flares during this
period.) Of particular interest is the 51-day period from JD 2454901 to 2454962 during which
flares 3-8 took place. As seen in Figure \ref{fig1}, the optical polarization position angle
$\chi$ rotated by $\sim~720^\circ$. (Since there is no distinction between $\chi$ and
$\chi\pm 180^\circ n$, we select $n$ such that the jump in $\chi$ is minimized.) The ratio
of $\gamma$-ray 0.1-200 GeV integrated flux to bolometric synchrotron flux \citep[see][]{Mar10}
for flares 1--8 was 70, 30, 40, 40, 30, 10, 40, and 9, respectively.
Superluminal knot A passed through the core at the same time (within the uncertainties) as the
very sharp (intra-day), extremely high-amplitude $\gamma$-ray and optical flare on JD
2454962. The rotation of the optical polarization vector ended at the same time. We therefore
surmise that the $\gamma$-ray flares occurred upstream of and within the core. After this
point, the $\gamma$-ray flux remained elevated, which suggests that $\gamma$-ray emission
continued downstream of the core, as inferred previously from observations with EGRET, the
VLBA, and single-dish radio antennas \citep{J01,LV03}.
\section{Interpretation}
Although a turbulent magnetic field can produce such apparent rotations of the polarization
vector by $720^\circ$ or more via a random
walk \citep{all85,jones88,darc07}, the beginning and end of such a rotation should occur randomly,
whereas the observed event matches quite closely the duration of the flux outburst. We therefore
conclude that a single coherent emission feature was responsible for the entire outburst. We
appeal to the same model as for BL~Lac \citep{Mar08}: as a disturbance in the flow moves down
the jet, it follows a spiral streamline \citep[as predicted for a model in which the jet is
magnetically accelerated and collimated;][]{vlah06}. It first travels through the acceleration
and collimation zone, which possesses a tightly wound helical magnetic field. If the emission
feature covers most, but not all, of the cross-section of the jet, much of the polarization
is cancelled out from mutually perpendicular magnetic field orientations around the
circumference. The residual polarization vector rotates as the feature executes its
spiral motion.
We have performed a rough calculation of the rotation of $\chi$ in such a model, matching the
parameters (probably non-uniquely) to the PKS~1510$-$089 data. The bulk Lorentz
factor increases from 8
to 24 during the rotation event, causing the Doppler factor to increase from 15 to 38 for an
angle between the jet axis and the line of sight of $1.4^\circ$ \citep{J05}. The acceleration
of the jet causes an increase with time in the rate of rotation, as observed. The circulation
of the centroid of the emission feature around the jet axis causes $\chi$ (corrected for
relativistic aberration) to undulate with time about the trend of monotonic rotation. In the
quantitative model, the
core lies 17 pc downstream of the point where the outburst containing flares 3-8 began.
The magnetic field is in the range of 1 to 0.2 G during the outburst.
The dichotomy in the ratio of $\gamma$-ray to synchrotron flux of the various flares suggests
that different sources of seed photons dominate the inverse Compton scattering as the
emission feature moves down the jet. Flares with relatively low values of this ratio can
be caused by an increase in the number of GeV electrons, which causes enhanced synchrotron,
synchrotron self-Compton (SSC), and external Compton (EC) emission. If SSC is the main
mechanism, second-order scattering is probably important, since the lowest $\gamma$-ray to
synchrotron flux ratio, 9, still exceeds unity. There are some flares (1, 3, and 7) where there
is little or no increase in the optical flux. We infer that these correspond to the emission
feature encountering local sources of enhanced IR-optical-UV seed photon density. We
estimate the luminosity of these sources to be $\sim 3\times 10^{43}$ erg/s, too high
to correspond to any likely cosmic source except a slower (but still probably relativistic)
sheath of the jet \citep[see][]{gtc05}. The source could, for example, be a shock in the
outer periphery of the jet.
Figure \ref{fig3} sketches the general picture that we propose. The $\gamma$-ray emission undergoes
flares at various sites along the jet, with events 3-7 representing locations where either
the sheath produces large numbers of seed photons or electrons are accelerated to
GeV energies. Flare 8 takes place as the moving emission feature
becomes compressed and energized by standing shocks in the core. The elevated post-outburst
$\gamma$-ray emission first comes from knot A as it propagates downstream of the core. At
some point --- probably near JD 2454990 --- emission from knot B becomes dominant. This is
weaker than knot A, so the flares are of lower amplitude. But we do see a $\sim~180^\circ$
rotation of the polarization vector {\it in the same direction and with the same slope as
the previous, longer rotation.} This repetition of the pattern is a firm requirement of
our model: the physical structure of the spine or sheath of the jet
could change on a time scale of years, but not months. In contrast, a turbulent field
model predicts that both the sense and rate of rotation should vary from one event to another.
It is possible that flares 1-2 correspond to the disturbance passing through the rich
seed photon field near the accretion disk and within the broad emission-line region (BELR). If
this is the case, then there is an extended region in the jet beyond the BELR where
electrons are not accelerated efficiently to energies exceeding $\sim 1$ GeV.
\section{Conclusions}
We are finally at the point where the richness of our datasets is sufficient for us to draw
grand inferences about the locations and physical mechanisms of $\gamma$-ray flares. If
PKS~1510$-$089 and BL~Lac are typical, it is no wonder that previous, less comprehensive
monitoring programs left us confused! There are multiple sources of seed photons, some of
which are quite local (as opposed to more global, such as an IR-emitting hot dust torus)
and may not radiate at sufficiently high luminosities to be directly observable. These
include the accretion disk and BELR --- which may be important during the earliest part of a
multi-flare outburst --- shocks or other features in the sheath of the jet, and synchrotron
radiation from the fast spine of the jet that contains the electrons that scatter the seed
photons to $\gamma$-ray energies. A single superluminal knot can be responsible for
a number of flares and periods of sustained elevated $\gamma$-ray emission.
As we accumulate data for more blazars, we expect to see similar patterns of flares both
before and after a disturbance passes through the core of the jet. We also hope to
find deviations from this pattern that serve to provide further insight into the range
of physical behavior of blazars. In both cases, we anticipate a tremendous increase
in our understanding of the relativistic jets of blazars by combining the great power
of monitoring the flux, polarization, and sub-milliarcsecond structure of these exciting
objects.
\vspace{-5mm}
\begin{acknowledgments}
This research was funded in part by NASA grants NNX08AV65G, NNX08AV61G,
NNX08AW56G, and NNX08AJ64G, NSF grant AST-0907893,
Russian RFBR grant~09-02-0092, Spanish ``Ministerio de Ciencia e Innovaci\'on'' grant
AYA2007-67626-C03-03, and Georgian NSF grant GNSF/ST08/4-404. The VLBA is an
instrument of the National Radio Astronomy Observatory, a facility of the NSF,
operated under cooperative agreement by Associated Universities, Inc.
\end{acknowledgments}
\vspace{-5mm}
|
\section{Introduction}
\subsection{Definition of the partition algebra}
Let $M = \{1, 2, \ldots, n\}$ be
a set of $n$ symbols and $F = \{1', \ldots, n'\}$ another set
of $n$ symbols.
We assume that the elements of $M$ and $F$
are ordered by $1<2<\cdots <n$ and $1'<2'<\cdots<n'$ respectively.
Consider the following set of set partitions:
\begin{eqnarray}
\Sigma_n^1
&=&
\{\{T_1, \ldots, T_s\}\ |\ s=1,2, \ldots \ ,\nonumber\\
& & \quad T_j(\neq\emptyset)\subset M\cup F\
(j = 1, 2, \ldots, s),\\
& & \quad \cup T_j = M\cup F,\quad T_i\cap T_j = \emptyset \mbox{ if }
i\neq j\nonumber\}.
\end{eqnarray}
We call an element $w$ of $\Sigma_n^1$ {\em a seat-plan}
and each element of $w$ a {\em part} of $w$.
It is easy to see that the number of seat-plans is equal
to $B_{2n}$, the Bell number.
For $w\in\Sigma_n^1$
consider a rectangle with $n$ marked points on the bottom and the same
$n$ on the top as in Figure~\ref{fig:plan}.
\fig{fig:plan}{A seat-plan of $\Sigma_5$}{1.eps}
The $n$ marked points on the top are labeled by
$1, 2, \ldots n$
from left to right.
Similarly, the $n$ marked points on the bottom are labeled by
$1', 2', \ldots, n'$.
If $w$ consists of $s$ parts,
then put $s$ shaded circles
in the middle of the rectangle
so that they have no intersections.
Then we join the $2n$ marked points and the $s$ circles
with $2n$ shaded bands so that
each shaded circle represent a part of $w$.
Using these diagrams, for $w_1,w_2\in\Sigma_n^1$,
an arbitrary pair of seat-plans,
we can define a product $w_1 w_2$.
The product is obtained by placing $w_1$ on $w_2$,
gluing the corresponding boundaries
and shrinking half along the vertical axis.
We then have a new diagram possibly containing some
shaded regions which are not connected to the boundaries.
If the resulting diagram has $p$ such regions,
then the product is defined by
the diagram with such region removed
and multiplied by $Q^p$.
Here $Q$ is an indeterminate.
(It is easily checked that the product defined above is closed
in the linear span of the set of seat-plans $\Sigma_n^1$
over $\mathbb{Z}[Q]$.)
For example, if
\[
w_1 = \{\{1,1',4'\}, \{2,5\},\{3,4\}, \{2'\},\{3',5'\}\}\in\Sigma_5^1
\]
and
\[
w_2 = \{\{1,1',3',4'\}, \{2\}, \{3,5\}, \{4\}, \{2',5'\}\}\in\Sigma_5^1,
\]
then we have
\[
w_1w_2 =
Q^2\{\{ 1,1',3',4'\}, \{2,5\}, \{3,4\}, \{2',5'\}\}
\in \mathbb{Z}[Q]\Sigma_5^1
\]
as in Figure~\ref{fig:prod}.
\fig{fig:prod}{The product of seat-plans}{2.eps}
By this product,
the set of linear combinations of the elements of $\Sigma_n^1$
over $\mathbb{Z}[Q]$ makes an algebra $A_{n}(Q)$
called the {\em partition algebra}.
The identity of $A_{n}(Q)$
is a diagram which corresponds to the partition
\[
1 = \{\{1, 1'\}, \{2, 2'\}, \ldots, \{n, n'\}\}.
\]
We put $A_{0}(Q) = A_{1}(Q)= \mathbb{Z}[Q]$.
We can define $A_{n}(Q)$ more rigorously in terms of the set partitions
(See P.~P.~Maritin's paper~\cite{Ma2}).
Next we define special elements $s_i, f_i$ ($1\leq i \leq n-1$)
and $e_i$ ($1\leq i\leq n$) of $\Sigma_n^1$
by
\begin{eqnarray*}
s_i &=& \{\{1,1'\},\ldots, \{i-1,(i-1)'\},
\{i+2, (i+2)'\},\ldots, \{n, n'\},\\
& &\quad \{i, (i+1)'\}, \{i+1, i'\}\}\\
f_i &=& \{\{1,1'\},\ldots, \{i-1,(i-1)'\},
\{i+2, (i+2)'\},\ldots, \{n, n'\},\\
& &\quad \{i, i+1, i', (i+1)'\}\}\\
e_i &=& \{\{1,1'\},\ldots, \{i-1,(i-1)'\}, \{i\}, \{i'\}
\{i+1, (i+1)'\},\ldots, \{n, n'\}\}.
\end{eqnarray*}
The diagrams of these special elements are
illustrated by the figures in Figure~\ref{fig:gen}.
Note that in the picture of $e_i$,
there exist ``a male'' only part and
``a female'' only part.
We call such a part ``defective'' (see Section~3.1).
\fig{fig:gen}{Special elements}{3.eps}
We easily find that they satisfy the following basic relations.
\begin{equation}
\begin{array}{rcl}\tag{$R0$}
f_{i+1} &=& s_is_{i+1}f_is_{i+1}s_i \quad (i = 1, 2, \ldots, n-2),\\
e_{i+1} &=& s_ie_{i}s_i \quad (i = 1, 2, \cdots, n-1)
\end{array}
\end{equation}
\begin{equation}
\begin{array}{rcl}\tag{$R1$}
s_i^2 &=& 1 \quad (i = 1, 2, \ldots, n-1),\\
s_is_{i+1}s_i &=& s_{i+1}s_is_{i+1} \quad (i = 1, 2, \ldots, n-2),\\
s_is_j &=& s_js_i \quad (|i-j|\geq 2),
\end{array}
\end{equation}
\begin{equation}
f_i^2 = f_i,\ f_if_j = f_jf_i,\tag{$R2$}
\end{equation}
\begin{equation}
f_is_i = s_if_i=f_i,\tag{$R3$}
\end{equation}
\begin{equation}
f_is_j = s_jf_i \quad (|i -j|\geq 2),\tag{$R4$}
\end{equation}
\begin{equation}
e_{i}^2 = Qe_i ,\tag{$E1$}
\end{equation}
\begin{equation}
s_ie_{i}e_{i+1}
= e_ie_{i+1}s_i
= e_ie_{i+1}
\quad (i = 1, 2, \ldots, n-1),\tag{$E2$}
\end{equation}
\begin{equation}
e_is_j = s_je_i
\quad(j-i\geq1,\ i-j\geq 2),
\quad e_ie_j
= e_je_i,\tag{$E3$}
\end{equation}
\begin{equation}
\begin{array}{rcl}\tag{$E4$}
e_if_ie_i = e_i
& e_{i+1}f_{i}e_{i+1} = e_{i+1}
& (i = 1, 2, \ldots, n-1),\\
f_ie_{i}f_i = f_{i},
&f_{i}e_{i+1}f_{i} = f_{i}
&(i = 1, 2, \ldots, n-1),
\end{array}
\end{equation}
\begin{equation}
e_if_j = f_je_i
\quad(j-i\geq1,\ i-j\geq 2).\tag{$E5$}
\end{equation}
Here we make a remark on the special elements above.
\begin{rem}\label{rem:gen}
The relation $(R0)$ implies that the special elements
$\{f_i\}$
and $\{e_i\}$
are generated by $f=f_1$, $e = e_1$ and $s_1,\ldots, s_{n-1}$.
\end{rem}
In this note, firstly we show that
the special elements and the basic relations ($R0$)-($R4$)
and ($E1$)-($E5$) above characterize
the partition algebra $A_{n}(Q)$, {\em i.e.}
the special elements generate $A_{n}(Q)$,
and all the possible relations in $A_{n}(Q)$ are
obtained from the basic relations.
By Remark~\ref{rem:gen},
the basic relations will be translated into the relations
among the symbols $f$, $e$ and $s_i$s.
Characterizations will be stated by these symbols.
\subsection{Characterization for $A_{n}(Q)$}
Since generators $\{s_i\ | \ 1\leq i\leq n-1\}$
of the partition algebra $A_{n}(Q)$
satisfy the relations of the symmetric group ${\mathfrak S}_n$,
we can understand that $f_i$ and $e_i$ are ``conjugate'' to $f$ and $e$
respectively.
Hence the basic relations $(R2)$-$(R4)$ and $(E1)$-$(E5)$
among the special elements
are translated into the relations $(R2')$-$(R4')$
and $(E1')$-$(E5')$ among the generators
as follows.
\begin{thm}\label{th:main}
The partition algebra $A_{n}(Q)$ is
characterized by the generators
\begin{equation*}
f, e, s_1, s_2, \ldots, s_{n-1},
\end{equation*}
and the relations
\begin{equation}
\begin{array}{rcl}\tag{$R1$}
s_i^2 &=& 1 \quad (i = 1, 2, \ldots, n-1),\\
s_is_{i+1}s_i &=& s_{i+1}s_is_{i+1} \quad (i = 1, 2, \ldots, n-2),\\
s_is_j &=& s_js_i \quad (|i-j|\geq 2, \ i, j =1,2, \ldots, n-1),
\end{array}
\end{equation}
\begin{equation}
f^2 = f,\ fs_2fs_2 = s_2fs_2f,\ fs_2s_1s_3s_2fs_2s_1s_3s_2 = s_2s_1s_3s_2fs_2s_1s_3s_2f,\tag{$R2'$}
\end{equation}
\begin{equation}
fs_1 = s_1f=f,\tag{$R3'$}
\end{equation}
\begin{equation}
fs_i = s_if \quad (i = 3, 4, \ldots, n-1),\tag{$R4'$}
\end{equation}
\begin{equation}
e^2 = Qe,\tag{$E1'$}
\end{equation}
\begin{equation}
es_1es_1 = s_1es_1e = es_1e,\tag{$E2'$}
\end{equation}
\begin{equation}
es_i=s_ie \quad (i = 2, 3, \ldots, n-1),\tag{$E3'$}
\end{equation}
\begin{equation}
efe = e,\ fef = f,\tag{$E4'$}
\end{equation}
\begin{equation}
fs_2s_1es_1s_2
= s_2s_1es_1s_2f.\tag{$E5'$}\end{equation}
\end{thm}
In Sections~2-4
we prove this theorem not using
the generators and the relations in the theorem
but using the special elements and the basic relations $(R0)$-$(R4)$
and $(E1)$-$(E5)$.
The partition algebras $A_n(Q)$ were introduced in early 1990s
by Martin~\cite{Ma1,Ma2} and Jones~\cite{Jo} independently
and have been studied, for example, in the papers~\cite{Ma3,DW,HR}.
The theorem above has already shown in the paper~\cite{HR}.
Here we give another poof defining a ``standard'' expression of
a word of the special elements of $A_{n}(Q)$
according to the papers~\cite{Ko1,Ko3,Ko4}.
From this standard expression,
we will find that the partition algebra $A_{n}(Q)$
is cellular in the sense of Graham and Lehrer~\cite{GL}.
Thus, applying the general representation of cellular
algebras to the partition algebras,
we will get a description of the irreducible
modules of $A_{n}(Q)$ for any field of arbitrary characteristic.
(For the cell representations, we also refer the paper~\cite{KL}.)
Further, we can make the character table of $A_{n}(Q)$
using the standard expressions.
These topics will be studied in near future.
For the present we refer the notes~\cite{Ko3,Na1}
and the results about the partition algebras~\cite{DW,Xi}.
\section{Local moves deduced from the basic relations}
Let
\[
{\cal L}_n^1 =
\{
s_1, s_2, \ldots, s_{n-1},
f_1, f_2, \ldots, f_{n-1},
e_1, e_2, \ldots, e_{n}
\}
\]
be the set of symbols
whose words satisfy the basic relations $(R0)$-$(R4)$
and $(E1)$-$(E5)$.
There are many relations among these symbols
which are deduced from the basic relations.
These relations are pictorially
expressed as local moves.
Among them, we frequently use relations
$f_{i+1}s_is_{i+1}=s_is_{i+1}f_i$ ($R0$),
$f_is_{i+1}f_i = f_if_{i+1}$ ($R2''$)
and
$e_is_i = e_if_ie_{i+1} =s_ie_{i+1}$ ($E4''$)
as in Figure~\ref{fig:fss},\ref{fig:fsf} and \ref{fig:efe}
respectively.
The latter two relations are deduced from the relations ($R0$)-($R3$)
and ($R0$), ($R3$), ($E4$) respectively.
\fig{fig:fss}{$f_{i+1}s_is_{i+1}=s_is_{i+1}f_i$ ($R0$)}{4.eps}
\fig{fig:fsf}{$f_is_{i+1}f_i = f_if_{i+1}$ ($R2''$)}{5.eps}
\fig{fig:efe}{$e_is_{i}=e_if_ie_{i+1} = s_ie_{i+1}$ ($E4''$)}{6.eps}
As we noted in the previous paper~\cite{Ko4},
these basic relations
are invariant under the transpositions of indices
$i\leftrightarrow n-i+1$ as well as
the $\mathbb{Z}[Q]$-linear involution $*$ defined by
$(xy)^{*} = y^*x^*$ ($x, y\in A_{n}(Q)$).
This implies that if one local move is allowed
then other three moves ---obtained by
reflections with respect to the vertical and the horizontal lines
and their composition---
are also allowed.
Further, we note that if we put
\[
e^{[r]} = f_1f_2\cdots f_{r-1}e_1e_2\cdots e_rf_1f_2\cdots f_{r-1}
\]
then we can check that $e^{[r]}$, $f$ and $s_i$ ($1\leq i\leq n-1$)
satisfy the defining relations of $P_{n,r}(Q)$,
the $r$-modular party algebra, defined in the paper~\cite{Ko4}.
This means that the local moves shown in the paper~\cite{Ko4}
also hold in $A_n(Q)$
(in fact, these local moves are more easily verified
in $A_n(Q)$).
Some of them are pictorially expressed
in Figures~\ref{fig:dpjrE},\ref{fig:roeE},\ref{fig:dpsE}
and \ref{fig:dpeE}.
\fig{fig:dpjrE}{Defective part jump rope\ ($R13'$)}{7.eps}
\fig{fig:roeE}{Removal (addition) of excrescences\ ($R14'$)}{8.eps}
\fig{fig:dpsE}{Defective part shift\ ($R16'$)}{9.eps}
\fig{fig:dpeE}{Defective part exchange\ ($R17'$)}{10.eps}
\section{Standard expressions of seat-plans}
In this section, for a seat-plan $w$ of $\Sigma_n^1$,
we define a {\em basic expression},
as a word in the alphabet $\Gamma_n^1$.
Then we define more general forms
called {\em crank form expression}s.
As a special type of the crank form expression,
we define the {\em standard expression}.
In the next section,
we show that any
two crank form expressions of a seat-plan
will be moved to each other by using the basic relations
$(R0)$-$(R4)$ and $(E1)$-$(E5)$
finite times.
Consequently, we find that
any seat-plan can be moved to its standard expression.
To define these expressions,
we introduce some terminologies.
\subsection{Propagating number}
Let $w = \{T_1, T_2, \ldots, T_s\}$
be a seat-plan of $A_{n}(Q)$.
For a part $T_i\in w$, the intersection with $M$, or $T_i^M = M\cap T_i$,
is called the {\em upper part} of $T_i$.
Similarly, $T_i^F = F\cap T_i$
is called the {\em lower part} of $T_i$.
If $T_i^M\neq\emptyset$ and $T_i^F\neq\emptyset$ hold
simultaneously,
$T_i$ is called {\em propagating},
otherwise, it is called {\em non-propagating},
or {\em defective}.
Let
$\pi(w) := \{T\in w\ |\ T:\mbox{propagating}\}$
be the set of propagating parts.
The number of propagating parts $|\pi(w)|$ is
called the {\em propagating number} (of $w$).
If $T_i\in\pi(w)$ then the upper [resp. lower]
part $T^M_i$ [resp. $T^F_i$] of $T_i$ is also called {\em propagating}.
If $T_i\in w\setminus\pi(w)$ and $T^M_i = T_i$ [resp. $T^F_i = T_i$],
then $T^M_i$ [resp. $T^F_i$] is called {\em defective}.
For example, in Figure~\ref{fig:plan},
$\pi(w) = \{T_1, T_4\}$.
Hence $|\pi(w)| = 2$.
On the other hand $T_2$, $T_3$ and $T_5$ are defective.
The upper and the lower propagating parts are $\{1\}$, $\{5\}$
and $\{1',2',4'\}$, $\{3'\}$ respectively.
The upper defective parts are $T_2$ and $T_3$.
The lower defective part is $T_5$.
\subsection{A basic expression of a seat-plan}
For a part $T_i\in w$, define $t(T_i)$ by
\[
t(T_i) =
\left\{
\begin{array}{ll}
1&\mbox{if $T_i$ is propagating,}\\
0&\mbox{if $T_i$ is defective.}
\end{array}
\right.
\]
Similarly we define $t(T_i^M)$ [resp. $t(T_i^F)$]
to be $1$ or $0$ in accordance with that $T_i^M$ [resp. $T_i^F$] is propagating
or not.
Using the terminologies above,
first we define a {\em basic expression} of an seat-plan.
Let $w\in\Sigma_n^1$ be a seat-plan
and $\rho_w = (T_1, \ldots, T_s)$ be an arbitrary sequence of all parts of $w$.
For the sequence $\rho_w$,
we define the sequence of the upper [resp. lower]
parts $\mathbb{M} = \mathbb{M}(\rho_w) = (T^M_{i_1}, \ldots, T^M_{i_u})$
($i_1<\cdots<i_u$, $u\leq s$)
[resp. $\mathbb{F} = \mathbb{F}(\rho_w) = (T^F_{j_1}, \ldots, T^F_{j_v})$
($j_1<\cdots<j_v$, $v\leq s$)] omitting empty parts.
Using these data, we define {\em cranks}
$C_{\mathbb{M}}[i]$, $C^*_{\mathbb{F}}[i]$
and $C^{\mathbb{M}}_{\mathbb{F}}[\sigma])$
as products of the generators as in
Figure~\ref{fig:mcrank},
\ref{fig:fcrank}
and \ref{fig:midcrank}
respectively.
Here $\sigma$ is a word in the alphabet
$\{s_1,\ldots, s_{|\pi(w)|-1}\}$.
\fig{fig:mcrank}{$C_{\mathbb{M}}[l]$}{11.eps}
\fig{fig:fcrank}{$C^*_{\mathbb{F}}[l]$}{12.eps}
\fig{fig:midcrank}{$C^{\mathbb{M}}_{\mathbb{F}}[\sigma]$}{13.eps}
Further we define the ``product of cranks''
${C}[\mathbb{M}]$ and ${C}[\mathbb{F}]$
by
\[
{C}[\mathbb{M}] =
C_{\mathbb{M}}[1]C_{\mathbb{M}}[2]\cdots C_{\mathbb{M}}[u-1]
\]
and
\[
{C}^*[\mathbb{F}] =
C^*_{\mathbb{F}}[v-1]\cdots C^*_{\mathbb{F}}[2]C^*_{\mathbb{F}}[1]
\]
respectively.
We note that $C_{\mathbb{M}}[l]$ [resp. $C^*_{\mathbb{F}}[l]$]
is defined by a composition $\mathbb{E} = (E_1,\ldots, E_s)$ of $n$
whose components have labels either ``propagating'' or ``defective''.
For example if $\mathbb{M} = (2,1,2,2,3)$,
$(t(M_i))_{1\leq i \leq 5} = (0,1,0,1,1)$,
$\mathbb{F} = (3,4,3)$, $(t(F_i))_{i=1,2,3} = (1,1,1)$
and $\sigma=(1,2)(2,3)\in\mathfrak{S}_3$,
then the product of cranks
${C}[\mathbb{M}]C^{\mathbb{M}}_{\mathbb{F}}[\sigma]C^*[\mathbb{F}]$
is presented as in Figure~\ref{fig:crank}.
\fig{fig:crank}{Product of cranks}{14.eps}
Let $\overline{\mathbb M}$
be the sequence of $n$ symbols obtained from
$\mathbb{M} = \mathbb{M}(\rho_w)$ by
arranging all elements of $T^M_{i_k}$s
in accordance with the sequence $\mathbb{M}$
so that all elements of each $T^M_{i_k}$ are increasingly lined up
from left to right.
For example, if $\mathbb{M} = (\{3,1,7\},\{6,4\},\{5,2\})$,
then $\overline{\mathbb{M}} = (1,3,7,4,6,2,5)$.
Similarly $\overline{\mathbb{F}}$ is defined
from $\mathbb{F} = \mathbb{F}(\rho_w)$.
Then the following product becomes
an expression of a seat-plan $w$.
\[
{\cal C}(\mathbb{M}, id, \mathbb{F})
= x_{\overline{\mathbb{M}}}{C}[\mathbb{M}]
C^{\mathbb{M}}_{\mathbb{F}}[id]
C^*[\mathbb{F}]x^*_{\overline{\mathbb{F}}}.
\]
Here $x_{\overline{\mathbb{M}}}$ [resp. $x^*_{\overline{\mathbb{F}}}$]
is a permutation
which maps $j$ to the number in the $j$-th coordinate
of $\overline{\mathbb{M}}$.
[resp. the number written in the $j$-th coordinate of
$\overline{\mathbb{F}}$ to $j'$].
We call this expression a {\em basic expression} of $w$.
We note that for a seat-plan $w$
there are several ways to choose $\rho_w$, a sequence of the parts of $w$.
{\em i.e.}
Several basic expressions can be defined for one seat-plan.
\subsection{The standard expressions of a seat-plan}
Our claim is that
any basic expression of a seat-plan $w$
can be moved to a special expression
called the {\em standard expression}
by using the relations $(R0)$-$(R4)$ and $(E1)$-$(E5)$ finite times.
In order to show this claim,
next we introduce the notion of a {\em crank form expression} of $w$.
Consider the propagating parts
$\pi(w)=\{T_{i_1},\ldots, T_{i_p}\}\ (p = |\pi(w)|)$ of $w$.
Let $(M_1, \ldots, M_p)$ be
a sequence of the upper parts of $\pi(w)$
and $(F_1, \ldots, F_p)$
the one of the lower parts.
Then there exists a permutation $\sigma\in{\mathfrak S_p}$
such that $\{M_{\sigma(k)}\sqcup F_{k}\ |\ k = 1, \ldots, p \} = \pi(w)$.
As is well known, a permutation $\sigma$ of degree $p$ is presented by
$p$-strings braid which connects the lower $k$-th point
to the upper $\sigma(k)$-th point.
Now we define a {\em crank form expression} of $w$.
Let $\mathbb{M} = (M_{1}, \ldots, M_{u})$
[resp. $\mathbb{F} = (F_1, \ldots, F_v)$]
be any fixed sequence of the upper [resp. lower] parts of $w$
(whose empty parts are omitted and propagating parts are specified).
From the sequences $\mathbb{M}$ and $\mathbb{F}$,
we obtain products of cranks $C[\mathbb{M}]$ and $C^*[\mathbb{F}]$.
Further, from $\pi(\mathbb{M})$ and $\pi(\mathbb{F})$,
we obtain a permutation $\sigma\in{\mathfrak S}_p$
such that $\{M_{i_{\sigma(k)}}\sqcup F_{j_{k}}\ ;\ k = 1, \ldots, p \}
= \pi(w)$.
Then
the product
\[
{\cal C}(\mathbb{M},\sigma,\mathbb{F})
=
x_{\mathbb{\overline{M}}}
{C}[\mathbb{M}]C^{\mathbb{M}}_{\mathbb{F}}[\sigma]C^*[\mathbb{F}]
x^*_{\overline{F}}
\]
becomes a presentation of $w$.
We call this presentation a {\em crank form expression of $w$
defined by ${\mathbb M}$ and ${\mathbb F}$}.
If a crank form expression is
made from sequences
$(M_1, \ldots, M_u)$ and $(F_1, \ldots, F_v)$ such that
\begin{enumerate}
\item
$M_1, \ldots, M_p$ and $F_1, \ldots, F_p$ are propagating,
\item
$M_{p+1}, \ldots, M_{u}$
and $F_{p+1}, \ldots, F_{v}$ are defective.
\end{enumerate}
then we call it {\em in normal form}.
Finally, we define the {\em standard expression} of $w$,
as a special expression of crank form expressions in normal form
by properly choosing the sequences $(M_1, \ldots, M_u)$
and $(F_1, \ldots, F_v)$.
For this purpose
first we sort the parts $T_1, \ldots, T_s$ of $w$ so that they satisfy:
\begin{enumerate}
\item
$\pi(w) = \{T_1, T_2, \ldots, T_p\},$
\item
$\{T_i\ |\ i = p+1, p+2, \ldots, u\}$
is the set of all upper defective parts,
\item
$\{T_i\ |\ i = u+1, u+2, \ldots, u+(v-p)\}$
is the set of all lower defective parts.
\end{enumerate}
For an ordered set $E$, let $\min E$ be the minimum element in $E$.
Let $T_1, T_{2}, \ldots, T_{p}$ be the parts of $\pi(w)$.
Define $(M_1, M_{2}, \ldots, M_{p})$
so that they satisfy
\[
\{M_1, M_{2}, \ldots, M_{p}\}
= \{T_1^M, T_{2}^M, \ldots, T_{p}^M\}
\]
and
\[
\min M_1 <\min M_{2} <\cdots <\min M_{p}.
\]
Similarly $(F_1, F_{2}, \ldots, F_{p})$
are defined using the lower parts of $\pi(w)$.
In such a method, the sequences of the upper parts $(M_1,\ldots, M_p)$ and
the lower parts $(F_1, \ldots, F_p)$ are uniquely defined
from a seat-plan $w$.
Now we define $(M_{p+1}, \ldots, M_{u})$
so that they satisfy
\[
\{M_{p+1}, M_{p+2}, \ldots, M_{u}\}
= \{T_{p+1}, T_{p+2}, \ldots, T_{u}\}
\]
and
\[
\min M_{p+1} <\min M_{p+2} <\cdots <\min M_{u}.
\]
Similarly we define $(F_{p+1}, \ldots, F_{v})$
so that they satisfy
\[
\{F_{p+1}, F_{p+2}, \ldots F_{v}\}
= \{T_{u+1}, T_{u+2}, \ldots, T_{u+(v-p)}\}
\]
and
\[
\min F_{p+1} <\min F_{p+2} <\cdots <\min F_{v}.
\]
Using these upper and lower sequences defined above,
we can obtain a crank from expression in normal form
called the {\em standard expression} of $w$.
\section{Proof of Theorem~1.2}
In the previous section,
we have defined the standard expression
of a word in the alphabet ${\cal L}_n^1$
as a special expression of the crank form expressions in normal form.
In this section, first we show that
any two crank form expressions of a seat-plan $w$
are transformed to each other by finitely using the local moves
shown in Section~2.
Then we show
that any word in the alphabet ${\cal L}_n^1$
is moved to a scalar multiple of one of the crank form expressions.
Thus we can find that
any word in the alphabet ${\cal L}_n^1$
is reduced to a scalar multiple of a the standard expression.
Since the set of seat-plans
makes a basis of $A_{n}(Q)$
and since every seat-plan has its standard expression,
this proves that the partition algebra $A_{n}(Q)$ is characterized
by the generators and relations in Theorem~\ref{th:main}.
First we show that
any two crank form expressions
are transformed to each other.
For $w\in\Sigma_n^1$, let ${\mathbb M} = (M_1, \ldots, M_u)$
and ${\mathbb F} = (F_1, \ldots, F_v)$ be
sequences of the upper and the lower parts of $w$ respectively.
Assume that the subsequence $\pi(\mathbb{M}) = (M_{i_1}, \ldots, M_{i_p})$
($i_1<\cdots <i_p$) of $\mathbb{M}$
is the sequence of the upper propagating parts
and $\pi(\mathbb{F}) = (F_{j_1}, \ldots, F_{j_p})$
($j_1<\cdots <j_p$) is that of the lower propagating parts.
Then
there exists a permutation $\sigma$ of degree $p = |\pi(w)|$
which specifies how the propagating parts of $w$
are recovered from $\pi(\mathbb{M})$ and $\pi(\mathbb{F})$.
Let $\mathbb{E} = (E_1, \ldots, E_s)$
be a sequence of the upper or lower parts.
Suppose that $\tau\in{\mathfrak S}_s$ acts on $\mathbb{E}$
by $\tau\mathbb{E} = (E_{\tau^{-1}(1)}, \ldots, E_{\tau^{-1}(s)})$.
Then the following lemma holds.
\begin{lem}\label{lem:defect}
Let ${\mathbb M} = (M_1, \ldots, M_u)$
and ${\mathbb F} = (F_1, \ldots, F_v)$ be
sequences of the upper and the lower (non-empty) parts of
a seat-plan respectively.
If $M_i$ [resp. $F_i$] is defective
and $\sigma_i = (i,i+1)$, the $i$-th adjacent transposition,
then the crank form expression
${\cal C}(\mathbb{M}, \sigma, \mathbb{F})$ is moved to
another crank form expression
${\cal C}(\sigma_i\mathbb{M}, \sigma, \mathbb{F})$
[resp.
${\cal C}(\mathbb{M}, \sigma, \sigma_i\mathbb{F})$
].
\end{lem}
\begin{proof}
We consider the case $M_i$ is defective.
In case $F_i$ is defective, the similar proof will hold.
Let $P_{\mathbb{M},i}\in{\mathfrak S}_n$ be
a permutation defined by
\[
P_{\mathbb{M},i}(x):=
\left\{
\begin{array}{ll}
x + |M_{i+1}|&
\mbox{if}\ \sum_{j=1}^{i-1}|M_{j}| <x \leq \sum_{j=1}^{i}|M_{j}|,\\
x - |M_{i}|&
\mbox{if}\ \sum_{j=1}^{i}|M_{j}| <x \leq \sum_{j=1}^{i+1}|M_{j}|,\\
x&
\mbox{otherwise}.
\end{array}
\right.
\]
Then we find that
$x_{\overline{\mathbb{M}}}P^{-1}_{\mathbb{M},i}$
maps $j$ to the $j$-th coordinate of $\overline{\sigma_i\mathbb{M}}$.
Hence we have $x_{\overline{\mathbb{M}}}P^{-1}_{\mathbb{M},i}
= x_{\overline{\sigma_i\mathbb{M}}}$.
(For the definition of $\overline{\mathbb{M}}$, see Section~3.2.)
On the other hand, since $M_i$ is defective,
we have $P_{\mathbb{M},i}C[\mathbb{M}] = C[\sigma_i\mathbb{M}]$
by removing an excrescence of $M_i$
and iteratively using ``defective part exchange'' ($R17'$)
in Figure~\ref{fig:dpeE} (if $M_{i+1}$ is defective)
or iteratively using ``defective part shift'' ($R16'$)
in Figure~\ref{fig:dpsE} (if $M_{i+1}$ is propagating),
and then adding an excrescence to $M_i$ just moved.
Thus we obtain
\begin{eqnarray*}
{\cal C}(\mathbb{M}, \sigma, \mathbb{F})
&=& x_{\overline{\mathbb{M}}}
C[\mathbb{M}]C^{\mathbb{M}}_{\mathbb{F}}[\sigma]C^*[\mathbb{F}]
x^*_{\overline{\mathbb{F}}}\\
&=& (x_{\overline{\mathbb{M}}}P^{-1}_{\mathbb{M},i})
(P_{\mathbb{M},i}C[\mathbb{M}])
C^{\mathbb{M}}_{\mathbb{F}}[\sigma]C^*[\mathbb{F}]
x^*_{\overline{\mathbb{F}}}\\
&=& x_{\overline{\sigma_i\mathbb{M}}}
C[\sigma_i\mathbb{M}]
C^{\mathbb{M}}_{\mathbb{F}}[\sigma]C^*[\mathbb{F}]
x^*_{\overline{\mathbb{F}}}\\
&=&{\cal C}(\sigma_i\mathbb{M}, \sigma, \mathbb{F}).
\end{eqnarray*}
\end{proof}
\begin{rem}\label{rem:defect2}
Lemma~\ref{lem:defect} also holds
if $M_{i+1}$ [resp. $F_{i+1}$] is defective.
\end{rem}
By Lemma~\ref{lem:defect} and Remark~\ref{rem:defect2}
we may assume that any crank form expression
is given in normal form.
\begin{lem}\label{lem:crex}
Let ${\cal C}(\mathbb{M}, \sigma, \mathbb{F})$
be a crank form expression of $w$ in normal form.
If $M_i$ and $M_{i+1}$ are propagating
then ${\cal C}(\mathbb{M},\sigma,\mathbb{F})$
is moved to another crank form expression
${\cal C}(\sigma_i\mathbb{M}, \sigma_i\sigma,\mathbb{F})$ in normal form.
Similarly if $F_i$ and $F_{i+1}$ are propagating,
then ${\cal C}(\mathbb{M},\sigma,\mathbb{F})$
is moved to ${\cal C}(\mathbb{M},\sigma\sigma_i,\sigma_i\mathbb{F})$.
\end{lem}
\begin{proof}
Let $C_{\mathbb{M}}[i]$ and $C_{\mathbb{M}}[i+1]$
be $i$-th and $(i+1)$-st cranks of $C[\mathbb{M}]$.
By Figure~\ref{fig:crex}, we have
\[
P_{\mathbb{M},i}C_{\mathbb{M}}[i]C_{\mathbb{M}}[i+1]
=C_{\sigma_i\mathbb{M}}[i]C_{\sigma_i\mathbb{M}}[i+1]\sigma_i.
\]
\fig{fig:crex}{Crank form exchange}{15.eps}
Thus we obtain
\begin{eqnarray*}
{\cal C}(\mathbb{M}, \sigma, \mathbb{F})
&=&
x_{\overline{\mathbb{M}}}
C[\mathbb{M}]
C^{\mathbb{M}}_{\mathbb{F}}[\sigma]C^*[\mathbb{F}]
x^*_{\overline{\mathbb{F}}}\\
&=&
(x_{\overline{\mathbb{M}}}P^{-1}_{\mathbb{M},i})
(P_{\mathbb{M},i}C[\mathbb{M}])
C^{\mathbb{M}}_{\mathbb{F}}[\sigma]C^*[\mathbb{F}]
y^*_{\overline{\mathbb{F}}}\\
&=&
x_{\overline{\sigma_i\mathbb{M}}}
C[\sigma_i\mathbb{M}]C^{\mathbb{M}}_{\mathbb{F}}[\sigma_i\sigma]
C^*[\mathbb{F}]
x^*_{\overline{\mathbb{F}}}\\
&=& {\cal C}(\sigma_i\mathbb{M}, \sigma_i\sigma, \mathbb{F}).
\end{eqnarray*}
\end{proof}
By Lemma~\ref{lem:defect}, Remark~\ref{rem:defect2} and Lemma~\ref{lem:crex}
we obtain the following.
\begin{prop}\label{prop:normalize}
A crank form expression of a seat-plan
is moved to its standard expression.
\end{prop}
Now we prove that any word in the alphabet ${\cal L}_n^1$
is moved to a crank form expression.
By the above proposition, we will find that
any word can be moved to its standard expression.
\begin{prop}
If ${\cal C}(\mathbb{M},\sigma,\mathbb{F})$ is the standard
expression of a seat-plan $w$,
then $s_i{\cal C}(\mathbb{M},\sigma,\mathbb{F})$ is
moved to a crank form expression of $s_iw$.
\end{prop}
\begin{proof}
If $i$ and $i+1$ are both included one of the (upper) parts of $w$,
say $M_k$, then we have
\[
\sum_{j=1}^{k-1}|M_j|
<x^{-1}_{\overline{\mathbb{M}}}(i) < x^{-1}_{\overline{\mathbb{M}}}(i+1)
= x^{-1}_{\overline{\mathbb{M}}}(i)+1
\leq\sum_{j=1}^{k}|M_j|
\]
and
\[
(x^{-1}_{\overline{\mathbb{M}}}(i), x^{-1}_{\overline{\mathbb{M}}}(i+1))
C_{\mathbb{M}}[k]
= (x^{-1}_{\overline{\mathbb{M}}}(i), x^{-1}_{\overline{\mathbb{M}}}(i)+1)
C_{\mathbb{M}}[k]
= {\cal C}_{\mathbb{M}}[k].
\]
Since
\[
s_ix_{\overline{\mathbb{M}}} = (i,i+1)x_{\overline{\mathbb{M}}}
= x_{\overline{\mathbb{M}}}
(x^{-1}_{\overline{\mathbb{M}}}(i), x^{-1}_{\overline{\mathbb{M}}}(i+1)),
\]
we find that $s_i{\cal C}(\mathbb{M}, \sigma, \mathbb{F})
= {\cal C}(\mathbb{M}, \sigma, \mathbb{F})$ is a crank form expression.
If $i$ is included in $M_j$ and $i+1$ is included in $M_k$ ($j\neq k$),
then we have
$s_ix_{\overline{\mathbb{M}}} = x_{\overline{\mathbb{M}'}}$.
Here $\mathbb{M}'$ is the sequence of the upper parts
obtained from $\mathbb{M}=(M_1,\ldots, M_u)$
by replacing $M_j$ with $M_j' = (M_j\setminus\{i\})\cup\{i+1\}$
and $M_k$ with $M_k' = (M_k\setminus\{i+1\})\cup\{i\}$.
Hence we find that $s_i{\cal C}(\mathbb{M},\sigma,\mathbb{F})$
is moved to ${\cal C}(\mathbb{M}',\sigma,\mathbb{F})$,
a crank form expression.
In particular this expression again becomes
the standard expression, unless $k = j+1$,
$t(M_j) = t(M_{j+1})$, and $i = \min M_j$, $i+1 = \min M_{j+1}$.
\end{proof}
\begin{prop}
If ${\cal C}(\mathbb{M},\sigma,\mathbb{F})$ is the standard
expression of a seat-plan $w$,
then $f{\cal C}(\mathbb{M},\sigma,\mathbb{F})$ is
moved to a crank form expression of $fw$.
\end{prop}
\begin{proof}
First consider the case $\{1, 2\}\subset M_k$ for some $k$.
In this case, there exists an integer $i$
such that $i = x^{-1}_{\overline{\mathbb{M}}}(1)$
and $i+1 = x^{-1}_{\overline{\mathbb{M}}}(2)$.
Hence in this case we have
$fx_{\overline{\mathbb{M}}} = x_{\overline{\mathbb{M}}}f_i$
and
$f_i{\cal C}_{\mathbb{M}}[k] = {\cal C}_{\mathbb{M}}[k]$.
Thus we obtain $f{\cal C}(\mathbb{M}, \sigma, \mathbb{F})
= {\cal C}(\mathbb{M}, \sigma, \mathbb{F})$.
Next consider the case $1\in M_j$ and $2\in M_k$ ($j\neq k$).
In the following we assume that $M_j$ and $M_k$ are both propagating.
Even if either $M_j$ or $M_k$ or both of them are defective,
the similar proof will hold.
Proposition~\ref{prop:normalize} implies that
the standard expression ${\cal C}(\mathbb{M}, \sigma, \mathbb{F})$
is moved to a crank form expression
${\cal C}(\mathbb{M}', id, \mathbb{F}')$
so that the first
and the second components of $\mathbb{M}'$
are $M_j$ and $M_k$ respectively
and the first and the second components of $\mathbb{F}'$
are jointed to $M_j$ and $M_k$ respectively.
Using the relations ($R2''$), ($R2$)
and ($R12''$),
we find that the first and the second components of
$\mathbb{M}'$ and those of $\mathbb{F}'$
are merged by the action of $f$.
For example, if $|M_j| = 5$ and $|M_k| = 4$
then we have Figure~\ref{fig:fwE}.
\fig{fig:fwE}{Action of $f$ on $w$}{16.eps}
The merged propagating parts will be moved to a
crank form expression ${\cal C}(\mathbb{M}'', id, \mathbb{F}'')$
by ``bumping'' as in Figure~\ref{fig:bump}.
Here $\mathbb{M}''$ [resp. $\mathbb{F}''$]
is a sequence of upper [resp. lower] parts
obtained from $\mathbb{M}$ [resp. $\mathbb{F}$]
by merging the first two components.
\fig{fig:bump}{Bumping}{17.eps}
\end{proof}
\begin{prop}
If ${\cal C}(\mathbb{M},\sigma,\mathbb{F})$ is the standard
expression of a seat-plan $w$,
then $e{\cal C}(\mathbb{M},\sigma,\mathbb{F})$ is
moved to a crank form expression of $ew$.
\end{prop}
\begin{proof}
By the same argument in the previous proposition,
we may assume that ${\cal C}(\mathbb{M},\sigma,\mathbb{F})$
is moved to a crank form expression
\[
{\cal C}(\mathbb{M}'', id, \mathbb{F}'')
\]
such that the first component $M_1''$ of $\mathbb{M}''$
contains $\{1\}$.
First consider the case $|M_1''|>1$.
In this case, it is easy to check that
$e{\cal C}(\mathbb{M}'', id, \mathbb{F}'')$
is again a crank form expression
of $ew$ as it is.
Next consider the case $|M_1''| = 1$.
If $M_1''$ is defective,
then we have a scalar multiple of a crank form expression
$e{\cal C}(\mathbb{M}'', id, \mathbb{F}'') =
Q{\cal C}(\mathbb{M}'', id, \mathbb{F}'')$.
If $M_1''$ is propagating,
then applying ``addition of excrescences ($R14'$)''
and ``bumping''
in Figures~\ref{fig:roeE} and \ref{fig:bump}
we can move $e{\cal C}(\mathbb{M}'', id, \mathbb{F}'')$ to a crank form
expression.
\end{proof}
\begin{proof}[Proof of Theorem~1.2]
Let $\widetilde{A_{n}(Q)}$
be the associative algebra over $\mathbb{Z}[Q]$
abstractly defined by the generators and the relations
in Theorem~1.2.
So there exists a surjective morphism $\psi$
from $\widetilde{A_{n}(Q)}$ to $A_{n}(Q)$.
As we noted in Section~1,
we may assume that $\widetilde{A_{n}(Q)}$
is generated by the alphabets ${\cal L}_n^1$
which satisfy the relations ($R0$)-($R4$) and ($E1$)-($E5$).
Here we note that the ``geometrical moves''
we have shown previously
can be applied to any algebra which satisfies
the relations ($R0$)-($R4$) and ($E1$)-($E5$).
Hence if we associate the alphabets in ${\cal L}_n^1$
with the diagrams in Figure~\ref{fig:gen},
then we can apply the notion of {\em basic expressions},
{\em crank form expressions}
and {\em standard expressions} to the words in the alphabets
${\cal L}_n^1$ of $\widetilde{A_{n}(Q)}$.
Let $w$ be a word in the alphabet ${\cal L}_n^1$ of $\widetilde{A_{n}(Q)}$.
Suppose that $w$ is presented in a standard expression.
Then by Proposition~4.5-4.7,
$s_iw$, $fw$ and $ew$ are all moved to
(possibly scalar multiples of) crank form expressions.
By Proposition~4.4, they are moved to the standard expressions.
Since $s_i$ ($1\leq i\leq n-1$), $f$, and $e$ are
crank form expressions as they are,
by induction on the lengths of the words in the alphabets ${\cal L}_n$,
any word turn out to be equal to (a scalar multiple) of
the standard expression of a seat-plan $w$ of $\Sigma_n^1$.
Hence we have
\[
\mbox{rank}\ \widetilde{A_{n}(Q)} \leq |\Sigma_n^1|.
\]
As Tanabe showed in \cite{Ta},
$\Sigma_n^1$ makes a basis of
$\mbox{${\mathbb C}$}\otimes A_{n}(k) = \mbox{${\mathbb C}$}\otimes \psi(\widetilde{A_{n}(k)})$
if $k\geq n$.
Hence $\mbox{rank}\ \mbox{${\mathbb C}$}\otimes A_{n}(z) = |\Sigma_n^1|$
holds as far as $z$ takes any integer value $k\geq n$.
This implies that $\psi$ is an isomorphism
and we find that the generators and the relations
in Theorem~1.2 characterize the partition algebra $A_{n}(Q)$.
\end{proof}
\section{Definition of $A_{n-\frac{1}{2}}(Q)$, a subalgebra of $A_n(Q)$}
\label{sec:5-1}
In this section,
we consider a subalgebra $A_{n-\frac{1}{2}}(Q)$ of $A_n(Q)$
generated by the special elements
$s_1, \ldots, s_{n-2}$, $f_1, \ldots, f_{n-1}$
and $e_1,\ldots, e_{n-1}$.
As we have noted in Remark~\ref{rem:gen},
$\{f_i\}$ ($1\leq n-2$)
and $\{e_i\}$ ($1\leq n-1$)
are written as products of $f=f_1$, $e = e_1$ and $s_1,\ldots, s_{n-2}$.
The special element $f_{n-1}$, however,
can not be expressed
as a product of other special elements in $A_{n-\frac{1}{2}}(Q)$,
since we deleted $s_{n-1}$ from the generators of $A_n(Q)$.
Hence $A_{n-\frac{1}{2}}(Q)$ can be defined
as a subalgebra of $A_n(Q)$
generated by the following elements:
$s_1, \ldots, s_{n-2}$, $f = f_1$, $f_* = f_{n-1}$
and $e = e_1$.
We can obtain the defining relations
among these generators just as in the case of $A_n(Q)$.
\begin{thm}\label{def:half-int-alg}
Let $\mathbb{Z}$
be the ring of rational integers
and $Q$ the indeterminate.
We put ${A}_{\frac{1}{2}}(Q) = \mathbb{Z}[Q]\cdot 1$.
For an integer $n\geq2$,
${A}_{n-\frac{1}{2}}(Q)$ is characterized by the
generators
\[
e, f, s_1, s_2, \ldots, s_{n-2}, f_{*} \mbox{(if $n>2$)}
\]
and the relations
($R0$), ($R1'$)-($R4'$) and ($E1'$)-($E5'$) omitting
the ones which involve $s_{n-1}$ and adding the
following relations:
\begin{gather}
f_{*}s_{n-2}s_{n-3}\cdots s_3s_2s_1
s_2s_3\cdots s_{n-3}s_{n-2}f_{*} \nonumber\\
\quad =\, f_{*}s_{n-2}s_{n-3}\cdots s_3s_2f
s_2s_3\cdots s_{n-3}s_{n-2}\phantom{,} \tag{$R2^*$} \\
\quad =\, s_{n-2}s_{n-3}\cdots s_3s_2f
s_2s_3\cdots s_{n-3}s_{n-2}f_{*},\nonumber
\end{gather}
\begin{equation}\tag{$R4^*$}
ff_{*} = f_{*}f,\quad
ef_{*} = f_{*}e,\quad
f_{*}s_i = s_if_{*}\ \mbox{($1\leq i \leq n-3$)},
\end{equation}
\begin{equation}\tag{$E4^*$}
\begin{array}{rcl}
f_{*}s_{n-2}s_{n-3}\cdots s_1es_1\cdots s_{n-3}s_{n-2}f_{*}
&=& f_{*},\\
es_{1}s_{2}\cdots s_{n-2}f_{*}s_{n-2}\cdots s_{2}s_{1}e
&=& e.
\end{array}
\end{equation}
We understand $A_{1+\frac{1}{2}}(Q) = A_{2-\frac{1}{2}}(Q)$
is defined by the generators $1$, $e$ and $f$ with the relations
$e^2 = Qe$, $f^2 = f$, $efe =e$, $fef = f$.
(Hence, $A_{2-\frac{1}{2}}(Q)$ is a rank 5 module with a basis
$\{1, e, f, ef, fe\}$.)
\end{thm}
The relations ($R2^*$) correspond to the relations
$f_{n-1}s_{n-2}f_{n-1} = f_{n-1}f_{n-2} = f_{n-2}f_{n-1}$.
We deduce $f_{*}s_{n-2}f_{*} = f_{*}s_{n-2}f_{*}s_{n-2}
= f_{*}s_{n-2}f_{*}s_{n-2}$ from ($R2^*$).
\begin{proof}
First we note that all the generators of $A_{n-\frac{1}{2}}(Q)$
have the part which contains $n$ and $n'$ simultaneously.
We consider the transpositions of indices
$i\leftrightarrow n-i+1$.
These transpositions make
$A_{n-\frac{1}{2}}(Q)$ a subalgebra of $A_{n}(Q)$
generated by
\[
{\cal L}_{n-\frac{1}{2}}^1 = \{f_1, \ldots, f_{n-1},
s_2, \ldots, s_{n-1}, e_2, \ldots, e_n\}.
\]
By the relation ($R0$), $A_{n-\frac{1}{2}}(Q)$ is actually
generated by letters
$\{f_1$, $f_2$, $e_2$ and $s_2,\ldots, s_{n-1}\}$.
Each of these generators has a part which includes $\{1, 1'\}$.
In the following in this section, we suppose that $A_{n-\frac{1}{2}}(Q)$
is generated by the letters in ${\cal L}^1_{n-\frac{1}{2}}$.
The $\mathbb{Z}[Q]$ bases of $A_{n-\frac{1}{2}}(Q)$
consist of $\Sigma^1_{n-\frac{1}{2}}$
a subset of seat-plans in $\Sigma^1_n$
which have at least one propagating part
which contains $1$ and $1'$ simultaneously.
In the diagram of the standard expression of a seat-plan of
$\Sigma^1_{n-\frac{1}{2}}$,
the vertices $1$ and $1'$ are joined by a vertical line.
Shrinking this vertical line to one vertex,
we have one to one correspondences between $\Sigma^1_{n-\frac{1}{2}}$
and
the set of the set-partitions of order $2n-1$.
(Hence we find $|\Sigma^1_{n-\frac{1}{2}}| = B_{2n-1}$, the Bell number.)
Under this preparation, we prove the theorem.
Since the relations in the theorem allow us to use
all the required local moves, we can show
just in the course of the arguments of Section~4
that any word in the alphabet ${\cal L}^1_{n-\frac{1}{2}}$
is equal to (possibly a scalar multiple of) a standard expression
in the abstract algebra $\widetilde{A_{n-\frac{1}{2}}(Q)}$.
Hence we have
\[
\mbox{rank}\ \widetilde{A_{n-\frac{1}{2}}}(Q)
\leq |\Sigma_{n-\frac{1}{2}}^1|.
\]
As Murtin and Rollet showed in \cite{MR},
$\Sigma_{n-\frac{1}{2}}^1$ makes a basis of
$\mbox{${\mathbb C}$}\otimes A_{n-\frac{1}{2}}(k)
= \mbox{${\mathbb C}$}\otimes \psi(\widetilde{A_{n-\frac{1}{2}}(k)})$
if $k>n$.
Hence $\mbox{rank}\ \mbox{${\mathbb C}$}\otimes A_{n}(z) = |\Sigma_{n-\frac{1}{2}}^1|$
holds as far as $z$ takes any integer value $k> n$.
This implies that $\psi$ is an isomorphism
and we find that the generators and the relations
in the theorem characterize the subalgebra
$A_{n-\frac{1}{2}}(Q)$.
\end{proof}
\section{Bratteli diagram of the partition algebras}\label{sec:bra}
In this section, we get back to the original definition
of $A_{n-\frac{1}{2}}(Q)$.
({\it i. e.} $A_{n-\frac{1}{2}}(Q)$ is
generated by $s_1, \ldots, s_{n-2}$, $f_1, \ldots, f_{n-1}$
and $e_1,\ldots, e_{n-1}$.)
Since, $A_{n-\frac{1}{2}}(Q)$ contains all the generators of
${A}_{n-1}(Q)$, it becomes
a subalgebra of ${A}_{n-\frac{1}{2}}(Q)$.
Hence we obtain the sequence of inclusions
$A_0(Q)\subset A_\frac{1}{2}(Q) \subset \cdots \subset A_{i-\frac{1}{2}}(Q)
\subset A_{i}(Q)\subset A_{i+\frac{1}{2}}(Q)\subset \cdots$.
First we define a graph $\Gamma_n$ [resp. $\Gamma_{n+\frac{1}{2}}$]
for a non-negative integer $n\in\mathbb{Z}_{\geq 0}$.
Then we define the sets of {\em tableaux} as sets of paths on this graph.
Figure~\ref{fig:brad} will help the reader to understand
the recipe.
\fig{fig:brad}{$\Gamma_4$}{18.eps}
For the moment, we assume that $Q$ is a sufficiently large integer.
Let
$\lambda = (\lambda_1, \lambda_2, \ldots, \lambda_l)$
be a partition.
For this $\lambda$,
define
\begin{eqnarray*}
\widetilde{\lambda}
&=& (Q-|\lambda|, \lambda_1, \lambda_2, \ldots, \lambda_l)\\
\big[\mbox{resp.}\ \widehat{\lambda}
&=& (Q-1-|\lambda|, \lambda_1, \lambda_2, \ldots, \lambda_l)\big]
\end{eqnarray*}
to be a partition of size $Q$ [resp. $Q-1$].
Pictorially,
$\widetilde{\lambda}$
[resp. $\widehat{\lambda}$]
is obtained by adding $Q-|\lambda|$
[resp. $Q-1-|\lambda|$] boxes on the top of $\lambda$.
Let
$
P_{\leq i}
= \bigcup_{j=0}^i\{ \lambda\ |\ \lambda\vdash j\}
$
be a set of Young diagrams of size less than or equal to $i$.
We define $\mbox{\boldmath $\Lambda$}_i$ and $\mbox{\boldmath $\Lambda$}_{i+\frac{1}{2}}$ to be
\[
\mbox{\boldmath $\Lambda$}_i = \{\widetilde{\lambda}\ |\ \lambda\in P_{\leq i}\}
\mbox{ and }
\mbox{\boldmath $\Lambda$}_{i+\frac{1}{2}} = \{\widehat{\lambda}\ |\ \lambda\in P_{\leq i}\},
\]
which are set of Young diagrams of size $Q$ and $Q-1$ respectively.
Under these preparations
we define a graph $\Gamma_n$ [resp. $\Gamma_{n+\frac{1}{2}}$]
which consists of the vertices labeled by:
\[
\left(
\bigsqcup_{i=0,1, \ldots, n-1}
(\mbox{\boldmath $\Lambda$}_i \sqcup \mbox{\boldmath $\Lambda$}_{i+\frac{1}{2}})
\right)
\bigsqcup\mbox{\boldmath $\Lambda$}_n
\quad
\left[ \mbox{resp.}\
\left(
\bigsqcup_{i=0,1, \ldots, n}
(\mbox{\boldmath $\Lambda$}_i \sqcup \mbox{\boldmath $\Lambda$}_{i+\frac{1}{2}})
\right)
\right]
\]
and the edges joined by either of the following rule:
\begin{itemize}
\item join $\widetilde{\lambda}\in\mbox{\boldmath $\Lambda$}_{i}$
and $\widehat{\mu}\in\mbox{\boldmath $\Lambda$}_{i+\frac{1}{2}}$ if
$\widehat{\mu}$ is obtained from $\widetilde{\lambda}$
by removing a box ($i = 0, 1, 2, \ldots n-1$)
[resp. ($i=0, 1, 2, \dots, n$)],
\item
join $\widehat{\mu}\in\mbox{\boldmath $\Lambda$}_{i-\frac{1}{2}}$
and $\widetilde{\lambda}\in\mbox{\boldmath $\Lambda$}_i$ if
$\widetilde{\lambda}$ is obtained from $\widehat{\mu}$
by adding a box ($i = 1, 2, \ldots n$).
\end{itemize}
For a pair of Young diagrams $(\mbox{\boldmath $\alpha$}, \mbox{\boldmath $\beta$})$,
if $\mbox{\boldmath $\beta$}$ is obtained from $\mbox{\boldmath $\alpha$}$ by one of the
method above, we write this as $\mbox{\boldmath $\alpha$}\smile\mbox{\boldmath $\beta$}$.
Finally, we define the sets of the tableaux.
For a half integer $n\in\frac{1}{2}\mathbb{Z}$
and $\mbox{\boldmath $\alpha$}\in\mbox{\boldmath $\Lambda$}_n$,
we define ${\mathbb T}(\mbox{\boldmath $\alpha$})$, {\em tableaux of shape $\mbox{\boldmath $\alpha$}$},
to be
\begin{eqnarray*}
{\mathbb T}(\mbox{\boldmath $\alpha$})&=&
\{P = (\mbox{\boldmath $\alpha$}^{(0)}, \mbox{\boldmath $\alpha$}^{(1/2)}, \ldots, \mbox{\boldmath $\alpha$}^{(n)})\ |
\ \mbox{\boldmath $\alpha$}^{(j)} \in \mbox{\boldmath $\Lambda$}_j\ (j = 0, 1/2, \ldots, n),\\
& & \quad\mbox{\boldmath $\alpha$}^{(n)} = \mbox{\boldmath $\alpha$},
\mbox{\boldmath $\alpha$}^{(j)}\smile\mbox{\boldmath $\alpha$}^{(j+1/2)}
\ (j = 0, 1/2, \ldots, n-1/2)\}.
\end{eqnarray*}
\section{Construction of representation}\label{sec:rep}
Now we have defined the sets of tableaux,
we define linear transformations among the tableaux.
Let ${\mathbb Q}$ be the field of rational numbers
and $K_0 = {\mathbb Q}(Q)$ its extension.
In the following, the linear transformations are defined over $K_0$.
If they preserve the relations defined in the previous sections,
they define representations
of ${A}_n = {A}_n(Q)\otimes K_0$.
Similar methods are used for example
in the references~\cite{AK,GHJ,Mu,W1,W2,Ko2}.
Let ${\mathbb V}(\mbox{\boldmath $\alpha$})
= \oplus_{P \in {\mathbb T}(\mbox{\boldmath $\alpha$}) }K_0 v_P$ be
a vector space over $K_0$ with the standard basis
$\{v_P|P\in {\mathbb T}(\mbox{\boldmath $\alpha$})\}$.
For generators $e_i$, $f_i$ and $s_i$ of ${A}_n$,
we define linear maps $\rho_{\mbox{\boldmath $\alpha$}}(e_i)$, $\rho_{\mbox{\boldmath $\alpha$}}(f_i)$
and $\rho_{\mbox{\boldmath $\alpha$}}(s_i)$
on ${\mathbb V}(\mbox{\boldmath $\alpha$})$ giving the matrices
$E_i$ $F_i$ and $M_i$ respectively with respect to
the basis $\{ v_P | P\in {\mathbb T}(\mbox{\boldmath $\alpha$}) \}$.
\subsection{Definition of $\rho_{\mbox{\boldmath $\alpha$}}(e_i)$}
Firstly, we define a linear map for $e_i$.
For a tableaux
$P = (\mbox{\boldmath $\alpha$}^{(0)}, \mbox{\boldmath $\alpha$}^{(1/2)}, \ldots, \mbox{\boldmath $\alpha$}^{(n)})$
of ${\mathbb T}(\mbox{\boldmath $\alpha$})$, we define
$\rho_{\mbox{\boldmath $\alpha$}}(e_i)(v_P)
= \sum_{Q \in {\mathbb T}(\mbox{\boldmath $\alpha$})}(E_i)_{QP}v_Q$.
Let $Q = (\mbox{\boldmath $\alpha$}^{\prime(0)}, \mbox{\boldmath $\alpha$}^{\prime(1/2)},
\ldots, \mbox{\boldmath $\alpha$}^{\prime(n)})$.
If there is an
$i_0 \in \{1/2, 1, \ldots, n-1/2 \} \setminus \{i-1/2\}$
such that $\mbox{\boldmath $\alpha$}^{(i_0)}\neq \mbox{\boldmath $\alpha$}^{\prime(i_0)}$,
then we put
\[
(E_i)_{QP} = 0.
\]
In the following, we consider the case that
$\mbox{\boldmath $\alpha$}^{(i_0)} = \mbox{\boldmath $\alpha$}^{\prime(i_0)}$
for $i_0\in\{0, 1/2, 1, \ldots, n-1/2\}\setminus\{i-1/2\}$.
If
$\mbox{\boldmath $\alpha$}^{(i-1)}$ and $\mbox{\boldmath $\alpha$}^{(i)}$
are not labeled by the same Young diagram,
then we put
\[
(E_i)_{QP} = 0.
\]
We consider the case $\mbox{\boldmath $\alpha$}^{(i-1)}$ and $\mbox{\boldmath $\alpha$}^{(i)}$
have the same label $\widetilde{\lambda}$.
In this case, the possible vertices as $\mbox{\boldmath $\alpha$}^{(i-1/2)}$
have labels $\{\widetilde{\lambda}^{-}_{(s)}\}$,
which are obtained by removing one box from $\widetilde{\lambda}$.
Let $\{Q_s\}$ be the set of tableaux
obtained from $P$ by replacing $\mbox{\boldmath $\alpha$}^{(i-1/2)}$
with $\widetilde{\lambda}^{-}_{(s)}$.
Then we define $(E_i)_{QP}$ to be
\[
(E_i)_{Q_sP}
= \frac{h(\widetilde{\lambda})}{h(\widetilde{\lambda}^{-}_{(s)})}.
\]
Here $h(\lambda)$ is the product of hook lengths defined by
\[
h(\lambda) = \prod_{x\in\lambda} h_{\lambda}(x)
\]
and $h_{\lambda}(x)$ is the {\em hook-length} at $x\in\lambda$.
Note that the matrix $E_i$ is determined
by
the label $\widetilde{\lambda}$ itself
not by the vertex at which the tableau $P$ goes through.
In other words, if another vertex in different level, say $i'$,
has the same label $\widetilde{\lambda}$, then $E_{i'}$ becomes
the same matrix.
Let $v(\lambda^-_{(s)}, \lambda)$ be the standard vector
which corresponds to a tableau whose $(i-1)$-st, $(i-1/2)$-th and $i$-th
coordinate $(\mbox{\boldmath $\alpha$}^{(i-1)}, \mbox{\boldmath $\alpha$}^{(i-1/2)}, \mbox{\boldmath $\alpha$}^{(i)})$
are labeled by $(\lambda, \lambda^-_{(s)}, \lambda)$.
Then for a tableau $P$ which goes through
$\lambda$ at the $(i-1)$-st and the $i$-th coordinate of $P$,
we have
\[
\rho(e_i)(v_P)
=
\sum_{s'} \frac{h(\lambda)}{h(\lambda^{-}_{(s')})}v(\lambda^-_{(s')}, \lambda).
\]
Here $\lambda^{-}_{(s')}$ runs through Young diagrams
obtained from $\lambda$ by removing one box.
\fig{fig:repE4a}{Representation spaces for $\rho(e_i)$}{19.eps}
\fig{fig:repE4b}{Representation spaces for $\rho(e_i)$}{20.eps}
\begin{example}
Suppose that tableaux $\{p_r\}$ goes through
paths in pictures illustrated in Figure~\ref{fig:repE4a}
or \ref{fig:repE4b}.
Then we have
\begin{eqnarray*}
\rho(e_i)(v_0) &=&
\frac{h(\widetilde{\emptyset})}{h(\widehat{\emptyset})}v_0 = Qv_0,\\
\rho(e_i)(v_1\ v_2) &=& (v_1\ v_2)
\begin{pmatrix}
h(\widetilde{\mbox{\tiny\yng(1)}})/h(\widehat{\emptyset})
&h(\widetilde{\mbox{\tiny\yng(1)}})/h(\widehat{\emptyset})\\
h(\widetilde{\mbox{\tiny\yng(1)}})/h(\widehat{\mbox{\tiny\yng(1)}})
&h(\widetilde{\mbox{\tiny\yng(1)}})/h(\widehat{\mbox{\tiny\yng(1)}})
\end{pmatrix}\\
&=& (v_1\ v_2)
\begin{pmatrix}
\frac{Q}{Q-1} &\frac{Q}{Q-1}\\
\frac{Q(Q-2)}{Q-1} &\frac{Q(Q-2)}{Q-1}
\end{pmatrix}
\\
\rho(e_i)(v_3\ v_4) &=& (v_3\ v_4)
\begin{pmatrix}
h(\widetilde{\mbox{\tiny\yng(2)}})/h(\widehat{\mbox{\tiny\yng(1)}})
&h(\widetilde{\mbox{\tiny\yng(2)}})/h(\widehat{\mbox{\tiny\yng(1)}})\\
h(\widetilde{\mbox{\tiny\yng(2)}})/h(\widehat{\mbox{\tiny\yng(2)}})
&h(\widetilde{\mbox{\tiny\yng(2)}})/h(\widehat{\mbox{\tiny\yng(2)}})
\end{pmatrix},\\
&=& (v_3\ v_4)
\begin{pmatrix}
\frac{2(Q-2)}{Q-3} &\frac{2(Q-2)}{Q-3}\\
\frac{(Q-1)(Q-4)}{Q-3} &\frac{(Q-1)(Q-4)}{Q-3}
\end{pmatrix}\\
\rho(e_i)(v_5\ v_6) &=& (v_5\ v_6)
\begin{pmatrix}
h(\widetilde{\mbox{\tiny\yng(1,1)}})/h(\widehat{\mbox{\tiny\yng(1)}})
&h(\widetilde{\mbox{\tiny\yng(1,1)}})/h(\widehat{\mbox{\tiny\yng(1)}})\\
h(\widetilde{\mbox{\tiny\yng(1,1)}})/h(\widehat{\mbox{\tiny\yng(1,1)}})
&h(\widetilde{\mbox{\tiny\yng(1,1)}})/h(\widehat{\mbox{\tiny\yng(1,1)}})
\end{pmatrix},\\
&=& (v_5\ v_6)
\begin{pmatrix}
\frac{2Q}{Q-1} &\frac{2Q}{Q-1}\\
\frac{Q(Q-3)}{Q-1} &\frac{Q(Q-3)}{Q-1}
\end{pmatrix}.
\end{eqnarray*}
Here $v_i$ is the standard vector which corresponds to $p_i$.
Similarly for the bases $\langle v_7, v_8\rangle$,
$\langle v_9, v_{10}, v_{11}\rangle$
and
$\langle v_{12}, v_{13}\rangle$,
we have the following matrices respectively:
\begin{eqnarray*}
&&\begin{pmatrix}
\frac{3(Q-4)}{Q-5} &\frac{3(Q-4)}{Q-5}\\
\frac{(Q-2)(Q-6)}{Q-5} &\frac{(Q-2)(Q-6)}{Q-5}
\end{pmatrix},\\
&&
\begin{pmatrix}
\frac{3(Q-1)}{2(Q-2)} &\frac{3(Q-1)}{2(Q-2)} &\frac{3(Q-1)}{2(Q-2)}\\
\frac{3(Q-3)}{2(Q-4)} &\frac{3(Q-3)}{2(Q-4)} &\frac{3(Q-3)}{2(Q-4)}\\
\frac{(Q-1)(Q-3)(Q-5)}{(Q-2)(Q-4)}&
\frac{(Q-1)(Q-3)(Q-5)}{(Q-2)(Q-4)}&
\frac{(Q-1)(Q-3)(Q-5)}{(Q-2)(Q-4)}
\end{pmatrix},\\
&&
\begin{pmatrix}
\frac{3Q}{Q-1} &\frac{3Q}{Q-1}\\
\frac{Q(Q-4)}{Q-1} &\frac{Q(Q-4)}{Q-1}
\end{pmatrix}.
\end{eqnarray*}
\end{example}
\subsection{Definition of $\rho_{\mbox{\boldmath $\alpha$}}(f_i)$}
Next, we define a linear map for $f_i$.
For a tableaux
$P = (\mbox{\boldmath $\alpha$}^{(0)}, \mbox{\boldmath $\alpha$}^{(1/2)}, \ldots, \mbox{\boldmath $\alpha$}^{(n)})$
of ${\mathbb T}(\mbox{\boldmath $\alpha$})$, we define
$\rho_{\mbox{\boldmath $\alpha$}}(f_i)(v_P)
= \sum_{Q \in {\mathbb T}(\mbox{\boldmath $\alpha$})}(F_i)_{QP}v_Q$.
Let $Q = (\mbox{\boldmath $\alpha$}^{\prime(0)}, \mbox{\boldmath $\alpha$}^{\prime(1/2)},
\ldots, \mbox{\boldmath $\alpha$}^{\prime(n)})$.
If there is an
$i_0 \in \{1/2, 1, \ldots, n-1/2 \} \setminus \{i\}$
such that $\mbox{\boldmath $\alpha$}^{(i_0)}\neq \mbox{\boldmath $\alpha$}^{\prime(i_0)}$,
then we put
\[
(F_i)_{QP} = 0.
\]
In the following, we consider the case that
$\mbox{\boldmath $\alpha$}^{(i_0)} = \mbox{\boldmath $\alpha$}^{\prime(i_0)}$
for $i_0\in\{0, 1/2, 1, \ldots, n-1/2\}\setminus\{i\}$.
If
$\mbox{\boldmath $\alpha$}^{(i-1/2)}$ and $\mbox{\boldmath $\alpha$}^{(i+1/2)}$
are not labeled by the same Young diagram,
then we put
\[
(F_i)_{QP} = 0.
\]
We consider the case $\mbox{\boldmath $\alpha$}^{(i-1/2)}$ and $\mbox{\boldmath $\alpha$}^{(i+1/2)}$
have the same label $\widehat{\mu}$.
In this case, the possible vertices as $\mbox{\boldmath $\alpha$}^{(i)}$
have labels $\{\widehat{\mu}^{+}_{(r)}\}$,
which are obtained by adding one box to $\widetilde{\mu}$.
Suppose that $\mbox{\boldmath $\alpha$}^{(i)}$,
the $i$-th coordinate of $P$,
has its label $\widetilde{\mu}^{+}_{(r_0)}$.
Let $Q$ be a tableau
obtained from $P$ by replacing $\mbox{\boldmath $\alpha$}^{(i)}$
with one of $\{\widehat{\mu}^{+}_{(r)}\}$.
Then we define $(F_i)_{QP}$ to be
\[
(F_i)_{Q_rP} = \frac{h(\widehat{\mu})}{h(\widehat{\mu}^{+}_{(r_0)})}.
\]
Let $v(\mu^+_{(r)}, \mu)$ be the standard vector
which corresponds to a tableau whose $(i-1/2)$-th, $i$-th and $(i+1/2)$-th
coordinate $(\mbox{\boldmath $\alpha$}^{(i-1/2)}, \mbox{\boldmath $\alpha$}^{(i)}, \mbox{\boldmath $\alpha$}^{(i+1/2)})$
are labeled by $(\mu, \mu^+_{(r)}, \mu)$.
Then for a tableau $P$ which goes through
$\mu$ at the $(i-1/2)$-th and the $(i+1/2)$-th coordinate of $P$,
we have
\[
\rho(f_i)(v_P)
=
\sum_{r} \frac{h(\mu)}{h(\mu^{+}_{(r_0)})}v(\mu^+_{(r)}, \mu).
\]
Here $\mu^{+}_{(r)}$ runs through Young diagrams
obtained from $\mu$ by adding one box.
\fig{fig:repF4}{Representation spaces for $\rho(f_i)$}{21.eps}
\begin{example}
Suppose that tableau $\{q_r\}$ go through
paths in the picture illustrated in Figure~\ref{fig:repF4}.
Then we have
\[
\rho(f_i)(v_0\ v_1)
= (v_0\ v_1)
\begin{pmatrix}
h(\widehat{\emptyset})/h(\widetilde{\emptyset})
&h(\widehat{\emptyset})/h(\widetilde{\mbox{\tiny\yng(1)}})\\
h(\widehat{\emptyset})/h(\widetilde{\emptyset})
&h(\widehat{\emptyset})/h(\widetilde{\mbox{\tiny\yng(1)}})
\end{pmatrix}
= (v_0\ v_1)
\begin{pmatrix}
\frac{1}{Q} &\frac{Q-1}{Q}\\
\frac{1}{Q} &\frac{Q-1}{Q}
\end{pmatrix}
\]
and
\begin{eqnarray*}
\rho(f_i)(v_2\ v_{3}\ v_{4})
&=& (v_2\ v_{3}\ v_{4})
\begin{pmatrix}
h(\widehat{\mbox{\tiny\yng(1)}})/h(\widetilde{\mbox{\tiny\yng(1)}})
&h(\widehat{\mbox{\tiny\yng(1)}})/h(\widetilde{\mbox{\tiny\yng(2)}})
&h(\widehat{\mbox{\tiny\yng(1)}})/h(\widetilde{\mbox{\tiny\yng(1,1)}})\\
h(\widehat{\mbox{\tiny\yng(1)}})/h(\widetilde{\mbox{\tiny\yng(1)}})
&h(\widehat{\mbox{\tiny\yng(1)}})/h(\widetilde{\mbox{\tiny\yng(2)}})
&h(\widehat{\mbox{\tiny\yng(1)}})/h(\widetilde{\mbox{\tiny\yng(1,1)}})\\
h(\widehat{\mbox{\tiny\yng(1)}})/h(\widetilde{\mbox{\tiny\yng(1)}})
&h(\widehat{\mbox{\tiny\yng(1)}})/h(\widetilde{\mbox{\tiny\yng(2)}})
&h(\widehat{\mbox{\tiny\yng(1)}})/h(\widetilde{\mbox{\tiny\yng(1,1)}})
\end{pmatrix}\\
&=& (v_2\ v_{3}\ v_{4})
\begin{pmatrix}
\frac{Q-1}{Q(Q-2)} &\frac{Q-3}{2(Q-2)} &\frac{Q-1}{2Q}\\
\frac{Q-1}{Q(Q-2)} &\frac{Q-3}{2(Q-2)} &\frac{Q-1}{2Q}\\
\frac{Q-1}{Q(Q-2)} &\frac{Q-3}{2(Q-2)} &\frac{Q-1}{2Q}
\end{pmatrix}.
\end{eqnarray*}
Here $v_i$ is the standard vector which corresponds to $q_i$.
Similarly, for the bases
$\langle v_5, v_6, v_7\rangle$ and
$\langle v_8, v_{9}, v_{10}\rangle$
we have the following matrices respectively:
\[
\begin{pmatrix}
\frac{Q-3}{(Q-1)(Q-4)} &\frac{Q-5}{3(Q-4)} &\frac{2(Q-2)}{3(Q-1)}\\
\frac{Q-3}{(Q-1)(Q-4)} &\frac{Q-5}{3(Q-4)} &\frac{2(Q-2)}{3(Q-1)}\\
\frac{Q-3}{(Q-1)(Q-4)} &\frac{Q-5}{3(Q-4)} &\frac{2(Q-2)}{3(Q-1)}
\end{pmatrix},\quad
\begin{pmatrix}
\frac{Q-1}{Q(Q-3)} &\frac{2(Q-4)}{3(Q-3)} &\frac{Q-1}{3Q}\\
\frac{Q-1}{Q(Q-3)} &\frac{2(Q-4)}{3(Q-3)} &\frac{Q-1}{3Q}\\
\frac{Q-1}{Q(Q-3)} &\frac{2(Q-4)}{3(Q-3)} &\frac{Q-1}{3Q}
\end{pmatrix}.
\]
\end{example}
\subsection{Definition of $\rho_{\mbox{\boldmath $\alpha$}}(s_i)$}
Finally, we define linear maps for $s_i$.
Unfortunately,
we do not have uniform description for $\rho_{\mbox{\boldmath $\alpha$}}(s_i)$,
except for ``non-reductive'' paths.
So first we define $\rho_{\mbox{\boldmath $\alpha$}}(s_i)$
for the non-reductive paths.
Then we define $\rho_{\mbox{\boldmath $\alpha$}}(s_1)$
and $\rho_{\mbox{\boldmath $\alpha$}}(s_2)$ for ``reductive'' paths one by one.
\subsubsection*{Non-Reductive Case}
In the following, we use notation $\mu\mbox{$\vartriangleleft$}\lambda$
if a Young diagram $\lambda$
is obtained from a Young diagram $\mu$ by
adding one box.
For $1\leq j\leq i$,
let ${\nu}$, ${\mu}$, ${\lambda}$
be Young diagrams of size $j-1$, $j$ and $j+1$ respectively
such that $\nu\mbox{$\vartriangleleft$}\mu\mbox{$\vartriangleleft$}\lambda$.
If a tableau $P$ of $\mathbb{T}(\mbox{\boldmath $\alpha$})$
goes through $\widetilde{\nu}$, $\widetilde{\mu}$ and $\widetilde{\lambda}$
at the $(i-2)$-nd, the $(i-1)$-st and the $i$-th coordinate,
then $P$ goes through $\widehat{\nu}$ and $\widehat{\mu}$
at the $(i-3/2)$-th and the $(i-1/2)$-th coordinate.
We call such a tableau {\em non-reductive} at $i$.
If a tableau $P$ does not satisfy the property above,
then we call $P$ {\em reductive} at $i$.
Recall that if $\nu\mbox{$\vartriangleleft$}\mu\mbox{$\vartriangleleft$}\lambda$,
then we can define the {\em axial distance} $d = d(\nu, \mu, \lambda)$.
Namely, if $\mu$ differs from $\nu$
in the $r_0$-th row and the $c_0$-th column only,
and $\lambda$ differs from $\mu$
in the $r_1$-th row and the $c_1$-th column only,
then $d = d(\nu, \mu, \lambda)$ is defined by
\begin{equation*}
d = d(\nu, \mu, \lambda) = (c_1 - r_1) - (c_0 - r_0)
= \left\{
\begin{array}{ll}
h_{\lambda}(r_1, c_0) -1 & \mbox{ if } r_0 \geq r_1,\\
1 - h_{\lambda}(r_0, c_1) & \mbox{ if } r_0 < r_1.
\end{array}
\right.
\end{equation*}
Here $h_{\lambda}(i,j)$ is the {\em hook-length} at $(i,j)$ in $\lambda$.
If $|d|\geq 2$,
then there is a unique Young diagram
$\mu^{\prime}\neq \mu$
which satisfies
$\nu\mbox{$\vartriangleleft$}\mu^{\prime}\mbox{$\vartriangleleft$}\lambda$.
Let $P'$
be a tableau of shape $\mbox{\boldmath $\alpha$}$ which are obtained from $P$
by replacing $(i-1)$-st and $(i-1/2)$-th coordinates
of $P$ with $\widetilde{\mu'}$ and $\widehat{\mu'}$ respectively.
For the standard vectors $v_P$ and $v_{P'}$ which correspond to $P$ and $P'$,
we define the linear map $\rho_{\mbox{\boldmath $\alpha$}}(s_i)$ as follows:
\begin{equation}\label{eq:non-red}
\rho_{\mbox{\boldmath $\alpha$}}(s_i)\ :\ (v_{P}, v_{P'})
\longmapsto (v_{P}, v_{P'})
\left(
\begin{array}{cc}
1/d & \left(1-1/d^2\right)/c\\
c & -1/d
\end{array}
\right),
\end{equation}
where we can arbitrarily chose $c\in K_0\setminus\{0\}$.
If we put
\begin{equation}\label{eq:ad}
a_d = 1/d\quad \mbox{and}\quad b_d = 1-a_d^2,
\end{equation}
then the matrix in the expression~\eqref{eq:non-red} is written as follows:
\begin{equation*}
\left(
\begin{array}{rr}
a_{d} & b_{d}/c\\
c & -a_{d}
\end{array}
\right).
\end{equation*}
If $|d_1|=1$,
then there does not exist a
distinct Young diagram
$\mu^{\prime}$
which satisfies
$\nu\mbox{$\vartriangleright$}\mu^{\prime}\mbox{$\vartriangleright$}\lambda$
other than $\mu$.
In this case, we define $\rho_{\mbox{\boldmath $\alpha$}}(s_i)$ to be
\[
\rho_{\mbox{\boldmath $\alpha$}}(s_i)\ :\ v_{P}
\longmapsto a_d v_{P}.
\]
Here $a_d$ is the one defined by \eqref{eq:ad}.
\begin{example}
Suppose that a tableau $p_1$ of $\mathbb{T}(\mbox{\boldmath $\alpha$})$ goes through
$\widetilde{\emptyset}$, $\widetilde{\mbox{\tiny\yng(1)}}$ and $\widetilde{\mbox{\tiny\yng(2)}}$
at the 0-th, the 1-st and the 2-nd coordinates respectively,
then for the standard vector $u_1$ which corresponds to $p_1$ we have
\[
\rho_{\mbox{\boldmath $\alpha$}}(s_1) u_1 = u_1.
\]
For the standard vector $v_2$ which corresponds to $p_2$,
a tableau of $\mathbb{T}(\mbox{\boldmath $\alpha$})$ which goes through
$\widetilde{\emptyset}$, $\widetilde{\mbox{\tiny\yng(1)}}$ and $\widetilde{\mbox{\tiny\yng(1,1)}}$
at the 0-th, the 1-st and the 2-nd coordinates respectively,
we have
\[
\rho_{\mbox{\boldmath $\alpha$}}(s_1) u_2 = -u_2.
\]
\end{example}
\begin{example}
Let $\lambda^{(1)} = (3)$, $\lambda^{(2)} = (2,1)$
and $\lambda^{(3)} = (1,1,1)$ be partitions of 3.
Suppose that
tableaux $q_1$ and $q_2$ of $\mathbb{T}(\mbox{\boldmath $\alpha$})$ both go through
$\widetilde{\mbox{\tiny\yng(1)}}$ and $\widetilde{\mbox{\tiny\yng(2)}}$
at the 1-st and the 2-nd coordinates respectively,
and
tableaux $q_3$ and $q_4$ of $\mathbb{T}(\mbox{\boldmath $\alpha$})$ both go through
$\widetilde{\mbox{\tiny\yng(1)}}$ and $\widetilde{\mbox{\tiny\yng(1,1)}}$
at the 1-st and the 2-nd coordinates respectively.
Further, the tableaux $q_1$, $q_2$, $q_3$ and $q_4$
go through $\widetilde{\lambda^{(1)}}$, $\widetilde{\lambda^{(2)}}$,
$\widetilde{\lambda^{(2)}}$ and $\widetilde{\lambda^{(3)}}$
at the 3-rd coordinates respectively.
Then we have
\[
\rho_{\mbox{\boldmath $\alpha$}}(s_2) (v_1\ v_2\ v_3\ v_4)
= (v_1\ v_2\ v_3\ v_4)
\begin{pmatrix}
1 & 0 & 0 & 0\\
0 & -1/2& 3/(4c) & 0\\
0 & c & 1/2 & 0\\
0 & 0 & 0 & -1
\end{pmatrix}.
\]
Here $v_i$ is the standard vector which corresponds to $q_i$.
\end{example}
\subsubsection*{Reductive Case}
Consider the case a tableau $P$ is reductive at $i$.
So far,
we do not have uniform description for $\rho_{\mbox{\boldmath $\alpha$}}(s_i)$.
So we define $\rho_{\mbox{\boldmath $\alpha$}}(s_1)$
and $\rho_{\mbox{\boldmath $\alpha$}}(s_2)$ one by one.
First we define $\rho_{\mbox{\boldmath $\alpha$}}(s_1)$.
For tableaux $p_1$ and $p_2$ of $\mathbb{T}(\mbox{\boldmath $\alpha$})$
which go through
$(\widetilde{\emptyset}, \widehat{\emptyset},
\widetilde{\emptyset}, \widehat{\emptyset}, \widetilde{\emptyset})$
and
$(\widetilde{\emptyset}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset}, \widetilde{\emptyset})$
at the 0-th, the $1-\frac{1}{2}$-th, the 1-st,
the $2-\frac{1}{2}$-th and the 2-nd coordinate
respectively,
let $u_1$ and $u_2$ be the corresponding standard vectors.
Then we define $\rho_{\mbox{\boldmath $\alpha$}}(s_1)(u_1\ u_2)$ by
\[
\rho_{\mbox{\boldmath $\alpha$}}(s_1)(u_1\ u_2)
= (u_1\ u_2)
\begin{pmatrix}
1 & 0\\
0 & 1
\end{pmatrix}.
\]
For tableaux $p_3$, $p_4$ and $p_5$ of $\mathbb{T}(\mbox{\boldmath $\alpha$})$
which go through
$(\widetilde{\emptyset}, \widehat{\emptyset},
\widetilde{\emptyset}, \widehat{\emptyset}, \widetilde{\mbox{\tiny\yng(1)}})$,
$(\widetilde{\emptyset}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset}, \widetilde{\mbox{\tiny\yng(1)}})$
and
$(\widetilde{\emptyset}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(1)}})$
at the 0-th, the $1-\frac{1}{2}$-th, the 1-st,
the $2-\frac{1}{2}$-th and the 2-nd coordinate
respectively,
let $u_3$, $u_4$ and $u_5$ be the corresponding standard vectors.
Then we define $\rho_{\mbox{\boldmath $\alpha$}}(s_1)(u_1\ u_2\ u_3)$ by
\[
\rho_{\mbox{\boldmath $\alpha$}}(s_1)(u_1\ u_2\ u_3)
= (u_1\ u_2\ u_3)
\begin{pmatrix}
0 & 1 & 1\\
\frac{1}{Q-1} & \frac{Q-2}{Q-1} & \frac{-1}{Q-1}\\
\frac{Q-2}{Q-1} & -\frac{Q-2}{Q-1} & \frac{1}{Q-1}
\end{pmatrix}.
\]
Next we define $\rho_{\mbox{\boldmath $\alpha$}}(s_2)$.
In the following, we write
\[
p = (\lambda^{(1)}, \lambda^{(2)}, \lambda^{(3)},
\lambda^{(4)}, \lambda^{(5)})
\]
to mean the tableau $p$
goes through $\lambda^{(1)}$, $\lambda^{(2)}$, $\lambda^{(3)}$,
$\lambda^{(4)}$, $\lambda^{(5)}$
at the 1-st, the $(2-\frac{1}{2})$-th,
the 2-nd, the $(3-\frac{1}{2})$-th
and the 3-rd coordinates respectively.
Suppose that
\[
\begin{array}{lr}
\begin{array}{rcl}
q_1 &=& (\widetilde{\emptyset}, \widehat{\emptyset},
\widetilde{\emptyset}, \widehat{\emptyset}, \widetilde{\emptyset}),\\
q_2 &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset},
\widetilde{\emptyset}, \widehat{\emptyset}, \widetilde{\emptyset}),\\
q_3 &=& (\widetilde{\emptyset}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset}, \widetilde{\emptyset}),
\end{array}
&
\begin{array}{rcl}
q_4 &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset}, \widetilde{\emptyset}),\\
q_5 &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset}, \widetilde{\emptyset}).
\end{array}
\end{array}
\]
Then for the standard vectors $(v_j)_{j=1}^5$
which correspond to $(q_j)_{j=1}^5$
we define $\rho_{\mbox{\boldmath $\alpha$}}(s_2)(v_1\ v_2\ v_3\ v_4\ v_5)$ by
\[
\rho_{\mbox{\boldmath $\alpha$}}(s_2)(v_1\ v_2\ v_3\ v_4\ v_5)
= (v_1\ v_2\ v_3\ v_4\ v_5)
\begin{pmatrix}
1 & 0 & 0& 0 &0\\
0 & 0 & 0& 1 &1\\
0 & 0 & 1& 0 &0\\
0 & \frac{1}{Q-1} & 0& \frac{Q-2}{Q-1} &\frac{-1}{Q-1}\\
0 & \frac{Q-2}{Q-1} & 0& -\frac{Q-2}{Q-1} &\frac{1}{Q-1}
\end{pmatrix}.
\]
Assume that
\[
\begin{array}{lr}
\begin{array}{rcl}
q_6 &=& (\widetilde{\emptyset}, \widehat{\emptyset},
\widetilde{\emptyset}, \widehat{\emptyset}, \widetilde{\mbox{\tiny\yng(1)}}),\\
q_7 &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset},
\widetilde{\emptyset}, \widehat{\emptyset}, \widetilde{\mbox{\tiny\yng(1)}}),\\
q_8 &=& (\widetilde{\emptyset}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset}, \widetilde{\mbox{\tiny\yng(1)}}),\\
q_9 &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset}, \widetilde{\mbox{\tiny\yng(1)}}),\\
q_{10} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset}, \widetilde{\mbox{\tiny\yng(1)}}),
\end{array}
&
\begin{array}{rcl}
q_{11} &=& (\widetilde{\emptyset}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(1)}}),\\
q_{12} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(1)}}),\\
q_{13} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(1)}}),\\
q_{14} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(2)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(1)}}),\\
q_{15} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(1,1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(1)}}).
\end{array}
\end{array}
\]
Then for the standard vectors $(v_j)_{j=6}^{15}$
which correspond to $(q_j)_{j=6}^{15}$
we define $\rho_{\mbox{\boldmath $\alpha$}}(s_2)(v_6\ v_8\ v_{11})$ and
$\rho_{\mbox{\boldmath $\alpha$}}(s_2)(v_7\ v_9\ v_{10} \ v_{12}\ v_{13}\ v_{14}\ v_{15})$
by
\[
\rho_{\mbox{\boldmath $\alpha$}}(s_2)(v_6\ v_8\ v_{11})
= (v_6\ v_8\ v_{11})
\begin{pmatrix}
0&1&1\\
\noalign{\medskip}
\frac{1}{(Q-1)}
&\frac{Q-2}{Q-1}
&\frac{-1}{(Q-1)}\\
\noalign{\medskip}
\frac{Q-2}{Q-1}
&-\frac{Q-2}{Q-1}
&\frac{1}{Q-1}
\end{pmatrix}
\]
and
\[
\rho_{\mbox{\boldmath $\alpha$}}(s_2)(v_7\ v_9\ v_{10}
\ v_{12}\ v_{13}\ v_{14}\ v_{15})
= (v_7\ v_9\ v_{10}
\ v_{12}\ v_{13}\ v_{14}\ v_{15})M_i
\]
Here the matrix $M_{i}$ is
\[
\begin{pmatrix}
\noalign{\medskip}
\frac{1}{Q-1}
&\frac{Q-2}{Q-1}
&\frac{-1}{Q-1}
&\frac{-1}{Q-1}
&\frac{1}{(Q-1)(Q-2)}
&\frac{(Q-1)(Q-2)-2}{2(Q-1)(Q-2)}
&-1/2\\
\noalign{\medskip}
\frac{Q-2}{(Q-1)^{2}}
&\frac{Q^2-3Q+3}{(Q-1)^2}
&\frac{1}{(Q-1)^2}
&\frac{1}{(Q-1)^2}
&\frac{-1}{(Q-1)^2(Q-2)}
&\frac{-Q(Q-3)}{2(Q-1)^2(Q-2)}
&\frac{1}{2(Q-1)}\\
\noalign{\medskip}
\frac{-(Q-2)}{(Q-1)^2}
&\frac{Q-2}{(Q-1)^2}
&\frac{-1}{(Q-1)^2}
&\frac{Q(Q-2)}{(Q-1)^2}
&\frac{1}{(Q-1)^2(Q-2)}
&\frac{Q(Q-3)}{2(Q-1)^2(Q-2)}
&\frac{-1}{2(Q-1)}\\
\noalign{\medskip}
\frac{-(Q-2)}{(Q-1)^2}
&\frac{Q-2}{(Q-1)^2}
&\frac{Q(Q-2)}{(Q-1)^2}
&\frac{-1}{(Q-1)^2}
&\frac{1}{(Q-1)^2(Q-2)}
&\frac {Q(Q-3)}{2(Q-1)^2(Q-2)}
&\frac{-1}{2(Q-1)}\\
\noalign{\medskip}
\frac{Q-2}{(Q-1)^2}
&\frac{-(Q-2)}{(Q-1)^2}
&\frac{1}{(Q-1)^2}
&\frac{1}{(Q-1)^2}
&\frac{(Q-1)^2(Q-2)-1}{(Q-1)^2(Q-2)}
&\frac{-Q(Q-3)}{2(Q-1)^2(Q-2)}
&\frac{1}{2(Q-1)}\\
\noalign{\medskip}
\frac{Q-2}{Q-1}
&\frac{-(Q-2)}{Q-1}
&\frac{1}{Q-1}
&\frac{1}{Q-1}
&\frac{-1}{(Q-1)(Q-2)}
&\frac{Q^2-3Q+4}{2(Q-1)(Q-2)}
&1/2\\
\noalign{\medskip}
\frac{-(Q-2)}{Q-1}
&\frac{Q-2}{Q-1}
&\frac{-1}{Q-1}
&\frac{-1}{Q-1}
&{\frac{1}{(Q-1)(Q-2)}}
&\frac{Q(Q-3)}{2(Q-1)(Q-2)}
&1/2
\end{pmatrix}.
\]
Next assume that
\[
\begin{array}{lr}
\begin{array}{rcl}
q_{16} &=& (\widetilde{\emptyset}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(2)}}),\\
q_{17} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(2)}}),\\
q_{18} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(2)}}),
\end{array}
&
\begin{array}{rcl}
q_{19} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(2)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(2)}}),\\
q_{20} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(1,1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(2)}}),\\
q_{21} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(2)}}, \widehat{\mbox{\tiny\yng(2)}}, \widetilde{\mbox{\tiny\yng(2)}}).
\end{array}
\end{array}
\]
Then for the standard vectors $(v_j)_{j=16}^{21}$
which correspond to $(q_j)_{j=16}^{21}$
we define $\rho_{\mbox{\boldmath $\alpha$}}(s_2)(v_{16}\ v_{17}\ v_{18}\ v_{19}\ v_{20}\ v_{21}
)$ by
\[
\rho_{\mbox{\boldmath $\alpha$}}(s_2)(v_{16}\ v_{17}\ v_{18}\ v_{19}\ v_{20}\ v_{21})
= (v_{16}\ v_{17}\ v_{18}\ v_{19}\ v_{20}\ v_{21})M_i.
\]
Here the matrix $M_i$ is
\begin{eqnarray*}
\begin{pmatrix}
1
&0
&0
&0
&0
&0\\
\noalign{\medskip}0
&\frac{-1}{(Q-1)}
&\frac{1}{(Q-1)(Q-2)}
&\frac{Q(Q-3)}{2(Q-1)(Q-2)}
&-1/2
&0\\
\noalign{\medskip}0
&\frac{1}{(Q-1)}
&\frac{-1}{(Q-1)(Q-2)}
&\frac{{Q}^{2}-3\,Q+4}{2(Q-1)(Q-2)}
&1/2
&1\\
\noalign{\medskip}0
&1
&\frac{{Q}^{2}-3\,Q+4}{Q(Q-3)(Q-2)}
&\frac{(Q-1)( Q-4)}{2( Q-2)(Q-3)}
&\frac{(Q-1)(Q-4)}{2Q(Q-3)}
&\frac{-1}{(Q-3)}\\
\noalign{\medskip}0
&-1
&\frac{1}{(Q-2)}
&\frac {Q-4}{2(Q-2)}
&1/2
&\frac{-1}{(Q-1)}\\
\noalign{\medskip}0
&0
&\frac{(Q-1)^{2}(Q-4)}{Q(Q-3)(Q-2)}
&\frac{-(Q-1)(Q-4)}{2(Q-2)(Q-3)}
&\frac{-(Q-1)(Q-4)}{2Q(Q-3)}
&\frac{1}{(Q-3)}
\end{pmatrix}.
\end{eqnarray*}
Finally assume that
\[
\begin{array}{lr}
\begin{array}{rcl}
q_{22} &=& (\widetilde{\emptyset}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(1,1)}}),\\
q_{23} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\emptyset},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(1,1)}}),\\
q_{24} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(1,1)}}),
\end{array}
&
\begin{array}{rcl}
q_{25} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(2)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(1,1)}}),\\
q_{26} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(1,1)}}, \widehat{\mbox{\tiny\yng(1)}}, \widetilde{\mbox{\tiny\yng(1,1)}}),\\
q_{27} &=& (\widetilde{\mbox{\tiny\yng(1)}}, \widehat{\mbox{\tiny\yng(1)}},
\widetilde{\mbox{\tiny\yng(2)}}, \widehat{\mbox{\tiny\yng(1,1)}}, \widetilde{\mbox{\tiny\yng(1,1)}}).
\end{array}
\end{array}
\]
Then for the standard vectors $(v_j)_{j=22}^{27}$
which correspond to $(q_j)_{j=22}^{27}$
we define $\rho_{\mbox{\boldmath $\alpha$}}(s_2)(v_{22}\ v_{23}\ v_{24}\ v_{25}\ v_{26}\ v_{27}
)$ by
\[
\rho_{\mbox{\boldmath $\alpha$}}(s_2)(v_{22}\ v_{23}\ v_{24}\ v_{25}\ v_{26}\ v_{27})
= (v_{22}\ v_{23}\ v_{24}\ v_{25}\ v_{26}\ v_{27})M_i.
\]
Here the matrix $M_i$ is
\begin{eqnarray*}
\begin{pmatrix}
-1
&0
&0
&0
&0
&0\\
\noalign{\medskip}
0
&\frac{1}{(Q-1)}
&\frac{-1}{(Q-1)(Q-2)}
&\frac{-Q(Q-3)}{2(Q-1)( Q-2 )}
&1/2
&0\\
\noalign{\medskip}
0
&\frac{-1}{(Q-1)}
&\frac{1}{(Q-1)(Q-2)}
&\frac{Q(Q-3)}{2(Q-1)(Q-2)}
&1/2
&1\\
\noalign{\medskip}
0
&-1
&\frac{1}{(Q-2)}
&\frac{Q-4}{2(Q-2)}
&1/2
&\frac{-1}{(Q-3)}
\\
\noalign{\medskip}
0
&1
&\frac{1}{(Q-2)}
&\frac{Q(Q-3)}{2(Q-1)(Q-2)}
&\frac{Q-3}{2(Q-1)}
&\frac{-1}{(Q-1)}\\
\noalign{\medskip}
0
&0
&\frac{Q-3}{Q-2}
&\frac{-Q(Q-3)}{2(Q-1)(Q-2)}
&\frac{-(Q-3)}{2(Q-1)}
&\frac{1}{(Q-1)}
\end{pmatrix}.
\end{eqnarray*}
\section{Discussion}\label{sec:dec}
In the previous section, we gave linear maps
$\rho_{\mbox{\boldmath $\alpha$}}(e_i)$ and
$\rho_{\mbox{\boldmath $\alpha$}}(f_i)$
for all the tableaux on $\Gamma_n$.
and defined
$\rho_{\mbox{\boldmath $\alpha$}}(s_i)$ for non-reductive tableaux on $\Gamma_n$.
We also defined $\rho_{\mbox{\boldmath $\alpha$}}(s_1)$
and $\rho_{\mbox{\boldmath $\alpha$}}(s_2)$ for the reductive tableaux on $\Gamma_n$.
(So far, we have further obtained $\rho_{\mbox{\boldmath $\alpha$}}(s_3)$
for almost all reductive tableaux on $\Gamma_4$.)
These linear maps preserve the relations
in Theorem~1.2 and Theorem~5.3.
Hence they give representations of $A_n(Q)$ for
all $\mbox{\boldmath $\alpha$}\in\mbox{\boldmath $\Lambda$}_n$
($n = 2-\frac{1}{2}, 2, 3-\frac{1}{2}, 3, 4-\frac{1}{2}$)
and for almost all $\mbox{\boldmath $\alpha$}\in\mbox{\boldmath $\Lambda$}_4$.
These representations also coincide with the ones calculated
through the Murphy's operators which are introduced
in the paper~\cite{HR} and programmed by Naruse.
Moreover, the traces of the representation matrices above
coincide with the ``characters'' which is defined
by Naruse in the paper~\cite{Na1}.
This means that
the representations we have presented in this note
will be irreducible and define Young's seminormal form
representations of the partition algebras $A_n(Q)$.
|
\section{Introduction}\label{sec:introduction}
The question of how planets form is one of the key questions in modern astronomy today. While it has been studied for centuries, the problem is still far from being solved. The agglomeration of small dust particles to larger ones is believed to be at least the first stage of planet formation. Both laboratory experiments \citep{Blum:2000p8110} and observations of the 10~$\mu$m spectral region \citep{Bouwman:2001p8118,vanBoekel:2003p8117} conclude that grain growth must take place in circumstellar disks. The growth from sub-micron size particles to many thousand kilometer sized planets covers 13 orders of magnitude in spatial scale and over 40 orders of magnitude in mass. To assemble a single 1 km diameter planetesimal requires the agglomeration of about $10^{27}$ dust particles. These dynamic ranges are so phenomenal that one has to resort to special methods to study the growth of particles though aggregation in the context of planet(esimal) formation.
A commonly used method for this purpose makes use of particle size distribution functions. The time dependent evolution of these particle size distribution functions has been studied by \citet{Weidenschilling:1980p4572}, \citet{Nakagawa:1981p4533}, \citet{Dullemond:2005p378}, \citet{Brauer:2008p215} (hereafter \citetalias{Brauer:2008p215}) and others. It was concluded that dust growth by coagulation can be very quick initially (in the order of thousand years to grow to centimeter sized aggregates at 1 AU in the solar nebula), but it stalls around decimeter to meter size due to what is known as the ``meter size barrier'': a size range within which particles achieve large enough velocities to undergo destructive collisions and fast radial inward drift toward the central star \citep{Weidenschilling:1977p865,Nakagawa:1986p2048}.
While the existence of this meter size barrier (ranging in fact from a couple of centimeters to tens of meters at 1 AU) has been known for over 30~years, a thorough study of this barrier, including all known mechanisms that induce motions of dust grains in protoplanetary disks, and at all regions in the disk, for various conditions in the disk and for different properties of the dust (such as sticking efficiency and critical fragmentation velocity), has been only undertaken recently \citepalias[see][]{Brauer:2008p215}. It was concluded that the barrier is indeed a very strong limiting factor in the formation of planetesimals from dust, and that special particle trapping mechanisms are likely necessary to overcome the barrier.
However, this work was based on a static, non-evolving gas disk model. It is known that over the duration of the planet formation process the disk itself also evolves dramatically \citep{LyndenBell:1974p1945,Hartmann:1998p664,Hueso:2005p685}, which may influence the processes of dust coagulation and fragmentation. It is the goal of this paper to introduce a combined disk-evolution and dust-evolution model which also includes additional physics: we include relative azimuthal velocities, radial dependence of fragmentation critical velocities and the Stokes-drag regime for small Reynolds numbers.
The aim is to find out what the effect of disk formation and evolution is on the process of dust growth, how the initial conditions affect the final outcome, and whether certain observable signatures of the disk (for instance, its degree of dustiness at a given time) can tell us something about the physics of dust growth.
Moreover, this model will serve as a supporting model for complementary modeling efforts such as the modeling of radiative transfer in protoplanetary disks (which needs information about the dust properties for the opacities) and modeling of the chemistry in disks (which needs information about the total amount of dust surface area available for surface chemistry). In this paper we describe our model in quite some detail, and thus provide a basis for future work that will be based on this model.
Furthermore, additional physics, such photoevaporation or layered accretion can be easily included, which we aim to do in the near future.
As outlined above, this model includes already many processes which influence the evolution of the dust and the gas disk. However, there are several aspects we do not include such as back-reactions by the dust through opacity or collective effects \citet{Weidenschilling:1997p4593}, porosity effects \citep{Ormel:2007p7127}, grain charging \citep{Okuzumi:2009p7473} or the ``bouncing barrier'' (Zsom et al., in press).
This paper is organized as follows: Section~\ref{sec:model} will describe all components of the model including the radial evolution of gas and dust, as well as the temperature and vertical structure of the disk and the physics of grain growth and fragmentation. In Section~\ref{sec:results} we will compare the results of our simulations with previous steady-state disk simulations and review the aforementioned growth barriers. As an application, we show how different material properties inside and outside the snow line can cause a strong enhancement of dust within the snow line. Section~\ref{sec:discussion} summarizes our findings.
A detailed description of the numerical method along with results for selected test cases can be found in the Appendix.
\section{Model}\label{sec:model}
The model presented in this work combines a 1D viscous gas disk evolution code and a dust evolution code, taking effects of radial drift, turbulent mixing, coagulation and fragmentation of the dust into account.
We model the evolution of gas and dust in a vertically integrated way. The gas disk is viscously evolving after being built up by in-falling material from a collapsing molecular cloud core.
The radial distribution of grains is subject to gas drag, radial drift, and turbulent mixing. To which extend each effect contributes, depends on the grain/gas coupling of each grain size. By simultaneously modeling about 100--200 different grain sizes, we are able to follow the detailed evolution of the dust sub-disk being the superposition of all sizes of grain distributions.
So far, the evolution of the dust distribution depends on the evolving gas disk but not vice versa. A completely self consistently coupled code is a conceptually and numerically challenging task which will be the subject of future work.
\subsection{Evolution of gas surface density}\label{sec:gas}
Our description of the viscous evolution of the gas disk follows closely the models described by \citet{Nakamoto:1994p798} and \citet{Hueso:2005p685}. In this paper we shall therefore be relatively brief and put emphasis on differences between those models and ours.
The viscous evolution of the gas disk can be described by the continuity equation,
\begin{equation}
\frac{\partial \ensuremath{\Sigma_\mathrm{g}}\xspace }{\partial t} - \frac{1}{r}\frac{\partial}{\partial r}\left( \ensuremath{\Sigma_\mathrm{g}}\xspace \, r \, u_\mathrm{g}\right) = S_\mathrm{g},
\label{eq:ssd}
\end{equation}
where the gas radial velocity $u_\mathrm{g}$ is given by \citep[see][]{LyndenBell:1974p1945}
\begin{equation}
u_\mathrm{g} = - \frac{3}{\ensuremath{\Sigma_\mathrm{g}}\xspace\sqrt{r}} \frac{\partial}{\partial r} \left( \ensuremath{\Sigma_\mathrm{g}}\xspace \nu_\mathrm{g} \sqrt{r} \right).
\label{eq:u_gas}
\end{equation}
$\ensuremath{\Sigma_\mathrm{g}}\xspace = \int_{-\infty}^{\infty} \ensuremath{\rho_\mathrm{g}}\xspace(z) \ensuremath{\mathrm{d}} z$ is the gas surface density, $r$ the radius along the disk mid-plane and \ensuremath{\nu_\mathrm{g}}\xspace the gas viscosity. The source term on the right hand side of Eq.~\ref{eq:ssd}, denoted by $S_\mathrm{g}$ can be either infall of material onto the disk or outflow.
The molecular viscosity of the gas is too small to account for relevant accretion onto the star, the time scale of viscous evolution would be in the order of several billion years.
Observed accretion rates and disk lifetimes can only be explained if turbulent viscosity drives the evolution of circumstellar disks. Therefore \citet{Shakura:1973p4854} parameterized the unknown viscosity as the product of a velocity scale and a length scale. The largest reasonable values for these scales in the disk are the pressure scale height $\ensuremath{H_\mathrm{p}}\xspace$
\begin{equation}
\ensuremath{H_\mathrm{p}}\xspace = \frac{\ensuremath{c_\mathrm{s}}\xspace}{\ensuremath{\Omega_\mathrm{k}}\xspace}
\end{equation}
and the sound speed $\ensuremath{c_\mathrm{s}}\xspace$. Therefore the viscosity is written as
\begin{equation}
\ensuremath{\nu_\mathrm{g}}\xspace = \alpha \: \ensuremath{c_\mathrm{s}}\xspace \: \ensuremath{H_\mathrm{p}}\xspace,
\end{equation}
where $\alpha$ is the turbulence parameter and $\alpha \leq 1$.
Today it is generally believed that disks transport angular momentum by turbulence, however the source of this turbulence is still debated. The magneto-rotational instability is the most commonly accepted candidate for source of turbulence \citep{Balbus:1991p4932}. Values of $\alpha$ are expected to be in the range of $10^{-3}$ to some $10^{-2}$ \citetext{see \citealp{Johansen:2005p8425}; Dzyurkevich et al., in press}. Observations confirm this range with higher probability for larger values \citep[see][]{Andrews:2007p4967}.
If the disk becomes gravitationally unstable, large scale mechanisms of angular momentum transport such as through the formation of spiral arms come into play. The stability of the disk can be described in terms of the Toomre parameter \citep{Toomre:1964p1002}
\begin{equation}
Q = \frac{\ensuremath{c_\mathrm{s}}\xspace \ensuremath{\Omega_\mathrm{k}}\xspace}{\pi \,G \, \ensuremath{\Sigma_\mathrm{g}}\xspace}.
\end{equation}
Values below a critical value of $Q_\text{cr} = 2$ describe a weakly unstable disk, which forms non-axisymmetric instabilities like spiral arms. $Q$ values below unity lead to fragmentation of the disk.
The effect of these non-axisymmetric structures is to transport angular momentum outward and rearranging the surface density in the disk so as to counteract the unstable configuration. This mechanism is therefore to a certain extent self-limiting. Values above $Q_\text{cr}$ are not influenced by instabilities, values below $Q_\text{cr}$ form instabilities which increase $Q$ until the disk is marginally stable again. Our model approximates this mechanism by increasing the turbulence parameter $\alpha$ according to the recipe of \cite{Armitage:2001p993},
\begin{equation}
\alpha(r) = \alpha_0 + 0.01 \left( \left(\frac{Q_\text{cr}}{\min(Q(r),Q_\text{cr})}\right)^2 - 1 \right),
\label{eq:alpha_of_Q}
\end{equation}
where $\alpha_0$ is a free parameter of the model which is taken to be $10^{-3}$, unless otherwise noted.
During the time of infall onto the disk, we use a constant, high value of $\alpha = 0.1$ to mimic the effective redistribution of surface density during the infall phase which also increases the overall stability of the disk. Once the infall stops, we gradually decrease the turbulence parameter on a timescale of 10~000 years until it reaches its input value.
Eq. \ref{eq:ssd} is a diffusion equation, which is stiff. This means, one faces the problem that the numerical step of an explicit integration scheme goes $\propto \Delta r^2$ (where $\Delta r$ is the radial grid step size) which would make the simulation very slow. One possible solution to this problem is using the method of implicit integration. This scheme keeps the small time scales of diffusion i.e. the fast modes in check. We are not interested in these high frequency modes, but they would become unstable if we used a large time step. With an implicit integration scheme (see Section~\ref{app:alg_advdif}) the time step can be chosen larger without causing numerical instabilities, thus increasing the speed of the computation.
\subsection{Radial evolution of dust}\label{sec:dust}
If the average dust-to-gas ratio in protoplanetary disks is in the order of $10^{-2}$ (as found in the ISM), then the dust does not dynamically influence the gas, while the gas strongly affects the dynamics of the dust.
Thermally, however, the dust has potentially a massive influence on the gas disk evolution through its opacity, but we will not include this in this paper. Therefore the evolution of the gas disk can, in our approximation, be done without knowledge of the dust evolution, while the dust evolution itself {\em does} need knowledge of the gas evolution.
There might be regions, where dust accumulates (such as the mid-plane of the disk or dead-zones and snow-lines) and its influence becomes significant or even dominant but we will not include feedback of such dust enhancements onto the disk evolution in this paper.
For now, we want to focus on the equations of motion of dust particles under the assumption that gas is the dominant material by mass. The interplay between dust and gas can then be described in terms of a dimensionless coupling constant, the \textit{Stokes number} which is defined as
\begin{equation}
\text{St} = \frac{\tau_\text{s}}{\tau_\text{ed}},
\label{eq:ST_general}
\end{equation}
where $\tau_\text{ed}$ is the eddy turn-over time and $\tau_\text{s}$ is the stopping time.
The stopping time of a particle is defined as the ratio of the momentum of a particle divided by the drag force acting on it. There are four different regimes for the drag force which determine the dust-to-gas coupling \citep[see][]{Whipple:1972p4621,Weidenschilling:1977p865}. Which regime applies to a certain particle, depends on the ratio between mean free path $\lambda_\text{mfp}$ of the gas molecules and the dust particle size $a$ (i.e. the Knudsen number) and also on the particle Reynolds-number $Re = {2 a u/\nu_\mathrm{mol}}$ with
$\nu_\mathrm{mol}$ being the gas molecular viscosity
\begin{equation}
\nu_\mathrm{mol} = \frac{1}{2} \, \bar u \, \lambda_\text{mfp},
\end{equation}
$\bar u$ the mean thermal velocity. The mean free path is taken to be
\begin{equation}
\lambda_\text{mfp} = \frac{1}{n\:\sigma_{\mathrm{H}_2}}
\end{equation}
where $n$ denotes the mid-plane number density and $\sigma_{\mathrm{H}_2} =2 \e{-15}$~cm$^2$.
The different regimes\footnote{It should be noted that ``Stokes regime'' refers to the regime where the drag force on a particle is described by the Stokes law -- this is not directly related to the Stokes number.} are
\begin{equation}
\tau_\text{s} =
\left\{
\begin{array}{lll}
\frac{\rho_\text{s}\: a}{\rho_\text{g}\:\bar u}& \text{for } \lambda_\text{mfp}/a\gtrsim\frac{4}{9} & \text{Epstein regime}\\
\\
\frac{2 \rho_\text{s}\: a^2}{9 \nu_\mathrm{mol} \:\rho_\text{g}}& \text{for } Re<1& \text{Stokes regime 1}\\
\\
\frac{2^{0.6}\:\rho_\text{s}\:a^{1.6}}{9 \nu_\mathrm{mol}^{0.6} \: \rho_\text{g}^{1.4}\:u^{0.4}} & \text{for } 1<Re<800& \text{Stokes regime 2}\\
\\
\frac{6 \rho_\text{s}\: a}{\rho_\text{g}\:u}& \text{for } Re>800& \text{Stokes regime 3}\\
\end{array}\right.
\label{eq:stopping_time}
\end{equation}
Here $u$ denotes the velocity of the dust particle with respect to the gas, $\bar u = c_\text{s} \sqrt{{\pi}/{8}}$ denotes the mean thermal velocity of the gas molecules, $\rho_\text{s}$ is the solid density of the particles and $\rho_\text{g}$ is the local gas density.
For now, we will focus on the Epstein regime. To calculate the Stokes number, we need to know the eddy turn-over time. As noted before, our description of viscosity comes from a dimensional analysis. We use a characteristic length scale $L_\text{c}$ and a characteristic velocity scale $V_\text{c}$ of the eddies. This prescription is ambiguous in a sense that it does not specify if angular momentum is transported by large, slow moving eddies or by small, fast moving eddies, that is
\begin{equation}
\ensuremath{\nu_\mathrm{g}}\xspace = (\alpha^{q} V_\text{c}) \cdot (\alpha^{1-q} L_\text{c}).
\end{equation}
This is rather irrelevant for the viscous evolution of the gas disk, since all values of $q$ lead to the same viscosity, but if we calculate the turn-over-eddy time, we get
\begin{equation}
\tau_\text{ed} = \frac{2\pi L_\text{c}}{V_\text{c}} = \alpha^{1-2q}\: \frac{1}{\ensuremath{\Omega_\mathrm{k}}\xspace}.
\end{equation}
The Stokes number and therefore the dust-to-gas coupling as well as the
relative particle velocities strongly depend on the eddy turnover time and
therefore on $q$ . In this work $q$ is taken to be 0.5
\citep[following][]{Cuzzi:2001p2167,Schrapler:2004p2394} which leads to a
turn-over-eddy time of
\begin{equation}
\tau_\text{ed} = \frac{1}{\ensuremath{\Omega_\mathrm{k}}\xspace}.
\end{equation}
The Stokes number then becomes
\begin{equation}
\text{St} = \frac{\rho_\text{s}\cdot a}{\ensuremath{\Sigma_\mathrm{g}}\xspace} \cdot \frac{\pi}{2} \qquad \text{for } a<\frac{9}{4}\lambda_\text{mfp}.
\label{eq:ST_epstein}
\end{equation}
The overall radial movement of dust surface density $\ensuremath{\Sigma_\mathrm{d}}\xspace$ can now be described by an advection-diffusion equation,
\begin{equation}
\ddel{\ensuremath{\Sigma_\mathrm{d}}\xspace}{t} + \frac{1}{r} \ddel{}{r} \Bigl( r \, F_\text{tot} \Bigr) = 0,
\end{equation}
where the total Flux, $F_\text{tot}$ has contributions from a diffusive and an advective flux.
The diffusive part comes from the fact that the gas is turbulent and the dust couples to the gas. The dust is therefore turbulently mixed by the gas. Mixing counteracts gradients in concentration, in this case it is the dust-to-gas ratio of each size that is being smoothed out by the turbulence. The diffusive flux can therefore be written as
\begin{equation}
F_\text{diff} = - D_\text{d} \: \ddel{}{r} \left(\frac{\ensuremath{\Sigma_\mathrm{d}}\xspace}{\ensuremath{\Sigma_\mathrm{g}}\xspace}\right) \cdot \ensuremath{\Sigma_\mathrm{g}}\xspace.
\end{equation}
The ratio of gas diffusivity $D_\text{g}$ over dust diffusivity $D_\text{d}$ is called the Schmidt number. We follow \cite{Youdin:2007p2021}, who derived
\begin{equation}
\text{Sc} \equiv \frac{D_\text{g}}{D_\text{d}} = {1+\text{St}^2},
\end{equation}
and assume the gas diffusivity to be equal to the turbulent gas viscosity \ensuremath{\nu_\mathrm{g}}\xspace.
The second contribution to the dust flux is the advective flux,
\begin{equation}
F_\text{adv} = \ensuremath{\Sigma_\mathrm{d}}\xspace \cdot u_\text{r},
\end{equation}
where $u_r$ is the radial velocity of the dust. There are two contributions to it,
\begin{equation}
u_\text{r} = \frac{u_\text{g}}{1 + \text{St}^2} - \frac{2 u_\text{n}}{\text{St} + \text{St}^{-1}}.
\label{eq:u_r_dust}
\end{equation}
The first term is a drag term which comes from the radial movement of the gas which moves with a radial velocity of $u_\mathrm{g}$, given by Eq.~\ref{eq:u_gas}. Since the dust is coupled to the gas to a certain extend, the radially moving gas is able to partially drag the dust along.
The second term in Eq. \ref{eq:u_r_dust} is the radial drift velocity with respect to the gas. The gas in a Keplerian disk does in fact move sub-keplerian, since it feels the force of its own pressure gradient which is usually pointing inwards. Larger dust grains do not feel this pressure and move on a keplerian orbit. Therefore, from a point of view of a (larger) dust particle, there exists a constant head wind, which causes the particle to loose angular momentum and to drift inwards. This depends on the coupling between the gas and the particle and is described by the second term in Eq. \ref{eq:u_r_dust}.
$u_n$ denotes the maximum drift velocity of a particle,
\begin{equation}
u_\text{n} = - E_\mathrm{d} \cdot \frac{ \ddel{P_g}{r}}{2 \: \rho_\text{g} \: \ensuremath{\Omega_\mathrm{k}}\xspace},
\label{eq:u_eta}
\end{equation}
which has been derived by \citet{Weidenschilling:1977p865}. Here, we introduced a radial drift efficiency parameter $E_\mathrm{d}$. This parameter describes how efficient the radial drift actually is, as several mechanisms such as zonal or meridional flows might slow down radial drift. This will be investigated in Section~\ref{sec:drift_barrier}.
Putting all together, the time dependent equation for the surface density of one dust species of mass $m_i$ is given by
\begin{equation}
\ddel{\ensuremath{\Sigma_\mathrm{d}}\xspace^i}{t} + \frac{1}{r}\ddel{}{r}
\left\lbrace r \cdot \left[ \ensuremath{\Sigma_\mathrm{d}}\xspace^i \cdot u_r^i - D^i_\text{d} \cdot \ddel{}{r} \left( \frac{\ensuremath{\Sigma_\mathrm{d}}\xspace^i}{\ensuremath{\Sigma_\mathrm{g}}\xspace}\right) \cdot \ensuremath{\Sigma_\mathrm{g}}\xspace \right] \right\rbrace = S_\mathrm{d}^i,
\label{eq:dustequation}
\end{equation}
where we have included a source term $S^i_\mathrm{d}$ on the right hand side which can be positive in the case of infall or re-condensation of grains and negative in the case of dust evaporation or outflows.
This source term does not include the sources caused by coagulation and fragmentation processes (see Section~\ref{sec:smolu}). All sources will be combined into one equation later which is implicitly integrated in an un-split scheme (see Section~\ref{app:alg_both}).
Note that both, the diffusion coefficient and the radial velocity depend on the Stokes number and therefore on the size of the particle.
\begin{figure}[thb]
\resizebox{\hsize}{!}{\includegraphics{plots/mass_loading}}
\caption{Total amount of in-fallen surface density as function of radius according to the Shu-Ulrich infall model (see Section~\ref{sec:collapse}) assuming a centrifugal radius of 8~AU (solid line), 33~AU (dashed line), and 100~AU (dash-dotted line).}
\label{fig:mass_loading}
\end{figure}
\subsection{Temperature and vertical structure}\label{sec:temperature}
The vertical structure can be assumed as being in hydrostatic equilibrium at all times if the disk is geometrically thin ($\ensuremath{H_\mathrm{p}}\xspace(r)/r\ll r$) and the vertical sound crossing time is much shorter than the radial drift time scale of the gas.
The isothermal vertical density structure is then given by
\begin{equation}
\ensuremath{\rho_\mathrm{g}}\xspace(z) = \rho_0 \: \exp\left( - \frac{1}{2} \: \frac{z^2}{\ensuremath{H_\mathrm{p}}\xspace^2} \right),
\end{equation}
where
\begin{equation}
\rho_0 = \frac{\ensuremath{\Sigma_\mathrm{g}}\xspace}{\sqrt{2\:\pi} \: \ensuremath{H_\mathrm{p}}\xspace}.
\label{eq:rho_midplane}
\end{equation}
The viscous heating is given by \cite{Nakamoto:1994p798}
\begin{equation}
Q_+ = \ensuremath{\Sigma_\mathrm{g}}\xspace \: \ensuremath{\nu_\mathrm{g}}\xspace \left( r\: \ddel{\ensuremath{\Omega_\mathrm{k}}\xspace}{r} \right)^2,
\end{equation}
were $\ensuremath{\nu_\mathrm{g}}\xspace$ denotes the turbulent gas viscosity and $\ensuremath{\Omega_\mathrm{k}}\xspace$ the Kepler frequency.
\citet{Nakamoto:1994p798} give a solution for the mid-plane temperature with an optically thick and an optical thin contribution,
\begin{equation}
{\ensuremath{T_\mathrm{mid}}\xspace^4} = \frac{9}{8 \ensuremath{\sigma_\mathrm{B}}\xspace} \left( \frac{3}{8}{\ensuremath{\tau_\mathrm{R}}\xspace} + \frac{1}{2\: {\ensuremath{\tau_\mathrm{P}}\xspace}} \right) \ensuremath{\Sigma_\mathrm{g}}\xspace \: \alpha \: \ensuremath{c_\mathrm{s}}\xspace^2\: \ensuremath{\Omega_\mathrm{k}}\xspace + T_\text{irr}^4
\label{eq:t_mid}
\end{equation}
where we used $\ensuremath{\nu_\mathrm{g}}\xspace = \alpha\, \ensuremath{c_\mathrm{s}}\xspace\, \ensuremath{H_\mathrm{p}}\xspace$ and the approximation $\tau_\text{R/P} = \kappa_\text{R/P} \,\ensuremath{\Sigma_\mathrm{g}}\xspace/2$. \ensuremath{\kappa_\mathrm{R}}\xspace and \ensuremath{\kappa_\mathrm{P}}\xspace are Rosseland and Planck mean opacities which will be discussed in the next section.
$T_\text{irr}$ contains contributions due to stellar or external irradiation.
Here, we use a formula derived by Ruden \& Pollack \citep[see][App. B]{Ruden:1991p1806},
\begin{equation}
T_\text{irr} = T_\star \cdot \left[ \frac{2}{3 \pi} \left(\frac{R_\star}{r}\right)^3 + \frac{1}{2}\: \left(\frac{R_\star}{r}\right)^2 \: \left(\frac{\ensuremath{H_\mathrm{p}}\xspace}{r}\right) \: \left(\frac{d\ln \ensuremath{H_\mathrm{p}}\xspace}{d\ln r} - 1 \right) \right]^{\frac{1}{4}},
\end{equation}
with a fixed $\mathrm{d ln}\ensuremath{H_\mathrm{p}}\xspace/\mathrm{d ln} r = 9/7$, following \citet{Hueso:2005p685}.
The main source of opacity is the dust. Due to viscous heating, the temperature will increase with surface density. If the temperature rises above $\sim 1500$ K, the dust (i.e. the source of opacity) will evaporate, which stops the disk from further heating until all dust is vaporized. To simulate this behavior in our model, we calculate a gas temperature (assuming a small, constant value for gas opacity) in the case where the dust temperature rises above the evaporation temperature. Then $T_\text{mid}$ is approximated by
\begin{equation}
T_\text{mid} = \mathrm{max}(T_\text{gas},T_\text{evap.}),
\end{equation}
only if $T_\text{mid}$ from Eq.~\ref{eq:t_mid} would be larger than $T_\text{evap}$.
This is a thermostat effect: if $T$ rises above 1500 K, dust will evaporate, the opacity will drop and the temperature stabilizes at $T=1500$ K. Once even the very small gas opacity is enough to get temperatures $>1500$ K, all the dust is evaporated and the temperature rises further.
\subsection{Opacity}\label{sec:opacities}
In the calculation of the mid-plane temperature we use Rosseland and Planck mean opacities which are being calculated from a given frequency dependent opacity table. The results are stored in a look up table and interpolated during the calculations. The opacity table is for a mixture of 50\% silicates and 50\% carbonaceous grains.
Since these are dust opacities, we convert them from \emph{cm$^2$/g dust} to \emph{cm$^2$/g gas} by multiplying the values with the dust-to-gas ratio $\epsilon_0$, which is a fixed parameter in our model, taken to be the canonical value of 0.01. This assumes that the mean opacity of the gas is very small and that the dust-to-gas ratio does not change with time. To calculate the opacity self-consistently, the total mass of dust and the distribution of grain sizes has to be taken into account, meaning that the dust evolution has a back reaction on the gas by determining the opacity. For now, our model does not take back-reactions from dust to gas evolution into account. Only in the case where the temperature rises above 1500 K, the drop of opacity due to dust evaporation is considered, as described above.
\subsection{Initial infall phase: cloud collapse}\label{sec:collapse}
The evolution of the protoplanetary disk also depends on the initial infall phase which builds up the disk from the collapse of a cold molecular cloud core. This process is still not well understood. First similarity solutions for a collapsing sphere were calculated by \cite{Larson:1969p2574} and \cite{Penston:1969p2601}. \cite{Shu:1977p843} found a self similar solution for a singular isothermal sphere. The Larson \& Penston solution predicts much larger infall rates compared to the inside-out collapse of Shu ($\dot m_\text{in} \approx 47\: c_\text{s}^3/G$ and $0.975\: c_\text{s}^3/G$ respectively).
More recent work has shown that the infall rates are not constant over time, but develop a peak of high accretion rates and drop to smaller accretion rates at later times. The maximum accretion rate is about $13\: c_\text{s}^3/G$ if opacity effects are included \citep[see][]{Larson:2003p3025}. Analytical, pressure-free collapse calculations of \cite{Myers:2005p4950} show similar behavior but with a smaller peak accretion rate of $\dot m_\text{in} = 7.07 \ensuremath{c_\mathrm{s}}\xspace^3/G$. They also argue that the effects of pressure and magnetic fields will further increase the time scales of cloud collapse.
This initial infall phase is important since it provides the initial condition of the disk and also influences the whole simulation by providing a source of small grains and gas at larger distances to the star during later times of evolution.
It should be noted that several groups perform 3D hydrodynamic simulations of
collapsing cloud cores which show more complicated evolution
\citep[e.g.,][]{Banerjee:2006p8491,Whitehouse:2006p8532}. However, to be able to study general trends of the infall phase, we use the Shu-model since it is adjustable by a few parameters whose influences are easy to understand. In this model the collapse proceeds with an infall rate of $\dot m_\text{in}=0.975 \: c_\text{s}^3/G$ which stays constant throughout the collapse.
We assume the singular isothermal sphere of mass $M_\text{cloud}$ to be in solid body rotation $\Omega_\text{s}$. If in-falling material is conserving its angular momentum, all matter from a shell of radius $r_\text{s}$ will fall onto the star and disk system (with mass $m_\mathrm{cent}(t)$) within the so called centrifugal radius,
\begin{equation}
r_\text{centr}(t) = \frac{\Omega_\text{s}^2\: r_\text{s}^4}{G \: m_\text{cent}(t)},
\label{eq:r_centri}
\end{equation}
where $G$ is the gravitational constant and $r_\text{s} = 0.975\cdot c_\text{s}\:t /2$. The path of every parcel of gas can then be described by a ballistic orbit until it reaches the equatorial plane. The resulting flow onto the disk is
\begin{equation}
\dot\Sigma_\text{d}(r,t) = 2\;\rho_1(r,t) \cdot u_\text{z}(r,t),
\end{equation}
where
\begin{equation}
u_\text{z}(r,t) = \sqrt{\frac{G \: m_\text{cent}(t)}{r}} \cdot \mu,
\end{equation}
and
\begin{equation}
\rho_1(r,t) = \frac{\dot m_\mathrm{in}}{8 \pi \sqrt{G \: m_\text{cent}(t) \: r^3}} \cdot \frac{r}{r_\text{centr}(r,t)} \cdot \frac{1}{\mu^2}
\end{equation}
as described in \cite{Ulrich:1976p856}.
Here, $\mu$ is given by
\begin{equation}
\mu = \sqrt{1-r/r_\text{centr}(r,t)}.
\end{equation}
The centrifugal radius can therefore be approximated by (cf. \citet{Hueso:2005p685})
\begin{equation}
r_\mathrm{centri}(t) \simeq
1.4 \left( \frac{\Omega_\mathrm{s}}{10^{-14} \mathrm{~s}^{-1}} \right)^2
\left(\frac{m_\mathrm{cent}(t)}{M_\odot}\right)^3
\left(\frac{\ensuremath{c_\mathrm{s}}\xspace}{3\times 10^4 \mathrm{~cm~s}^{-1}}\right)^{-8}\mathrm{ AU}.
\end{equation}
We admit that this recipe for the formation of a protoplanetary disk is perhaps oversimplified. Firstly, as shown by \citet{Visser:2009p9087}, the geometrical thickness of the disk changes the radial distribution of in-falling matter onto the disk surface, because the finite thickness may ``capture'' an in-falling gas parcel before it could reach the midplane. Secondly, star formation is likely to be messier than our simple single-star axisymmetric infall model. And even in such a simplified scenario, the Shu inside-out collapse model is often criticized as being unrealistic. However, it would be far beyond the scope of this paper to include a better infall model. Here we just want to get a feeling for the effect of initial conditions on the dust growth, and we leave more detailed modeling to future work.
\subsection{Grain growth and fragmentation}\label{sec:growth_and_frag}
\begin{figure*}[t]
\centering
\resizebox{0.85\hsize}{!}{\includegraphics{plots/rel_vel}}
\caption{Sources of relative particle velocities considered in this model (Brownian motion velocities are not plotted) at a distance of 10~AU from the star. The turbulence parameter $\alpha$ in this simulation was $10^{-3}$. It should be noted that relative azimuthal velocities do not vanish for very large and very small particles.}
\label{fig:rel_vel}
\end{figure*}
The goal of the model described in this paper is to trace the evolution of gas and dust during the whole lifetime of a protoplanetary disk, including the grain growth, radial drift and turbulent mixing.
Here, the problem arises that radial drift and coagulation ``counteract'' each
other in the regime of $\text{St}=1$ particles: coagulation of smaller sizes restores the population around $\text{St}=1$, whereas radial drift preferentially removes these particles. To be able to properly model this behavior, the time step has to be chosen small enough if the method of operator splitting is used.
The upper limit for this time step can be very small. If larger steps are used the solution will ``flip-flop'' back and forth between the two splitted sub-steps of motion and coagulation, and the results become unreliable. A method to allow the choice of large time steps while preserving the accuracy is to use a non-splitted scheme in which the integration is done ``implicitly''. \citetalias{Brauer:2008p215} already use this technique to avoid flip-flopping between coagulation and fragmentation of grains, and we refer to that paper for a description of the general method. What is new in the current paper is that this implicit integration scheme is extended to also include the radial motion of the particles. So now radial motion, coagulation and fragmentation are done all within a single implicit integration scheme. See Appendix~\ref{app:alg_both} for details.
\subsubsection{Smoluchowski equation}\label{sec:smolu}
The dust grain distribution $n(m,r,z)$, which is a function of mass $m$, distance to the star $r$ and height above the mid-plane $z$, describes the number of particles per cm$^3$ per gram interval in particle mass. This means that the total dust density in g cm$^{-3}$ is given by
\begin{equation}
\rho(r,z)=\int_0^\infty n(m,r,z) \cdot m \; \ensuremath{\mathrm{d}}{m}.
\end{equation}
With this definition of $n(m,r,z)$, the coagulation/fragmentation at one point in the disk can be described by a general two-body process,
\begin{equation}
\begin{split}
\frac{\partial}{\partial t} n(m,r,z) =& \iint_0^\infty M(m,m',m'') \times\\
&\phantom{ \iint_0^\infty}\times n(m',r,z) \cdot n(m'',r,z) \,\ensuremath{\mathrm{d}} m' \,\ensuremath{\mathrm{d}} m'',
\end{split}
\label{eq:smolu}
\end{equation}
where $M(m,m',m'')$ is called the kernel. In the case of coagulation and fragmentation, this is given by
\begin{equation}
\begin{split}
M(m,m',m'') = &\phantom{+}\frac{1}{2} K(m',m'') \cdot \delta(m'+m''-m)\\
&- K(m',m'') \cdot \delta(m''-m)\\
&+\frac{1}{2} L(m',m'') \cdot S(m,m',m'')\\
&- L(m',m'') \cdot \delta(m-m'').
\end{split}
\label{eq:combined_kernel}
\end{equation}
For better readability, the dependency of $M$ on radius and height above the mid-plane was omitted here. The combined coagulation/fragmentation kernel consists of four terms containing the coagulation kernel $K$, the fragmentation kernel $L$ and the distribution of fragments after a collision $S$.
The first two terms in Eq. \ref{eq:combined_kernel} correspond to gain (masses $m'$ and $m-m'$ coagulate) and loss ($m$ coagulates with $m'$) due to grain growth.
The third term describes the fragmentation of masses $m$ and $m'$, governed by the fragmentation kernel $L$ and the fourth term describes the fact that when masses $m'$ and $m''$ fragment, they distribute some of their mass via fragments to smaller sizes.
The coagulation and fragmentation kernels will be described in section
\ref{sec:coag_kernel}, the distribution of fragments, $S$, will be described in
the next section.
To be able to trace the size and radial evolution of dust in a combined way, we need to express all contributing processes in terms of the same quantity. Hence, we will formulate the coagulation/fragmentation equation in a vertically integrated way. The vertical integration can be done numerically \citepalias[as in][]{Brauer:2008p215}, however coagulation processes are most important near the mid-plane, which allows to approximate the kernels as being vertically constant (using the values at the mid-plane). If the vertical dust density distribution of each grain size is taken to be a Gaussian (see Section~\ref{sec:rel_velocities}, Eq.~\ref{eq:h_dust}), then the vertical integration can be done analytically, as discussed in Appendix~\ref{app:alg_coag}. Unlike the steady-state disk models of \citetalias{Brauer:2008p215} which have fixed surface density and temperature profiles, we need to recompute the coagulation and fragmentation kernels (which are functions of surface density and temperature) at every time step. Therefore this analytical integration also saves significant amounts of computational time.
We therefore define the vertically integrated dust surface density distribution
per logarithmic bin of grain radius, $\sigma(r,a)$ as
\begin{equation}
\sigma(r,a) = \int_{-\infty}^\infty n(r,z,a) \cdot m\cdot a\; \ensuremath{\mathrm{d}}{z},
\label{eq:def_sigma}
\end{equation}
where $n(r,z,a)$ and $n(r,z,m)$ are related through $m=4\pi/3 \rho_\text{s} a^3$.
The total dust surface density at any radius is then given by
\begin{equation}
\Sigma_\mathrm{d}(r) = \int_0^\infty \sigma(r,a)\; \mathrm{dln}a.
\end{equation}
Defining $\sigma(r,a)$ as in Eq.~\ref{eq:def_sigma} makes it a grid-independent density unlike the mass integrated over each numerical bin. This way, all plots of $\sigma(r,a)$ are meaningful without knowledge of the size grid that was used. Numerically, however we use the discretized values as defined in the appendix.
In our description of growth and fragmentation of grains, we always assume the
dust particles to have a constant volume density meaning that we do not trace the evolution of porosity of the particles as this is currently computationally too expensive with a statistical code as used in this work. This can be achieved with Monte-Carlo methods as in \citet{Ormel:2007p7127} or \citet{Zsom:2008p7126}, however these models have do not yet include the radial motion of dust and therefore cannot trace the global evolution of the dust disk.
\subsubsection{Distribution of fragments}\label{sec:distr_of_fragments}
The distribution of fragments after a collision, $S(m,m',m'')$, is commonly described by a power law,
\begin{equation}
n(m) \text{d}m \propto m^{-\xi} \text{d}m.
\label{eq:frag_powerlaw}
\end{equation}
The value $\xi$ has been investigated both experimentally and theoretically.
Typical values have been found in the range between 1 and 2, by both
experimental \citep[e.g.,][]{Blum:1993p4324,Davis:1990p7995} and theoretical
studies \citep{Ormel:2009p8002}. Unless otherwise noted, we will follow \citetalias{Brauer:2008p215} by using the value of $\xi=1.83$.
In the case where masses of the colliding particles differ by orders of magnitude, a complete fragmentation of both particles is an unrealistic scenario. More likely, cratering will occur \citep{Sirono:2004p8225}, meaning that the smaller body will excavate a certain amount of mass from the larger one. The amount of mass removed from the larger one is parameterized in units of the smaller body $m_\text{s}$,
\begin{equation}
m_\text{out} = \chi \: m_\text{s}.
\end{equation}
The mass of the smaller particle plus the mass excavated from the larger one is then distributed to masses smaller than $m_\text{s}$ according to the distribution described by Eq. \ref{eq:frag_powerlaw}. In this work, we follow \citetalias{Brauer:2008p215} by assuming $\chi$ to be unity, unless otherwise noted.
\subsubsection{Coagulation and fragmentation kernels}\label{sec:coag_kernel}
The coagulation kernel $K(m_1,m_2)$ can be factorized into three parts,
\begin{equation}
K(m_1,m_2) = \Delta u(m_1,m_2) \: \sigma_\mathrm{geo}(m_1,m_2) \: p_\text{c},
\end{equation}
and, similarly, the fragmentation kernel can be written as
\begin{equation}
L(m_1,m_2) = \Delta u(m_1,m_2) \: \sigma_\mathrm{geo}(m_1,m_2) \: p_\text{f}.
\end{equation}
Here, $\Delta u(m_1,m_2)$ denotes the relative velocity of the two particles, $\sigma_\mathrm{geo}(m_1,m_2)$ is the geometrical cross section of the collision and $p_\text{c}$ and $p_\text{f}$ are the coagulation and fragmentation probabilities which sum up to unity. In general, all these factors can also depend on other material properties such as porosity, however we always assume the dust grains to have a volume density of $\rho_\text{s}=1.6$~g~cm$^{-3}$.
The fragmentation probability is still not well known and subject of both theoretical \citep{Paszun:2009p8871,Wada:2008p4903} and experimental research \citep[see][]{Blum:2008p1920,Guttler:2009p8384}. In this work, we adopt the simple recipe
\begin{equation}
p_\text{f} = \left\{
\begin{array}{ll}
0& \text{if } \Delta u < \ensuremath{u_\text{f}}\xspace - \delta u\\
\\
1& \text{if } \Delta u > \ensuremath{u_\text{f}}\xspace\\
\\
1-\frac{\ensuremath{u_\text{f}}\xspace-\Delta u}{\delta u}& \text{else}
\end{array}
\right.
\end{equation}
with a transition width $\delta u$ and the fragmentation speed \ensuremath{u_\text{f}}\xspace as free parameter which is assumed to be 1 m~s$^{-1}$, following experimental work of \citet{Blum:1993p4324} and theoretical studies of \citet{Leinhardt:2009p5282}.
\subsubsection{Relative particle velocities}\label{sec:rel_velocities}
The different sources of relative velocities considered here are Brownian motion, relative radial and azimuthal velocities, turbulent relative velocities and differential settling. These contributions will be described in the following, an example of the most important contributions is shown in Figure~\ref{fig:rel_vel}.
\textit{Brownian motion}, the thermal movement of particles, dependents on the mass of the particle. Hence, particles of different mass have an average velocity relative to each other which is given by
\begin{equation}
\Delta u_\text{BM} = \sqrt{\frac{8 k_\text{B} \: T (m_1 + m_2)}{\pi \: m_1 \: m_2 }}.
\end{equation}
Particle growth due to Brownian relative motion is most effective for small particles.
\textit{Radial drift}, as described in section \ref{sec:dust} also induces relative velocities since particles of different size are differently coupled to the gas. The relative velocity is then
\begin{equation}
\Delta u_\text{RD} = \left| u_\text{r}(m_1) - u_\text{r}(m_2) \right|,
\end{equation}
where the radial velocity of the dust, $u_\text{r}$ is given by Eq. \ref{eq:u_r_dust}.
\textit{Azimuthal relative velocities} are induced by gas drag in a similar way as radial drift. However while only particles (plus/minus 2 orders of magnitude) around \text{St}=1 are significantly drifting, relative azimuthal velocities do not vanish for encounters between very large and vary small particles (see Figure~\ref{fig:rel_vel}). Consequently, large particles are constantly suffering high velocity impacts of smaller ones. According to \citet{Weidenschilling:1977p865} and \citet{Nakagawa:1986p2048}, the relative azimuthal velocities for gas-dominated drag are
\begin{equation}
u_\varphi = \left| u_\mathrm{n} \cdot \left( \frac{1}{1+\text{St}_1^2} - \frac{1}{1+\text{St}_2^2} \right) \right|,
\label{eq:dv_az}
\end{equation}
where $u_\mathrm{n}$ is defined by Eq.~\ref{eq:u_eta}.
\textit{Turbulent motion} as source of relative velocities is discussed in detail in \cite{Ormel:2007p801}. They also derive closed form expressions for the turbulent relative velocities which we use in this work.
\textit{Differential settling} is the fifth process we consider contributing to relative particle velocities. \cite{Dullemond:2004p390} constructed detailed models of vertical disk structure describing the depletion of grains in the upper layers of the disk. They show that the equilibrium settling speed for particles in the Epstein regime is given by
$u_\text{sett} = - z \: \ensuremath{\Omega_\mathrm{k}}\xspace \: \text{St}$ which can be derived by equating the frictional force $F_\text{fric} = - m \: u / t_\text{fric}$ and the vertical component of the gravity force, $F_\text{grav} = - m \: z \: \ensuremath{\Omega_\mathrm{k}}\xspace^2$.
To limit the settling speed to velocities smaller than half the vertically projected Kepler velocity, we use
\begin{equation}
u_\text{sett} = - z \: \ensuremath{\Omega_\mathrm{k}}\xspace\:\min\left(\text{St},\frac{1}{2}\right)
\label{eq:u_settling}
\end{equation}
for calculating the relative velocities.
Since we do not resolve the detailed vertical distribution of particles, we take the scale height of each dust size as average height above the mid-plane, which gives
\begin{equation}
\Delta u_\text{DS} = \left| h_i \cdot \min(\text{St}_i,1/2) - h_j \cdot \min(\text{St}_j,1/2) \right| \: \cdot \ensuremath{\Omega_\mathrm{k}}\xspace.
\end{equation}
\begin{figure}[thb]
\centering
\resizebox{0.8\hsize}{!}{\includegraphics{plots/gas_T_snapshots}}
\caption{Evolution of disk surface density distribution (top) and midplane temperature (bottom) of the fiducial model described in
\ref{sec:visc_evol_of_gas_disk}.}
\label{fig:gas_T_snapshots}
\end{figure}
The dust scale height is calculated by equating the time scale for settling,
\begin{equation}
t_\text{sett} = \frac{z}{u_\text{sett}}
\end{equation}
and the time scale for stirring \citep{Dubrulle:1995p300,Schrapler:2004p2394,Dullemond:2004p390},
\begin{equation}
t_\text{stir} = \frac{z^2}{D_\text{d}}.
\end{equation}
By limiting the vertical settling velocity as in Eq. \ref{eq:u_settling} and by constraining the dust scale height to be at most equal to the gas scale height, one can derive the dust scale height to be
\begin{equation}
h_\text{d} = \ensuremath{H_\mathrm{p}}\xspace \cdot \min\left(1, \sqrt{\frac{\alpha}{\min(\text{St},1/2)(1+\text{St}^2)}} \right).
\label{eq:h_dust}
\end{equation}
This prescription is only accurate for the dust close to the mid-plane, however most of the dust (and hence most of the coagulation/fragmentation processes) take place near the mid-plane, therefore this approximation is accurate enough for our purposes.
\section{Results}\label{sec:results}
\subsection{Viscous evolution of the gas disk}\label{sec:visc_evol_of_gas_disk}
We will now focus on the evolution of a disk around a T Tauri like star.
We assume the rotation rate of the parent cloud core to be $7\e{-14}$~s$^{-1}$,
which corresponds to 0.06 times the break up rotation rate of the core.
The disk is being built-up from inside out due to the Shu-Ulrich
infall model, with the centrifugal radius being 8~AU.
The parameters of our fiducial model are summarized in
Table~\ref{tab:model_parameters}.
\begin{figure}[thb]
\resizebox{\hsize}{!}{\includegraphics{plots/accretion}}
\caption{Evolution of disk mass and stellar mass (solid and dashed line on left scale respectively) and accretion rate onto the star (dash-dotted line on right scale). Adapted from Figure~5 in \citet{Hueso:2005p685}.}
\label{fig:accretion}
\end{figure}
Figure~\ref{fig:gas_T_snapshots} shows how the gas surface
density and the mid-plane temperature of this model evolve as the disk gets
built up, viscously spreads and accretes onto the star. It can be seen that
viscous heating leads to a strong increase of temperature at small radii. This
effect becomes stronger as the disk surface density increases during the infall
phase. After the infall has ceased, the surface density and therefore also the
amount of viscous heating falls off.
This effect also influences the position of the dust
evaporation radius, which is assumed to be at the radius where the dust
temperature exceeds 1500~K. This position moves outwards during the infall
(because of the stronger viscous heating described above). Once the infall
stops, the evaporation radius moves back to smaller radii as the large surface
densities are being accreted onto the star.
Figure~\ref{fig:accretion} shows the evolution of accretion rate onto the star, stellar mass
and disk mass. The infall phase lasts until
about 150\,000~years. At this point, the disk looses its source of gas and
small-grained dust and the disk mass drops off quickly until the disk has
adjusted itself to the new condition. This also explains the sharp drop of the
accretion rate. The slight increase in the accretion rate afterwards comes from
the change in $\alpha$ after the infall stops (see Section~\ref{sec:gas}).
\citet{Hueso:2005p685} find a steeper, power-law decline of the accretion rate
after the end of the infall phase because their model does not take the effects
of gravitational instabilities into account.
\subsection{Fiducial model without fragmentation}\label{sec:fiducial_model_nofrag}
\begin{figure*}[htp]
\centering
\resizebox{0.9\hsize}{!}{\includegraphics{plots/snapshots_f7}\includegraphics{plots/snapshots_fid}}
\caption{Snapshots of the vertically integrated dust density distributions (defined in Eq.~\ref{eq:def_sigma}) of a steady state disk as in \citetalias{Brauer:2008p215} (left column) and of an evolving disk (fiducial model, right column). No coagulation is calculated within the evaporation radius (denoted by the dash-dotted line), fragmentation is not taken into account in both simulations. The solid line shows the particle size corresponding to a Stokes number of unity.
Since $a_\mathrm{St=1} \propto \ensuremath{\Sigma_\mathrm{g}}\xspace$ (see Eq.~\ref{eq:ST_epstein}) this curve
in fact has the same shape as $\ensuremath{\Sigma_\mathrm{g}}\xspace(r)$, so it reflects, as a ``bonus'', what the gas disk looks like. The radius dividing the evolving disk into accreting and expanding regions is marked by the dotted line and the arrows. Particles which are located below the dashed line have a positive flux in the radial direction due to coupling to the expanding gas disk and turbulent mixing (particles within the closed contour in the upper right plot have an inward pointing flux).}
\label{fig:snapshots_fid}
\end{figure*}
\begin{table}
\centering
\caption[]{Parameters of the fiducial model.}
\label{tab:model_parameters}
$
\begin{array}{p{0.5\linewidth}llp{0.1\linewidth}}
\hline
\noalign{\smallskip}
parameter & \mathrm{symbol} & \mathrm{value} & unit\\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
turbulence parameter & \alpha & 10^{-3} & -\\
irradiation angle & \varphi & 0.05 & -\\
cloud mass & M_\text{cloud} & 0.5 &$M_\odot$\\
effect. speed of sound in core & \ensuremath{c_\mathrm{s}}\xspace &3\e{4} & cm~s$^{-1}$\\
rotation rate of cloud core & \Omega_\text{s} &7\e{-14} & s$^{-1}$\\
solid density of dust grains & \rho_\text{s} & 1.6 & g/cm$^3$\\
stellar radius & R_\star & 2.5 & $R_\odot$\\
stellar temperature & T_\star & 4000 & K\\
\noalign{\smallskip}
\hline
\end{array}
$
\end{table}
We will now focus on the dust evolution of the disk. This fiducial simulation includes only grain growth without fragmentation or
erosion. All other parameters as well as the evolution of the gas surface
density and mid-plane temperature are the same as discussed in the previous section. The evolution of this model is visualized in Figure~\ref{fig:snapshots_fid}.
The contour levels in Figure~\ref{fig:snapshots_fid} show the vertically
integrated dust surface density distribution per logarithmic bin of grain
radius, $\sigma(r,a)$, as defined in Eq.~\ref{eq:def_sigma}.
The left column of Figure~\ref{fig:snapshots_fid} shows the results of dust
evolution for a steady state (i.e. not viscously evolving) gas disk as described
in \citetalias{Brauer:2008p215}.
The right column shows the evolution of the dust density distribution of the fiducial model, in which the gas disk is gradually built up through infall from the parent molecular cloud core, and the gas disk viscously spreads and accretes matter onto the star. The solid line marks the grain size corresponding to \text{St}=1 at the given radius. This grain size will be called $a_{\text{St}=1}$ hereafter. In the Epstein regime, $a_{\text{St}=1}$ is proportional to the gas surface density of the disk, which can be seen from Eq.~\ref{eq:ST_epstein}.
There are several differences to the simulations of grain growth in steady-state disks, presented in \citetalias{Brauer:2008p215}: viscously evolving disks accrete onto the star by transporting mass inwards and angular momentum outwards. This leads to the fact that some gas has to be moving to larger radii because it is 'absorbing' the angular momentum of the accreting gas. This can be seen in numerical simulations, but also in the self similar solutions of \citet{LyndenBell:1974p1945}. They show that there is a radius $R_\pm$ which divides the disk between inward and outward moving gas. This radius itself is constantly moving outwards, depending on the radial profile of the viscosity.
The radius $R_\pm$ in the fiducial model was found to move from around 20~AU at the end of the infall to 42~AU at 1~Myr and is denoted by the dotted line in Figure~\ref{fig:snapshots_fid}.
Important here is that small particles are well enough coupled to the gas to be transported along with the outward moving gas while larger particles decouple from the gas and drift inwards. Those parts of the dust distribution which lie below the dotted line in Figure~\ref{fig:snapshots_fid} have positive flux in the radial direction due to the gas-coupling.
One might expect that the dotted and dashed lines always coincide for small grains as they are well coupled to the gas, however, it can be seen that this is not the case. The reason for this is that turbulent mixing also contributes to the total flux of dust of each grain size. The smallest particles in between the dotted line and the dashed line in the lower two panels of Figure~\ref{fig:snapshots_fid} do have positive velocities, but due to a gradient in concentration of these dust particles, the flux is still negative.
During the early phases of its evolution, a disk which is built up from inside out quickly grows large particles at small radii which are lost to the star by radial drift. During further evolution, growth timescales become larger and larger (since the dust-to-gas ratio is constantly decreasing) while only the small particles are mixed out to large radii.
The radial dependence of the dust-to-gas ratio after 1~Myr is shown in Figure~\ref{fig:FGI}.
These simulations show that the initial conditions of the stationary disk models (such as shown in \citetalias{Brauer:2008p215} and in the left column of Figure~\ref{fig:snapshots_fid}) are too optimistic since they assume a constant dust-to-gas ratio at the start of their simulation throughout the disk which is not possible if the disk is being built-up from inside out unless the centrifugal radius is very large (in which case the disk would probably fragment gravitationally) and grain fragmentation is included to prevent the grains from becoming large and start drifting strongly.
Removal of larger grains and outward transport of small grains lead to the fact that the dust-to-gas ratio is reduced by 0.5--1.5 orders of magnitude compared to a stationary model. This effect is also discussed in Section~\ref{sec:influences_of_infall_model}.
\subsection{Fiducial model with fragmentation}\label{sec:fiducial_model_frag}
The situation changes significantly, if we take grain fragmentation into account. As discussed in Section~\ref{sec:distr_of_fragments}, we consider two different kinds of outcomes for grain-grain collisions with relative velocities larger than the fragmentation velocity \ensuremath{u_\text{f}}\xspace: cratering (if the masses differ by less than one order of magnitude) and complete fragmentation (otherwise).
\begin{figure*}[htb]
\centering
\resizebox{0.9\hsize}{!}{\includegraphics{plots/F_GI}}
\caption{\com{NEW FIGURE}Comparison of the radial dependence of the dust-to-gas ratio for the stationary simulations (thick lines) and the evolving disk simulations (thin lines). The four panels compare the results after 1~Myr of evolution with/without fragmentation (left/right column) and with/without effects of non-axisymmetric instabilities (top/bottom row). It can be seen that the dust-to-gas ratio of the evolving disk simulations is almost everywhere lower than the one of the stationary simulations. See Section~\ref{sec:influences_of_infall_model} for more details.}
\label{fig:FGI}
\end{figure*}
\begin{figure}[h!]
\centering
\makeatletter
\if@referee
\resizebox{0.5\hsize}{!}{\includegraphics{plots/snapshots_frag}}
\else
\resizebox{0.85\hsize}{!}{\includegraphics{plots/snapshots_frag}}
\fi
\makeatother
\caption{Evolution of the dust density distribution of the fiducial model as in Figure~\ref{fig:snapshots_fid} but with fragmentation included. The dashed contour line (in the lower two panels) around the upper end of the size distribution and around small particles at $>60$~AU marks those parts of the distribution which have a positive (outward pointing) fluxes.}
\label{fig:snapshots_frag}
\end{figure}
For particles larger than about $\text{St}\approx 10^{-3}$, relative velocities are dominated by turbulent motion (and to a lesser extend by vertical settling). Since the relative velocities increase with Stokes number (and therefore with grain size), we can calculate the approximate position of the fragmentation barrier by equating the assumed fragmentation velocity \ensuremath{u_\text{f}}\xspace with the approximate relative velocities of the particles,
\begin{equation}
\text{St}_\text{max} \simeq \frac{\ensuremath{u_\text{f}}\xspace^2}{2\alpha\,\ensuremath{c_\mathrm{s}}\xspace^2}
\label{eq:st_max}
\end{equation}
Particles larger than this size are subject to high-velocity collisions which will erode or completely fragment those particles. This is only a rough estimate as the relative velocities also depend on the size of the smaller particle and radial drift also influence the grain distribution. However Eq.~\ref{eq:st_max} reproduces well the upper end of the size distribution in most of our simulations and therefore helps understanding the influence of various parameters on the outcome of these simulations.
The evolution of the grain size distribution including fragmentation is depicted in Figure~\ref{fig:snapshots_frag}. The initial condition is quickly forgotten since particles grow on very short timescales to sizes at which they start to fragment. The resulting fragments contribute again to the growth process until a semi-equilibrium of growth and fragmentation is reached.
It can be seen that particles stay much smaller than in the model without fragmentation. This means that
they are less affected by radial drift on the one hand and better transported along with the expanding gas disk on the other hand. Consequently, considerable amounts of dust can reach radii of the order of 100~AU, as seen in Figure~\ref{fig:fid_d2g_frag}.
The approximate maximum Stokes number, defined in Eq.~\ref{eq:st_max}, is inversely proportional to the temperature (since relative velocities are proportional to \ensuremath{c_\mathrm{s}}\xspace), which means that in regions with lower temperature, particles can reach larger Stokes numbers. By equating drift and drag velocities of the particles (cf. Eq.~\ref{eq:u_r_dust}), it can be shown that the radial velocities of particles with Stokes numbers larger than about $\alpha/2$, are being dominated by radial drift.
Due to the high temperatures below $\sim$5~AU (caused by viscous heating), $\text{St}_\text{max}$ stays below this value which prevents any significant radial drift within this radius, particles inside 5~AU are therefore only transported along with the accreting gas. Particles at larger radii and lower temperatures can drift (although only slightly), which means that there is a continuous transport of dust from the outer regions into the inner regions. Once these particles arrive in the hot region, they get ``stuck'' because their Stokes number drops below $\alpha/2$. The gas within about 5~AU is therefore enriched in dust, as seen in Figure~\ref{fig:fid_d2g_frag}. The dust-to-gas ratio at 1~AU after 1~Myr is increased by 25\% over the value of in-falling matter, which is taken to be 0.01.
Figure~\ref{fig:fid_d2g_frag} also shows a relatively sharp decrease in the dust to gas ratio at a few hundred AU. At these radii, the gas densities become so small that even the smallest dust particles decouple from the gas and start to drift inwards.
The thick line in Figure~\ref{fig:fid_d2g_frag} shows as comparison the dust-to-gas ratio of the stationary disk model (cf. left column of Figure~\ref{fig:snapshots_fid}) after 1~Myr, which starts with a radially constant initial dust-to-gas ratio of 0.01.
Figure~\ref{fig:distri_slices} shows the semi-equilibrium grain surface density distribution at 1, 10 and 100~AU in the fiducial model with fragmentation at 1~Myr.
The exact shape of these distributions depends very much on the prescription of fragmentation and cratering. In general the overall shape of these semi-equilibrium distributions is always the same: a power-law or nearly constant distribution (in $\sigma$) for small grains and a peak at some grain size $a_\mathrm{max}$, beyond which the distribution drops dramatically. The peak near the upper end of the distribution is caused by cratering. This can be understood by looking at the collision velocities: the relative velocity of two particles increases with the grain size but it is lower for equal-sized collisions than for collisions with particles of very different sizes (see Figure~\ref{fig:rel_vel}). The largest particles in the distribution have relative velocities with similar sized particles which lie just below the fragmentation velocity (otherwise they would fragment). This means that the relative velocities with much smaller particles (which are too small to fragment the bigger particles but can still damage them via cratering) are above this critical velocity. This inhibits the further growth of the big particles beyond $a_\mathrm{max}$, causing a ``traffic jam'' close to the fragmentation barrier. The peak in the distribution represents this traffic jam.
\begin{figure}[tbh]
\resizebox{\hsize}{!}{\includegraphics{plots/fid_d2g_frag}}
\caption{Evolution of the radial dependence of the dust-to-gas ratio in the fiducial model including fragmentation with the times corresponding to the snapshots shown in Figure~\ref{fig:snapshots_frag}. The initial dust-to-gas ratio is taken to be~0.01. The thick dashed curve represents the result at 1~Myr of the static disk model for comparison.}
\label{fig:fid_d2g_frag}
\end{figure}
\begin{figure}[thb]
\resizebox{\hsize}{!}{\includegraphics{plots/slices}}
\caption{Vertically integrated (cf. Eq.~\ref{eq:def_sigma}) grain surface density distributions as function of grain radius at a distance of 1~AU (solid), 10~AU (dashed) and 100~AU (dot-dashed) from the star. These curves represent slices through the bottom panel of Figure~\ref{fig:snapshots_frag}.}
\label{fig:distri_slices}
\end{figure}
\subsection{Influences of the infall model}\label{sec:influences_of_infall_model}
In the fiducial model without fragmentation, continuous resupply of material by infall is the cause why the disk has much more small grains than compared to the stationary disk model (cf. Figure~\ref{fig:snapshots_fid}), which relatively quickly consumes all available micrometer sized dust. The effect has already been found in \citet{Dominik:2008p4626}: if all grains start to grow at the same time, then the bulk of the mass grows in a relatively thin peak to larger sizes (see Figure~6 in \citetalias{Brauer:2008p215}). However if the bulk of the mass already resides in particles of larger size, then additional supply of small grains by infall is not swept up effectively because of the following: firstly, the number density of large particles is small (they may be dominating the mass, but not necessarily the number density distribution) and secondly, they only reside in a thin mid-plane layer while the scale height of small particles equals the gas scale height.
We studied how much the disk evolution depends on changes in the infall model.
For a given cloud mass, the so called centrifugal radius $r_\text{centr}$, which was defined in Eq.~\ref{eq:r_centri}, depends on the temperature and the angular velocity of the cloud. Both can be varied resulting in a large range of possible centrifugal radii reaching from a few to several hundred AU. Since the centrifugal radius is the relevant parameter, we varied only the rotation rate of the cloud core. We performed simulations with three different rotation rates which correspond to centrifugal radii of about 8 (fiducial model), 33~AU, and 100~AU. For each centrifugal radius, we performed two simulations: one which includes effects of gravitational instabilities (GI) -- i.e. increased $\alpha$ during infall and according to Eq.~\ref{eq:alpha_of_Q} -- and one which neglects them.
However for a centrifugal radius of 100~AU, too much matter is loaded onto the cold outer parts of the disk and consequently, the disk would fragment through gravitational instability. We cannot treat this in our simulations, hence, for the case of 100~AU, we show only results which neglect all GI effects.
The resulting dust-to-gas ratios are being shown in Figure~\ref{fig:FGI}.
Two general aspects change in the case of higher rotation rates: firstly, more of the initial cloud mass has to be accreted onto the star by going through the outer parts of the disk. Consequently, the disk is more extended and more massive than compared to the case of low rotation rate.
Secondly, as aforementioned the high surface densities in the colder regions at larger radii cause the disk to become less gravitationally stable.
If grain fragmentation is not taken into account in the simulations, both effects cause more dust mass to be transported to larger radii. Growth and drift timescales are increasing with radius and the dust disk with centrifugal radius of 33~AU (100~AU) can stay 5 (35)~times more massive than in the low angular momentum case after 1~Myr if GI effects are neglected.
If GI effects are included, matter is even more effectively transported outward, the dust-to-disk mass ratio for 8 and 33~AU is increased from 5 to 8.
However if fragmentation is included, it does not matter so significantly, where the dust mass is deposited onto the disk since grains stay so small during the build-up phase of the disk (due to grain fragmentation by turbulent velocities) that they are well coupled to the gas. Outwards of $\sim 10$~AU (without GI effects) or of a few hundred AU (if GI effects are included), the gas densities become so small that even the smallest grains start do decouple from the gas. They are therefore not as effectively transported outwards. In these regions, the amount of dust depends on the final centrifugal radius while at smaller radii, turbulent mixing quickly evens out all differences in the dust-to-gas ratio (see left column of Figure~\ref{fig:FGI}).
It can be seen, that in all simulations, the dust-to-gas ratio is lower than in the stationary disk model. The trend in the upper right panel in Figure~\ref{fig:FGI} suggests that for a centrifugal radius of 100~AU and the enhanced radial transport by GI effects might have a higher dust-to-gas ratio than the stationary disk model. However in this case, the disk would become extremely unstable and would therefore fragment.
The reason for this is the following: to be able to compare both simulations, the total mass of the disk-star system is the same as in the stationary disk models. How the total mass is distributed onto disk and star depends on the prescription of infall. If a centrifugal radius of 100~AU is used, the disk becomes so massive that it significantly exceeds the stability criterion $M_\mathrm{disk}/M_\star \lesssim 0.1$.
\begin{figure}[h]
\centering
\makeatletter
\if@referee
\resizebox{0.7\hsize}{!}{\includegraphics{plots/snapshots_drift}}
\else
\resizebox{\hsize}{!}{\includegraphics{plots/snapshots_drift}}
\fi
\makeatother
\caption{Evolution of the dust surface density distribution of the fiducial model at 200\,000 years for different drift efficiencies $E_\text{d}$, without fragmentation. The solid line denotes the grain size $a_{\text{St}=1}$ of particles with a Stokes number of unity. Gas outside of the radius denoted by the dotted line as well as particles below the dashed line have positive radial velocities. See section~\ref{sec:fiducial_model_nofrag} for more discussion.}
\label{fig:snapshots_drift}
\end{figure}
\subsection{The radial drift barrier revisited}\label{sec:drift_barrier}
According to the current understanding of planet formation, several mechanisms seem to prevent the formation of large bodies via coagulation quite rigorously. The most famous ones -- radial drift and fragmentation -- have already been introduced above.
Radial drift has first been discussed by \citet{Weidenschilling:1977p865}, while the importance of the fragmentation barrier (which may prevent grain growth at even smaller sizes) was discussed in \citetalias{Brauer:2008p215}.
In the following, we want to test some ideas about how to weaken or to overcome these barriers apart from those already studied in \citetalias{Brauer:2008p215}.
\citetalias{Brauer:2008p215} has quantified the radial drift barrier by equating the timescales of growth and radial drift which leads to the condition
\begin{equation}
\frac{\tau_\text{g}}{\tau_\text{d}} = \frac{1}{\epsilon_0} \left( \frac{\ensuremath{H_\mathrm{p}}\xspace}{r} \right)^2 \leq \frac{1}{\gamma},
\label{eq:drift_barrier}
\end{equation}
where $\epsilon_0$ is the dust-to-gas ratio and $\tau_\text{d}$ and $\tau_\text{g}$ are the drift and growth timescales respectively. The parameter $\gamma$ describes how much more efficient growth around \text{St}=1 must be, so that the particles are not removed by radial drift. To overcome the drift barrier, obviously either particle growth must be accelerated, or the drift efficiency has to be decreased. \citetalias{Brauer:2008p215} have numerically measured the parameter $\gamma$ to be around 12. In other words, this means that the growth timescales have to be decreased (e.g. by an increased dust-to-gas ratio) until the condition in Eq.~\ref{eq:drift_barrier} is fulfilled.
However, there are other ways of breaking through the drift barrier. Firstly, the drift timescale for \text{St}=1 particles also depends on the temperature (via the pressure gradient). A simple approximation from Eq.~\ref{eq:drift_barrier} with a 0.5~$M_\odot$ star and a dust-to-gas ratio of 0.01 gives
\begin{equation}
T < 103\text{ K } \left(\frac{r}{\text{AU}}\right)^{-1},
\end{equation}
which means that particles should be able to break through the drift barrier at 1~AU if the temperature is below $\sim$100~K (or 200~K for a solar mass star).
\citet{Dullemond:2002p399} have constructed vertical structure models of passively irradiated circumstellar disks using full frequency- and angle-dependent radiative transfer. They show that the mid-plane temperature of such a T Tauri like system at 1~AU can be as low as 60~K. Reducing the temperature by some factor reduces the drift time scale by the a factor of similar size which we will call the radial drift efficiency $E_\text{d}$ (cf. Eq.~\ref{eq:u_eta}).
Zonal flows as presented in \citet{Johansen:2006p7466} could be an alternative way of decreasing the efficiency of radial drift averaged over typical time scales of particle growth. \citet{Johansen:2006p7466} found a reduction of the radial drift velocity of up to 40\% for meter-sized particles.
Meridional flows \citep[e.g.,][]{Urpin:1984p1473,Kley:1992p7134} might also seem interesting in this context, however they do not directly influence the radial drift efficiency but rather reverse the gas-drag effect. This might be important for small particles (which, however are not strongly settling to the mid-plane) but for \text{St}=1 particles, $\alpha$ would have to be extremely high to have significant influence: even $\alpha=0.1$ would result in a reduction of the particles radial velocity by approximately only a few percent.
Another possibility to weaken the drift barrier is changing its radial dependence. The reason for this is the following: particle radial drift is only a barrier if it prevents particles to cross the size $a_{\text{St}=1}$. Since particles grow while drifting, the particle size corresponding to \text{St}=1 needs to increase as well, to be a barrier. Otherwise drifting particles would grow (at least partly) through the barrier while they are drifting. If $a_{\text{St}=1}$ is decreasing in the direction toward the star, then a particle that drifts inwards would have an increasing Stokes number even if the particle does not grow at all.
\begin{figure}[thb]
\vspace{0.35cm}
\centering
\resizebox{\hsize}{!}{\includegraphics{plots/snapshots_drift_ST}}
\caption{Dust grain surface density distribution as in Figure~\ref{fig:snapshots_drift} at 200\,000 years but including the Stokes drag regime. The drift efficiency is set to $E_\text{d} = 0.75$ and fragmentation is not taken into account. It can be seen that $a_\mathrm{\text{St}=1}$ (solid line) increases with radius until about 1~AU, which facilitates the break through the drift barrier.}
\label{fig:snapshots_drift_ST}
\end{figure}
In the Epstein regime, the size corresponding to \text{St}=1 is proportional to the gas surface density
\begin{equation}
a_\text{\text{St}=1} = \frac{2 \ensuremath{\Sigma_\mathrm{g}}\xspace}{\pi \rho_\text{s}},
\end{equation}
meaning that a relatively flat profile of surface density (or even a profile with positive slope) is needed to allow particles to grow through the barrier. However, our simulations of the viscous gas disk evolution does not yield surface density profiles with positive slopes outside the dust evaporation radius.
To quantify the arguments above, we have performed simulations with varying drift efficiency $E_\text{d}$ to test how much the radial drift has to be reduced to allow break through. We have additionally included the first Stokes drag regime to see how the radial drift of particles is influenced by it.
Figure~\ref{fig:snapshots_drift} shows the grain surface density distribution after 200\,000 years of evolution for three different drag efficiencies. The most obvious changes can be seen in the region where the $a_{\text{St}=1}$ line (solid line) is relatively flat: the grain distribution is shifted towards larger Stokes numbers. As explained above, particles can grow while drifting, which can be seen in the case of $E_\text{d}=0.5$. The Stokes number and size of the largest particles is significantly increasing towards smaller radii. However the radial drift efficiency has to be reduced by 80\% to produce particles which are large enough to escape the drift regime and are therefore not lost to the star.
\begin{figure*}[thb]
\centering
\vspace{0.45cm}
\resizebox{\hsize}{!}{\includegraphics{plots/snapshots_enhancement}}
\caption{Dust surface density distributions (top row) and the according solid-to-gas ratio (bottom row) for the case of radius-dependent fragmentation velocity after 2\e{5} years of evolution. In the upper row, the vertical dashed line denotes the position of the snow line at the mid-plane, the solid line corresponds to $a_\mathrm{St=1}$ and the dot-dashed line shows the approximate location of the fragmentation barrier according to Eq.~\ref{eq:st_max}.
The snow line on the right plot lies further outside since viscous heating is stronger if $\alpha$ is larger.
In the bottom row, the solid line denotes the vertically integrated dust-to-gas ratio while the dashed line denotes the dust-to-gas ratio at the disk mid-plane.
The icy dust grains outside the snow line are assumed to fragment at a critical velocity of 10~m~s$^{-1}$ while particles inside the snow line fragment at 1~m~s$^{-1}$. The plots on the left and right hand side differ in the amount of turbulence in the disk ($\alpha = 10^{-3}$ and $10^{-2}$, respectively).}
\label{fig:snapshots_enhancement}
\end{figure*}
The situation changes, if the Stokes drag is taken into account: if gas surface densities are high enough for the dust particles to get into a different drag regime, a change in the radial dependency of $a_\text{\text{St}=1}$ can be achieved. The Epstein drag regime for particles with Stokes number of unity is only valid if
\begin{equation}
\ensuremath{\Sigma_\mathrm{g}}\xspace \lesssim 108\frac{\text{g}}{\text{cm}^2} \left(\frac{T}{\text{200 K}}\right)^{\frac{1}{4}} \left(\frac{R}{\text{AU}}\right)^{\frac{3}{4}} \left(\frac{M_\star}{M_\odot}\right)^{-\frac{1}{4}}\left(\frac{\rho_\text{s}}{1.6\frac{\text{g}}{\text{cm}^3}}\right)^{\frac{1}{2}},
\label{eq:sig_threshold}
\end{equation}
otherwise, drag forces have to be calculated according to the Stokes drag law since the Knudsen number becomes smaller than 4/9 \citep[see][]{Weidenschilling:1977p865}. The Stokes number is then given by
\begin{equation}
\text{St} = \frac{\sqrt{2\pi}}{9} \frac{\rho_\text{s}\; \sigma_{\text{H}_2} \; a^2}{\mu\; \ensuremath{m_\mathrm{p}}\xspace} \: \ensuremath{H_\mathrm{p}}\xspace^{-1},
\end{equation}
with $\sigma_{\text{H}_2}$ being the cross section of molecular hydrogen. Interestingly, $a_{\text{St}=1}$ is independent of $\ensuremath{\Sigma_\mathrm{g}}\xspace$ and proportional to the square root of the pressure scale height which decreases towards smaller radii. This means that -- as long as the surface density is high enough -- it does not depend on the radial profile of the surface density. In this regime the radial drift itself could move particles over the drift barrier since drifting inwards increases the Stokes number of a particle without increasing its size.
Results of simulations which include the Stokes drag are shown in Figure~\ref{fig:snapshots_drift_ST}. In this case, particles can already break through the drift barrier if $E_\text{d} \lesssim 0.75$. This value and the position of the breakthrough depends on where the drag law changes from Epstein to Stokes regime and therefore on the disk surface density. As noted above, larger surface densities generally shift the position of regime change towards larger radii making it easier for particles to break through the drift barrier.
It should be noted that the physical way to avoid the Stokes drag regime is to decrease the surface densities, however we chose the same initial conditions for both cases -- with and without Stokes drag -- and just neglected the Stokes drag in the latter computations to be able to compare the efficiency factors independent of other parameters such as disk mass or temperature.
\subsection{The fragmentation barrier revisited}\label{sec:frag_barrier}
In the previous section, we have shown that several mechanisms could allow particles to break through the radial drift barrier, however fragmentation puts even stronger constraints on the formation of planetesimals.
As shown by \citet{Ormel:2007p801}, the largest relative velocities are of the order of
\begin{equation}
\Delta u_\text{max} \simeq \sqrt{2 \, \alpha} \, \ensuremath{c_\mathrm{s}}\xspace.
\end{equation}
If particles should be able to break through the fragmentation barrier, then they need to survive these large relative velocities, meaning that $\Delta u_\text{max}$ has to be smaller than the fragmentation velocity of the particles, or
\begin{equation}
\frac{\ensuremath{u_\text{f}}\xspace}{\ensuremath{c_\mathrm{s}}\xspace} \gtrsim \sqrt{2 \, \alpha}.
\end{equation}
This condition is hard to fulfill with reasonable fragmentation velocities, unless $\alpha$ is very small. E.g., for $\alpha = 10^{-5}$ and a temperature of 200 to 250~K, the fragmentation velocity needs to be higher than 4~m~s$^{-1}$, which could already seen in the simulations by \cite{Brauer:2008p212}, who have simulated particle growth near the snow-line.
Even in the case of very low turbulence, relative azimuthal velocities of large ($\text{St}\gtrsim 1$) and small grains ($\text{St}\ll 1$) are of the order of 30~m~s$^{-1}$, which means that large particles are constantly being 'sand-blasted' by small grains. The only way of reducing these velocities significantly is decreasing the pressure gradient (see Equations~\ref{eq:u_eta}~and~\ref{eq:dv_az}).
Another possibility to overcome this problem would be if high-velocity impacts of smaller particles would cause net growth, as has been found experimentally by \citet{Wurm:2005p1855} and \citet{Teiser:2009p7785}.
Taken together, these facts make environments as the inner edge of dead zones ideal places for grain growth \citep[see][]{Brauer:2008p212,Kretke:2007p697}: shutting of MRI leads to low values of $\alpha$, which are needed to reduce turbulent relative velocities and the low pressure gradients prevent radial drift and high azimuthal relative velocities.
\subsection{Dust enhancement inside the snow line}\label{sec:dustenh_in_snowline}
As already noted in a previous paper \citep{Birnstiel:2009p7135}, significant loss of dust by radial drift can be prevented by assuring that particles stay small enough and are therefore not influenced by radial drift. For typical values of $\alpha$ ($10^{-3}-10^{-2}$), this means that the fragmentation velocity must be smaller than about 0.5--5~m~s$^{-1}$. If particles have higher tensile strength, they can grow to larger sizes which are again affected by radial drift.
Typical fragmentation velocities for silicate grains determined both theoretically and experimentally are of the order of a few~m~s$^{-1}$ \citep[for a review, see][]{Blum:2008p1920}. The composition of particles outside the snow-line is expected to change due to the presence of ices. This can influence material properties and increase the fragmentation velocity \citep[see][]{Schafer:2007p7468,Wada:2009p8776}.
We have performed simulations with a radially varying fragmentation speed. We assume the fragmentation speed inside the snow-line to be 1~m~s$^{-1}$, outside the snow-line to be 10~m~s$^{-1}$. It should be noted that we do not follow the abundance of water in the disk or the composition of grains, we only assume particles outside the snow line to have larger tensile strength due to the presence of ice.
To be able to compare both simulations, we used the same 1~Myr old 0.09~$M_\odot$ gas disk around a solar mass star as initial condition.
The gas surface density profile of this disk was derived by a separate run of the disk evolution code. We used this gas surface density profile and a radially constant dust-to-gas ratio as initial condition for the simulations presented in this subsection. Apart from the level of turbulence, the input for both simulations is identical, the results are therefore completely independent of uncertainties caused by the choice of the infall model.
Results of the simulations are shown in Figure~\ref{fig:snapshots_enhancement}.
A one order of magnitude higher fragmentation velocity causes the maximum grain size to be about two orders of magnitude larger, which follows from Eq.~\ref{eq:st_max} since $\text{St}_\text{max} \propto a_\text{max}$ in the Epstein regime (all particles in these simulations are small enough to be in the Epstein regime). This effect can be seen in Figure~\ref{fig:snapshots_enhancement}.
Particles outside the snow-line are therefore more strongly drifting inwards (because they reach larger Stokes numbers) where they are being pulverized as soon as they enter the region within the snow-line.
Strong drift outside the snow line and weaker radial drift inside the snow line cause the dust-to-gas ratio within the snow line to increase significantly (see bottom row of Figure~\ref{fig:snapshots_enhancement}): in the case of $\alpha = 10^{-3}$, the dust-to-gas ratio reaches values between 0.39 and 0.10 in the region from 0.2 to 1.9~AU.
Simulations for a less massive star (0.5~$M_\odot$) show the same behavior, however the increase in dust-to-gas ratio is not as high as for a solar mass star (dust-to-gas ratio of 0.27--0.20 from 0.2--4~AU).
This effect is not as significant in the case of stronger turbulence, where the maximum dust-to-gas ratio is around 0.027. The evolution of the dust-to-gas ratio at a distance of 1~AU from the star is plotted for both cases in Figure~\ref{fig:d2g_at_1AU} (the minor bump is an artifact of the initial condition).
The reason for this difference lies in the locations of the drift and fragmentation barriers. The approximate position of the fragmentation barrier is represented by the dot-dashed line in Figure~\ref{fig:snapshots_enhancement}. The radial drift barrier cannot be defined as sharply, however radial drift is strongest at a Stokes number of unity, which corresponds to the solid line in Figure~\ref{fig:snapshots_enhancement}. An increase of $\alpha$ by one order of magnitude lowers the fragmentation barrier by about one order of magnitude in grain size.
In the lower turbulence case, the fragmentation barrier lies close to $a_\mathrm{St=1}$. Most particles are therefore drifting inwards before they are large enough to experience the fragmentation barrier. Hence, the outer parts of the disk are significantly depleted in small grains.
In contrast to this case, fragmentation is the stronger barrier for grain growth throughout the disk in the high turbulence simulation (right column in Figure~\ref{fig:snapshots_enhancement}). It can be seen that particles of smaller sizes are constantly being replenished by fragmentation.
With these results in mind, the evolution of the disk mass (bottom panel of Figure~\ref{fig:dust_mass_comparison}) seems counter-intuitive: the mass of the high turbulence dust disk is decaying faster than in the low turbulence case. This can be understood by looking at the \emph{global} dust-to-gas ratio of the disks (top panel of Figure~\ref{fig:dust_mass_comparison}) which does not differ much in both cases. This means that the increased dust mass loss in the high turbulence disk is due to the underlying evolution of the gas disk. Particles in the high turbulence disk may have smaller Stokes numbers (causing drift to be less efficient), however the inward dragging by the accreting gas is stronger in this case.
To show how much the dust evolution depends on the fragmentation velocity, we included the case of a lower fragmentation velocity throughout the disk in Figure~\ref{fig:dust_mass_comparison}. It can be seen that the dust mass is retained at its initial value for much longer timescales.
\begin{figure}[thb]
\centering
\vspace{0.45cm}
\resizebox{\hsize}{!}{\includegraphics{plots/d2g_at1AU}}
\caption{Dust-to-gas ratio at a distance of 1~AU from the central star as a function of time for the case of low ($\alpha=10^{-3}$, solid line) and high ($\alpha=10^{-2}$, dashed line) turbulence.}
\label{fig:d2g_at_1AU}
\end{figure}
\begin{figure}[thb]
\centering
\vspace{0.45cm}
\resizebox{\hsize}{!}{\includegraphics{plots/dustmass_comparison}}
\caption{Evolution of the global dust-to-gas ratio (top panel) and the dust disk mass (bottom panel) for the simulations shown in Figure~\ref{fig:snapshots_enhancement}. The solid and dashed lines correspond to the low and high turbulence case, respectively. The dash-dotted line shows the evolution of the disk if a low fragmentation velocity is assumed throughout the disk.}
\label{fig:dust_mass_comparison}
\end{figure}
\section{Discussion and conclusions}\label{sec:discussion}
We constructed a new model for growth and fragmentation of dust in circumstellar disks.
We combined the (size and radial) evolution of dust of \citetalias{Brauer:2008p215} with a viscous gas evolution code which takes into account the spreading and accretion, irradiation and viscous heating of the gas disk. The dust model includes the growth/fragmentation, radial drift/drag and radial mixing of the dust. We re-implemented and substantially improved the numerical treatment of the Smoluchowski equation of \citetalias{Brauer:2008p215} to solve for the combined size and radial evolution of dust in a fully implicit, un-split scheme (see Appendix~\ref{app:alg}).
In addition to that, we also included more physics such as relative azimuthal velocities, radial dependence of fragmentation critical velocities and the Stokes drag regime.
The code has been tested extensively and was found to be very accurate and mass-conserving (see Appendix~\ref{app:tests}).
We compared our results of grain growth in evolving protoplanetary disks to those of steady state disk simulations, similar to \citetalias{Brauer:2008p215}. In spite of many differences in details, we confirm the most general result of \citetalias{Brauer:2008p215}: radial drift and particle fragmentation set strong barriers to particle growth. If fragmentation is included in the calculations, then it poses the strongest obstacle for the formation of planetesimals. Very low turbulence ($\alpha\lesssim 10^{-5}$) and fragmentation velocities of more than a few m~s$^{-1}$ are needed to be able to overcome the fragmentation barrier in the case of turbulent relative velocities.
This model includes also the initial build-up phase of the disk, which is still a very poorly understood phase of disk evolution. We use the Shu-Ulrich infall model which represents a strong simplification. However, the following novel findings of this work do not or only weakly depend on the build-up phase of the disk:
\begin{itemize}
\item Apart from an increased dust-to-gas ratio \citepalias[as in][]{Brauer:2008p215}, other mechanisms such as streaming instabilities or a decreased temperature may be able to weaken this barrier by decreasing the efficiency of radial drift. We found that in simulations without fragmentation the radial drift efficiency needs to be reduced by 80\% to produce particles which crossed the meter-size barrier and are large enough to resist radial drift.
\item If the gas surface density is above a certain threshold (in our simulations about 140~g~cm$^{-2}$ at 1~AU or 330~g~cm$^{-2}$ at 5~AU, see Eq.~\ref{eq:sig_threshold}) then the drag force which acts on the dust particles has to be calculated according to the Stokes drag law, instead of the Epstein drag law. The drift barrier in this drag regime is shifting to smaller sizes for smaller radii (independent on the radial profile of the surface density) which means that pure radial drift can already transport dust grains over the drift barrier or at least to larger Stokes numbers even without simultaneous grain growth. In this case, the drift efficiency needs to be reduced only by about 25\% to produce large bodies.
\item If relative azimuthal velocities are included, then grains with $\text{St}>1$ are constantly 'sand-blasted' by small grains (if they are present) which (in our prescription of fragmentation) causes erosion and stops grain-growth even in the case of low turbulence. Only decreasing the radial pressure gradient significantly weakens both relative azimuthal and radial velocities. Low turbulence and a small radial pressure gradient together are needed to allow larger bodies to form. These conditions may be fulfilled at the inner edge of dead zones (\citealp{Brauer:2008p212}; \citealp{Kretke:2007p697}; see also Dzyurkevich et al., in press). Future work needs to investigate the disk evolution and grain growth of disks with dead zones.
Our prescription of fragmentation and erosion may also need rethinking since \citet{Wurm:2005p1855} and \citet{Teiser:2009p7785} find net-growth by high velocity impacts of small particles onto larger bodies.
\item Higher tensile strengths of particles outside the snow-line allows particles to grow to larger sizes, which are more strongly affected by radial drift. Particles therefore drift from outside the snow-line to smaller radii where they fragment and almost stop drifting (since their radial velocity is decreased by almost two orders of magnitude). This can cause an increase the dust-to-gas ratio inside the snow-line by more than 1.5 orders of magnitude.
\item The critical fragmentation velocity and its radial dependence (and to a lesser extent the level of turbulence) is a very important parameter determining if the dust disk is drift or fragmentation dominated. A drift dominated disk is significantly depleted in small grains and only a fragmentation dominated disk can retain a significant amount of dust for millions of years as is observed in T Tauri disks.
\end{itemize}
The following results depend on the build-up phase of the disk. However unless the collapse of the parent cloud is not inside-out or so fast to cause disk fragmentation, we expect only slight alteration of the results:
\begin{itemize}
\item Disk spreading causes small particles ($\lesssim 10$~$\mu$m) to be transported outward at radii of $\sim$60--190~AU even in 1~Myr old disks.
\item Small particles provided by infalling material are not effectively swept up by large grains if the bulk of the dust mass has already grown to larger sizes.
\item In an inside-out build-up of circumstellar disks, grains growth is very fast (timescales of some 100~years) because densities are high and orbital timescales are small. Large grains are quickly lost due to drift towards the star if fragmentation is neglected. Fragmentation is firstly needed to keep grains small enough to be able to transport a significant amount of dust to large radii by disk spreading and secondly to retain it in the disk by preventing strong radial inward drift.
\end{itemize}
\begin{acknowledgements}
We like to thank Thomas Henning, Hubert Klahr, Chris Ormel, and Andras Zsom for insightful discussions. We also thank the referee, Hidekazu Tanaka for his fast and insightful review which helped to improve this paper.
\end{acknowledgements}
\bibliographystyle{aa}
|
\section{Introduction}
In this work we consider discrete stationary processes over a finite alphabet $A.$ Markov chains of finite order are widely used to model stationary processes with finite memory. A problem with full Markov chains models of finite order $M$ is that the number of parameters $(|A|^M(|A|-1))$ grows exponentially with the order $M,$ where $|A|$ denotes the cardinal of the alphabet $A.$ Another characteristic is that the class of full Markov chains is not very rich, fixed the alphabet $A$ there is just one model for each order $M$ and in practical situations could be necessary a more flexible structure in terms of number of parameters. For an extensive discussion of those two problems se \citet{Buhlmann1999}. A richer class of finite order Markov models introduced by \citet{Rissanen1983} and \citet{Buhlmann1999} are the variable length Markov chain models (VLMC) which are mentioned in section \ref{GPCT}. In the VLMC class, each model is identified by a prefix tree $\cal T$ called context tree. For a given model with a context tree $\cal T$, the final number of parameters for the model is $|{\cal T}|(| A|-1)$ and depending on the tree, this produce a parsimonious model. In \citet{Csiszar2006} is proved that the bayesian information criterion (BIC) can be used to consistently choose the VLMC model in an efficient way using the context tree weighting (CTW) algorithm.
In this paper we introduce a larger class of finite order Markov models, and we address the problem of model selection inside this class, showing that the model can be selected consistently using the BIC criterion. In our class, each model is determined by choosing a partition of the state space, our class of models include the full Markov chain models and the VLMC models because a context tree can be seen as a particular partition of the state space (see for illustration the example \ref{CTpartition}).
In Section \ref{minimalmodels}, we define the minimal Markov models and show that this models can be selected in a consistently in theorems \ref{GeralREFTEO1} and \ref{GeralREFTEO2}.
In Section \ref{algoritmos} we show two algorithms that use the results in Section \ref{minimalmodels} to choose consistently a minimal Markov model for a sample and some simulations. Section \ref{conclu} have the conclusions and Section \ref{demostra} have the proofs.
\section{Minimal Markov models}\label{minimalmodels}
\subsection{Notation} ${ }$
Let $(X_t)$ be a discrete time order $M$ Markov chain on a finite alphabet $A$. Let us call ${\cal S}=A^M$ the state space. Denote the string $a_ma_{m+1} \ldots a_{n}$ by $a_m^n,$ where $\,a_i \in A,\,m\leq i \leq n.$
Let ${\cal L}=\{ L_1,L_2,\ldots, L_K\}$ be a partition of ${\cal S},$
\begin{eqnarray}
P(L,a)=\sum_{ s \in L}\mbox{Prob}(X_{t-M}^{t-1}=s,X_t=a), \,\,\, a \in A, \,\, L \in {\cal L};
\end{eqnarray}
\begin{eqnarray}
P(L)= \sum_{ s \in L} \mbox{Prob}(X_{t-M}^{t-1}=s),\,\,\, L \in {\cal L}.
\end{eqnarray}
Let $x_1^n$ be a sample of the process $\big( X_t \big),\,s \in {\cal S},$ $a \in A$ and $n>M.$ We denote by $N_n(s,a)$ the number of occurrences of the string $s$ followed by $a$ in the sample $x_1^n,$
\begin{eqnarray} \label{contaNsa}
N_n(s,a)= \big|\{t:M< t \leq n, x_{t-M}^{t-1}=s ,x_{t}=a \}\big|,
\end{eqnarray}
the number of occurrences of $s$ in the sample $x_1^n$ is denoted by $N_n(s)$ and
\begin{eqnarray}
N_n(s)=\big|\{t:M< t \leq n, x_{t-M}^{t-1}=s \}\big|.
\end{eqnarray}
The number of occurrences of elements into $L$ followed by $a$ is given by,
\begin{eqnarray}\label{defi1}
N^{\cal L}_n(L,a)=\sum_{s \in L} N_n(s,a),\,\, L \in {\cal L};
\end{eqnarray}
the accumulated number of $N_n(s)$ for $s$ in $L$ is denoted by,
\begin{eqnarray}\label{defi2}
N^{\cal L}_n(L)=\sum_{s \in L} N_n(s),\,\, L \in {\cal L}.
\end{eqnarray}
\subsection{Good partitions of $\cal S$}\label{pbuenas} ${ }$
\begin{definition}
Let $(X_t)$ be a discrete time order $M$ Markov chain on a finite alphabet $A,$ ${\cal S}=A^M$ the state space. A partition ${\cal L}=\{ L_1,L_2,\ldots, L_K\}$ of $\cal S$ is a good partition of $\cal S$ if for each $s, s^{\prime}\,\in L, \,\,\,L \in {\cal L},$ \[Prob(X_{t}=. \,|X^{t-1}_{t-M}=s)=Prob(X_{t}=.\,|X^{t-1}_{t-M}=s^{\prime}).\]
\end{definition}
\begin{Remark} For a discrete time order $M$ Markov chain on a finite alphabet $A$ with ${\cal S}=A^M$ the state space, ${\cal L}={\cal S}$ is a good partition of ${\cal S}.$
\end{Remark}
If ${\cal L}$ is a good partition of ${\cal S},$ we define for each category $L \in {\cal L}$
\begin{eqnarray}\label{goodpartitions}
P(a \vert L)=\mbox{Prob}(X_{t}=a|X_{t-M}^{t-1}=s) \,\,\, \forall a\in A,
\end{eqnarray}
where $s$ is some element into $L.$
As a consequence, if we write $P(x_1^n)=\mbox{Prob}(X_1^n=x_1^n),$ we obtain
\begin{eqnarray}\label{probsample2}
P(x_1^n)=P(x_1^M)\prod_{L \in {\cal L}, a \in A} P(a \vert L)^{N^{\cal L}_n(L,a)}.
\end{eqnarray}
In the same way that \citet{Csiszar2006} we will define our BIC criterion using a modified maximum likelihood.
We will call maximum likelihood to the maximization of the second term in the equation (\ref{probsample2}) for the given observation. For the sequence $x_1^n,$ will be
\begin{eqnarray}\label{MLgoodp}
\mbox{ML}({\cal L},{x_1^n})=\prod_{L \in {\cal L}, a \in A} \left( \frac{r_n(L,a)}{r_n(L)} \right)^{N_n^{\cal L}(L,a)},
\end{eqnarray}
where
\begin{eqnarray}\label{erres}
r_n\big(L,a\big)=\frac{N^{{\cal L}}_n(L,a)}{n}, \,\,\,\,\,\, a \in A, \,\,\, L \in {\cal L}\,\,&and&
r_n\big(L\big)=\frac{N^{{\cal L}}_n(L)}{n}, \,\,\,\, L \in {\cal L}.
\end{eqnarray}
The BIC is given by the next definition
\begin{definition}\label{BICML}
Given a sample $x_1^n,$ of the process $(X_t),$ a discrete time order $M$ Markov chain on a finite alphabet $A$ with ${\cal S}=A^M$ the state space and ${\cal L}$ a good partition of ${\cal S}.$ The BIC of the model (\ref{MLgoodp}) is given by
\begin{eqnarray}
BIC({\cal L}, x_1^n)=\ln\left(\mbox{ML}({\cal L},x_1^n)\right)-\frac{(|{A}|-1)|{\cal L}|}{2}\ln(n). \nonumber
\end{eqnarray}
\end{definition}
\subsection{Good partitions and context trees} \label{GPCT} ${ }$
Let $(X_t)$ be a finite order Markov chain taking values on $A$ and {\sl ${\cal T}$} a set of sequences of symbols from $A$ such that no string in ${\cal T}$ is a suffix of another string in ${\cal T},$ for each $s\in \cal T,$ $d({\cal T})=\max\big(l(s), s \in {\cal T}\big)$ where $l(s)$ denote the length of the string $s,$ with $l(\emptyset)=0$ if the string is the empty string.
\begin{definition}
${\cal T}$ is a context tree for the process $(X_t)$ if for any sequence of symbols in $A$, $x^n_1$ sample of the process with $n\geq d({\cal T}),$ there exist $s \in \cal T$ such that$$Prob(X_{n+1}=a|X_1^n=x_1^n)=Prob(X_{n+1}=a|X_{n-l(s)+1}^n=s)$$
\end{definition}
$d({\cal T})$ is the depth of the tree. \\
The context tree is the minimal state space of the variable length Markov chain (VLMC), \citet{Buhlmann1999}. The context tree for a VLMC with finite depth $M$ define a good partition on the space ${\cal S}=A^M$ as illustrated by the next example.
\begin{example} \label{CTpartition} Let be a VLMC
over the alphabet $A=\{0,1\}$ with depth $M=3$ and contexts,
$$\{ 0\}, \{01\}, \{ 011\}, \{ 111\} $$
This context tree correspond to the good partition
$\{ L_1, L_2, L_3, L_4 \}$ where\\
$L_1=\{\{ 000\},\{ 100\},\{ 010\}, \{110\}\},\,L_2=\{\{001 \}, \{101\}\},\,\,L_3=\{011\}$ and $L_4=\{ 111\}. $
\end{example}
\subsection{Smaller good partitions}
\begin{definition}\label{juntarLiLj}
Let ${\cal L}^{ij}$ denote the partition
\begin{eqnarray}
{\cal L}^{ij} = \{L_1, \ldots,L_{i-1},L_{ij},L_{i+1},\ldots,L_{j-1},L_{j+1},\ldots, L_K\},\nonumber
\end{eqnarray}
where ${\cal L} = \{L_1, \ldots, L_K\}$ is a good partition of $\cal S,$ and for $1\leq i<j\leq K$ with $ L_{ij}=L_i\cup L_j.$
\end{definition}
Now we adapt the notation established for the partition ${\cal L}$ to the new partition ${\cal L}^{ij} .$\\
\begin{Notation}\label{Notationconteoij}
for $a \in A$ we write,
\begin{eqnarray}
P(L_{ij},a)&=&P(L_i,a)+P(L_j,a)\nonumber;\\
P(L_{ij})&=& P(L_i)+P(L_j) \nonumber.
\end{eqnarray}
\begin{eqnarray}\label{nlij}
N_n^{{\cal L}^{ij}}(L_{ij},a)= N^{\cal L}_n(L_i,a)+N^{\cal L}_n(L_j,a);
\end{eqnarray}
\begin{eqnarray}
N_n^{{\cal L}^{ij}}(L_{ij})= N^{\cal L}_n(L_i)+N^{\cal L}_n(L_j);
\end{eqnarray}
\end{Notation}
If $P(. \vert L_i)=P(. \vert L_j)$ then ${\cal L}^{ij}$ is a good partition and (\ref{goodpartitions}) remains valid for ${\cal L}^{ij},$ just is necessary to change ${\cal L}$ by ${\cal L}^{ij}$ in equations (\ref{probsample2}), (\ref{MLgoodp}) and definition (\ref{BICML}). \\
In the following theorem, we show that the BIC criterion provides a consistent way of detecting smaller good partition.
\begin{thm} \label{GeralREFTEO1} Let $(X_t)$ be a Markov chain with order $M$ over a finite alphabet $A,$ ${\cal S}=A^{M}$ the state space.
If ${\cal L}=\{L_1, L_2, \ldots , L_K \}$ is a good partition of ${\cal S}$ and ${ L}_i \neq { L}_j,\, L_i, L_j \in {\cal L}.$ Then, eventually almost surely as $n \to \infty,$
$$ I_{\{BIC({\cal L}^{ij},x_1^n)>BIC({\cal L },x_1^n) \}}=1$$ if, and only if $$P(a|{ L}_i)=P(a|{ L}_j) \; \forall a \in A.$$
Where $I_{A}$ is the indicator function of $A,$ and the ${\cal L}^{ij}$ partition is defined under ${\cal L}$ by equation (\ref{juntarLiLj}).
\end{thm}
Next we extract from the previous theorem the relation that we use in the next section, in practice to find smaller good partitions.
\begin{definition} Let be $(X_t)$ a Markov chain of order $M,$ with finite alphabet $A$ and state space ${\cal S}=A^M$, $x_1^n$ a sample of the process and let ${\cal L}=\{ L_1,L_2,\ldots, L_K\}$ be a good partition of ${\cal S},$
\begin{eqnarray}
d_{\cal L}(i,j)= \frac{1}{\ln(n)} \sum_{ a \in A}\left\{
N^{{\cal L}}_n(L_i,a)\ln\left(\frac{N_n(L_i,a)}{N_n(L_i)}\right)\right. + N^{{\cal L}}_n(L_j,a)\ln\left(\frac{N_n(L_j,a)}{N_n(L_j)}\right) \nonumber \\
\left. -N^{{\cal L}^{ij}}_n(L_{ij},a)\ln\left(\frac{N_n(L_{ij},a)}{N_n(L_{ij})}\right)\right\}
\end{eqnarray}
\end{definition}
\begin{corollary}\label{GeralREFEQ1} Under the assumptions of theorem \ref{GeralREFTEO1},
\begin{eqnarray}
BIC({\cal L}, x_1^n)-BIC({{\cal L}^{ij}}, x_1^n)<0 &\iff& d_{\cal L}(i,j)<\frac{(|{A}|-1)}{2}. \nonumber
\end{eqnarray}
\end{corollary}
\begin{proof} From equation (\ref{dgeral}) we have the validity of the result.
\end{proof}
\begin{Remark}The results will remain valid if we replace the constant $\frac{(|A|-1)}{2}$ for some arbitrary constant, positive and finite value $v,$ into the definition (\ref{BICML}).
\end{Remark}
\begin{Remark} \label{GOODBADTEO} Under the assumptions of theorem \ref{GeralREFTEO1},
if $P(a \vert L_i) \neq P(a \vert L_j)$ for some $a \in A,$ then
eventually almost surely as $n \to \infty,$
$ BIC({\cal L},x_1^n)>BIC({\cal L}^{ij},x_1^n)$
where ${\cal L}^{ij}$ verified the definition (\ref{juntarLiLj}).
\end{Remark}
\subsection{ Minimal good partition} ${ }$
We want to find the smaller good partition into the universe of all possible good partitions of ${\cal S}.$ This special good partition could be defined as follows and it allows the definition of the most parsimonious model into the class considered in this paper.
\begin{definition}
Let $(X_t)$ be a discrete time order $M$ Markov chain on a finite alphabet $A,$ ${\cal S}=A^M$ the state space. A partition ${\cal L}=\{ L_1,L_2,\ldots, L_K\}$ of $\cal S$ is the minimal good partition of $\cal S$ if, $\forall L \in {\cal L},$
\[s, s^{\prime} \in L\,\,\mbox{ if, and only if}\,\, Prob(X_{t}= .\, |X^{t-1}_{t-M}=s)=Prob(X_{t}= . \,|X^{t-1}_{t-M}=s^{\prime}).\]
\end{definition}
\begin{Remark} For a discrete time order $M$ Markov chain on a finite alphabet $A$ with ${\cal S}=A^M$ the state space, $\exists !$ minimal good partition of $\cal S.$
\end{Remark}
In the next example we emphasize the difference between good partitions and the minimal good partition,
The next theorem shows that for $n$ large enough we achive the partition ${\cal L}^{*}$ which is the minimal good partition.
\begin{thm} \label{GeralREFTEO2} Let $(X_t)$ be a Markov chain with order $M$ over a finite alphabet $A,$ ${\cal S}=A^{M}$ the state space and let $\cal P$ be the set of all the partitions of $\cal S.$ Define,
$$ {{\cal L}_{n}^{*}}=argmax_{\cal L \in P} \{BIC({\cal L},x_1^n)\}$$
then, eventually almost surely as $n \to \infty,$
$$ {\cal L}^{*}={{\cal L}_{n}^{*}}$$
\end{thm}
\section{Minimal good partition estimation algorithm}\label{algoritmos}
\begin{Algorithm}\label{al2}
({\it MMM} algorithm for good partitions)
\\
Consider $x_1^n$ a sample of the Markov process $(X_t),$ with order $M$ over a finite alphabet $A,$ ${\cal S}=A^{M}$ the state space. \\ Let be $ {\cal L}=\{L_1,L_2,\ldots ,L_K\}$ a good partition of ${\cal S},$ for each $s \in {\cal S},$\\
\begin{itemize}
\item[1] for $i=1,2,\cdots,K-1,$
\begin{itemize} \item[] for $j=i+1,2,\cdots,K,$
\begin{itemize} \item[] Calculate $d_{\cal L}(i,j)$
\item[] $R_n^{i,j}= I_{\{d_{\cal L}(i,j)<\frac{(|{A}|-1)}{2}\}}$
\end{itemize}
\end{itemize}
\item[2]If $R_n^{i,j}=1,$ define $L_{ij}=L_i\cup L_j$ and ${\cal L}={\cal L}^{ij}$ . Else $i=i+1,$ Return to step 1
\end{itemize}
\end{Algorithm}
The algorithm allows to define the next relation based on the sample $x_1^n,$
\begin{definition}
for $r, s \in {\cal S}; \;\; r\sim_n s \iff R_n^{a(r),a(s)}=1.$
\end{definition}
For $n$ large enough, the algorithm return the minimal good partition.
\begin{corollary} Let {\sl $\{X_t, t=0,1,2,\ldots \}$} be a Markov chain with order $M$ over a finite alphabet $A$, ${\cal S}=A^M$ and $x_1^n$ a sample of the Markov process.
$\hat{\cal L}_n, $ given by the algorithm (\ref{al2}) converges almost surely eventually to ${\cal L}^*,$ where ${\cal L}^*$ is the minimal good partition of ${\cal S}.$
\end{corollary}
\begin{proof}
Because $K <\infty,$ for $n$ large enough, the algorithm return the minimal good partition.
\end{proof}
\begin{Remark}
In the worst case, which correspond to an initial good partition equal to $\cal S$, we need to calculate the term $\left( \frac{N^{{\cal L}}_n(L,a)}{N^{{\cal L}}_n(L)} \right)^{N_n^{\cal L}(L,a)}$ for each $s \in {\cal S}$ plus $K(K-1)/2$ divisions to implement the algorithm (\ref{al2}).
\end{Remark}
The next algorithm is a variation of the first. In this case the partitions are grow selecting the pair of elements with the minimal value of $\{d_{\cal L}(i,j)$, the algorithm stop when there is not $\{d_{\cal L}(i,j)$ lower than $(|A|-1)/2$.
\begin{Algorithm}\label{almin}
Consider $x_1^n$ a sample of the Markov process $(X_t),$ with order $M$ over a finite alphabet $A,$ ${\cal S}=A^{M}$ the state space. \\ Let $ {\cal L}=\{L_1,L_2,\ldots ,L_K\}$ be a good partition of ${\cal S}$
\begin{itemize}
\item[1] Calculate $$ (i^{*},j^{*})=arg \min_{i,j | 1\leq i <j\leq K} \{ d_{ {\cal L}(i,j)} \} $$
\item[2] If $ d_{ {\cal L}(i,j)}<\frac{|A|-1}{2}$ then $ {\cal L}= {\cal L}{i^{*}j^{*}},\;\; K=K-1$ and return to 1.\\ Else end.
\end{itemize}
\end{Algorithm}
This algorithm is consistent and always return a partition but have a greater computational cost. Taking in consideration that the cost depend on $K$ and that for a Markov chain of order $M$ we consider samples of size $n$ such that $\log(n)>M$. The two algorithms \ref{al2} and \ref{almin} have a computational cost that is linear in $n$ (the sample size).
\subsection{Dendrograms and {\it MMM} algorithm } \label{dendrogamas}
In practice, when the sample size is not large enough and the algorithm \ref{al2} has not converged, it is possible that the algorithm will not return a partition of $\cal S,$ independent of the value used in $v.$ In that case, a better approach can be to use for each $r, s \in {\cal S}$ the function $d_n(r,s)$ as a similarity measure between $r$ and $s$.
Then $d_n(r,s)$ can be used to produce a dendrogram and then use the partition defined by the dendrogram as the partition estimator.
Also in practice it is possible that the maximum number of free parameters in our model is limited by a number $K$. In that case, the logic choice will be to find a value of $d$ in the dendrogram such that the size of the partition obtained cutting the dendrogram in $d$ is less or equal to $K$, the chosen model will be the one defined by that partition.
\begin{example}\label{exmodel}
Consider a Markov chain of order $M=3$ on the alphabet $A=\{0,1,2\}$ with classes:
\begin{eqnarray*}
L_1&=&\{000,100,200,010,110,210,020,120,220,022,122,222 \},\\
L_2&=&\{001,101,201,011,111,211,021,121,221 \},\\
L_3&=&\{012, 112, 212, 002 \},\\
L_4&=&\{102\},\\
L_5&=&\{202\},
\end{eqnarray*}
and transition probabilities,
\begin{eqnarray*}
P(0|L_1)=0.2, \,\,\,\,P(1|L_1)=0.3,\\
P(0|L_2)=0.4, \,\,\,\, P(1|L_2)=0.3,\\
P(0|L_3)=0.4, \,\,\,\, P(1|L_3)=0.1,\\
P(0|L_4)=0.1, \,\,\,\, P(1|L_4)=0.4,\\
P(0|L_5)=0.3, \,\,\,\, P(1|L_5)=0.5.
\end{eqnarray*}
On this example, $|A|=3$ so the penalty constant is $1=\frac{|A|-1}{2}$.
We simulated samples of sizes $n=5000$ and $9000$, obtaining dendrograms on figure \ref{dendrogramas}. The dendrogram for the sample size of $9000$ gives the correct partition.
\begin{figure}
[h!]
\label{dendrogramas}
\centering
\includegraphics[width=3in]{dendr5000.pdf}
\includegraphics[width=3in]{dendr9000.pdf}
\caption{The figure shows the dendrograms for the model on example \ref{exmodel} estimated using algorithm \ref{REFALG1} for sample sizes of $5000$ (upper picture) and $9000$ (lower picture).}
\end{figure}
\end{example}
\subsection{Simulations}\label{simulacion1}
We implemented a simulation study for the model described on example \ref{exmodel}. More precisely we simulated $1000$ samples of the process for each of the sample sizes $4000, 6000, 8000$ and $10000$. For each sample we calculate the values $d_n(r,s)$ and build the corresponding dendrogram (using the R-project package hclust with linkage method complete). Table \ref{tablaerr1} show the results.
\begin{table}
[h!]
\caption{Number of errors on the partition estimated for the model on example \ref{exmodel}}\label{tablaerr1}
\begin{center}
\begin{tabular}{c|c}
Sample size & Proportion of errors\\
\hline
4000 & 0.801\\
6000 & 0.495\\
8000 & 0.252\\
10000 & 0.161\\
\end{tabular}
\end{center}
\label{default}
\end{table}%
\subsection{Simulations}
The VLMC corresponding to the partition on example (\label{exmodel}), have contexts:
\begin{eqnarray*}
T_1&=&\{0 \},\\
T_2&=&\{1\},\\
T_3&=&\{12 \},\\
T_4&=&\{102\},\\
T_5&=&\{202\},\\
T_6&=&\{22\},\\
T_7&=&\{002\}.
\end{eqnarray*}
We simulated $1000$ samples of the process for each of the sample sizes $4000, 6000, 8000$ and $10000$. Using the tree as a basic good partition, for each sample we calculate the values $d_n(L_i,L_j)$ corresponding to the algorithm (\ref{al2}) and build the corresponding partitions. Table (\ref{tablaerr2}) show the results.
\begin{table}[htdp]
\caption{proportion of errors on the partition estimated for the model of example (\ref{exmodel})}\label{tablaerr2}
\begin{center}
\begin{tabular}{c|c}
Sample size & Proportion of errors\\
\hline
4000 & 0.614\\
6000 & 0.206\\
8000 & 0.047\\
10000 & 0.007\\
\end{tabular}
\end{center}
\label{default}
\end{table}%
Starting from the good partition corresponding to the context tree, the number of possible models is substantially reduced compared to those in the simulation on section (\ref{simulacion1}) and because of that, the error rates on this simulation are much better than before.
\section{Conclusions}\label{conclu}
Our main motivation to define the minimal Markov models is, in the first place, the concept of partitioning the state space in classes in which the states are equivalent, this allow us to model the redundancy that appears in many processes in the nature as in genetics, linguistics, etc. Each class in the state space has a very specific, clear and practical meaning: any sequence of symbol in the same class has the same effect on the future distribution of the process. In other words, they activate the same random mechanism to choose the next symbol on the process. We can think of the resulting minimal partition as a list of the relevant contexts for the process and their synonymous.
In second place our motivation for developing this methodology is to demonstrate that for a stationary, finite memory process it is theoretically possible to find consistently a minimal Markov model to represent this process and that this can be accomplished in practice. The utilitarian implication of the fact that the model selection process can be started from a context tree partition, is that minimal Markov models can be easily fitted to stationary sources where the VLMC models already works.
It is clear that there are applications on which the natural partition to estimate is neither the minimal nor a context tree partition. As long as the partition particular properties are well defined, we can use theorem \ref{GeralREFTEO1} to estimate the minimal partition satisfying those properties.
Our theorems are still valid if we change the constant term in the penalization of the BIC criterion for any positive (and finite) number.
In the case of the VLMC model, the problem of finding a better constant has been addressed in diverse works as for example \citet{Buhlmann1999} and \citet{galves2009}.
\section{Proofs} \label{demostra}
\begin{definition}\label{entropy} Let be $P$ and $Q$ probability distributions on $A.$ The relative entropy between $P$ and $Q$ is given by,
\begin{eqnarray}
D(P(\cdot) || Q(\cdot))= \sum_{a \in A} P(a) \ln \left( \frac{P(a)}{Q(a)} \right). \nonumber
\end{eqnarray}
\end{definition}
\subsection{Proof of theorem \ref{GeralREFTEO1}}
\begin{eqnarray*}
BIC({\cal L}, x_1^n)
&=&\sum_{a \in A}\ln\left(\prod_{ L \in {\cal L}} \left( \frac{r_n(L,a)}{r_n(L)} \right)^{N_n^{\cal L}(L,a)}\right)-\frac{(|{A}|-1)|{\cal L}|}{2}\ln(n),
\end{eqnarray*}
as consequence,
\begin{eqnarray}
BIC({\cal L}, x_1^n)-BIC({{\cal L}^{ij}}, x_1^n)&=& \sum_{ a \in A}\left\{
N^{{\cal L}}_n(L_i,a)\ln\left(\frac{r_n(L_i,a)}{r_n(L_i)}\right)\right. \nonumber \\
&&+ N^{{\cal L}}_n(L_j,a)\ln\left(\frac{r_n(L_j,a)}{r_n(L_j)}\right) \nonumber \\
&&-\left. N^{{\cal L}^{ij}}_n(L_{ij},a)\ln\left(\frac{r_n(L_{ij},a)}{r_n(L_{ij})}\right)\right\}-\frac{(|{A}|-1)}{2}\ln(n). \label{dgeral}
\end{eqnarray}
We note that, the condition $\,\,I_{\{BIC({\cal L}^{ij}, x_1^n)>BIC({\cal L}, x_1^n)\}}=1$ is true if, and only if
\begin{eqnarray} \sum_{ a \in A}\left\{r_n(L_i,a)\ln \left( \frac {r_n(L_i,a)}{r_n(L_i)}\right)
+r_n(L_j,a)\ln \left( \frac {r_n(L_j,a)}{r_n(L_j)}\right) \right. \nonumber \\
\left.
-r_n(L_{ij},a)\ln\left(\frac {r_n(L_{ij},a)}{r_n(L_{ij})}
\right) \right\}< \frac{(|{A}|-1)\ln(n)}{2n}.
\label{Ind1}
\end{eqnarray}
Because $r_n(L,a)$ and $r_n(L)$ are non-negative, using Jensen we have that,
\begin{eqnarray}
r_n(L_{i},a)\ln \left( \frac {r_n(L_{i},a)}{r_n(L_{i})}\right)
+r_n(L_{j},a) \ln \left( \frac {r_n(L_{j},a)}{r_n(L_{j})} \right) \geq \nonumber \\
\left(r_n(L_{i},a)+ r_n(L_{j},a) \right) \ln \left( \frac {r_n(L_{i},a)+ r_n(L_{j},a)}{r_n(L_{i})+r_n(L_{j})} \right) \nonumber
\end{eqnarray}
or equivalently,
\begin{equation}
r_n(L_{i},a)\ln \left( \frac {r_n(L_{i},a)}{r_n(L_{i})} \right)
+r_n(L_{j},a)\ln \left( \frac {r_n(L_{j},a)}{r_n(L_{j})} \right) \geq r_n(L_{ij},a)\ln \left( \frac {r_n(L_{ij},a)}{r_n(L_{ij})} \right),
\label{Jensen}
\end{equation}
with equality if and only if $\frac {r_n(L_{i},a)}{r_n(L_{i})}=\frac {r_n(L_{j},a)}{r_n(L_{j})},\,\, \forall a \in A.$\\
As consequence, equation (\ref{Jensen}) $\Rightarrow$
\begin{eqnarray}\label{Jensenconseq}
\sum_{ a \in A}\left\{r_n(L_{i},a)\ln \left( \frac {r_n(L_{i},a)}{r_n(L_{i})}
\right) +r_n(L_{j},a)\ln \left( \frac {r_n(L_{j},a)}{r_n(L_{j})}\right) \right. \nonumber \\
\left. -r_n(L_{ij},a) \ln \left( \frac {r_n(L_{ij},a)}{r_n(L_{ij})} \right)
\right\} \geq 0,
\end{eqnarray}
with equality if and only if $\frac {r_n(L_{i},a)}{r_n(L_{i})}=\frac {r_n(L_{j},a)}{r_n(L_{j})} \; \forall a \in A.$\\
Considering that $\frac{(|{A}|-1)\ln(n)}{2n} \to 0,$ as $n \to \infty$ and from the equation (\ref{Ind1}), we have that if\\
$\lim_{n \to \infty} I_{\{BIC({\cal L}^{ij}, x_1^n)>BIC({\cal L}, x_1^n)\}}=1,$ then
\begin{eqnarray*}
\lim_{n \to \infty} \sum_{ a \in A}\left\{r_n(L_{i},a)\ln \left( \frac {r_n(L_{i},a)}{r_n(L_{i})} \right)
+r_n(L_{j},a)\ln \left( \frac {r_n(L_{j},a)}{r_n(L_{j})} \right) \right.\\
\left.-r_n(L_{ij},a)\ln \left( \frac {r_n(L_{ij},a)}{r_n(L_{ij})} \right)
\right\} \leq 0,
\end{eqnarray*}
from equation (\ref{Jensenconseq}) and taking the limit inside the sum we obtain
$$ \sum_{ a \in A}\left\{P(L_{i},a)\ln \left( \frac {P(L_{i},a)}{P(L_{i})}\right)
+P(L_{j},a)\ln \left( \frac {P(L_{j},a)}{P(L_{j})} \right)
-P(L_{ij},a)\ln \left( \frac {P(L_{ij},a)}{P(L_{ij})} \right)
\right\} = 0,$$
using Jensen again, this means that
$\frac {P(L_{i},a)}{P(L_{i})}=\frac {P(L_{j},a)}{P(L_{j})} \; \forall a \in A,$
or equivalently,
$P(a|L_i)=P(a|L_j) \; \forall a \in A.$\\
For the other half of the proof, suppose that
$P(a|L_i)=P(a|L_j) \; \forall a \in A,$
as a consequence we have that
\begin{eqnarray}
P(a|L_{ij})=P(a|L_i) \; \forall a \in A\label{paipb}
\end{eqnarray}
\begin{eqnarray*}
BIC({\cal L}, x_1^n)-BIC({{\cal L}^{ij}}, x_1^n)&=&
\ln\left(\prod_{ a \in A} \left(\frac{N^{{\cal L}}_n(L_{i},a)}{N^{{\cal L}}_n(L_{i})}\right)^{N^{{\cal L}}_n(L_{i},a)}\right)\\
&+& \ln\left(\prod_{ a \in A} \left( \frac{N^{{\cal L}}_n(L_{j},a)}{N^{{\cal L}}_n(L_{j})}\right)^{N^{{\cal L}}_n(L_{j},a)}\right)\\
&-& \ln\left(\prod_{ a \in A} \left(\frac{N^{{\cal L}^{ij}}_n(L_{ij},a)}{N^{{\cal L}^{ij}}_n(L_{ij})}\right)^{N^{{\cal L}^{ij}}_n(L_{ij},a)}\right)-\frac{(|{A}|-1)}{2}\ln(n).
\label{BIC1}
\end{eqnarray*}
Now, considering that $\frac{N^{{\cal L}^{ij}}_n(L_{ij},a)}{N^{{\cal L}^{ij}}_n(L_{ij})}$ is the maximum likelihood estimator of $P(a|L_{ij})$,
$$\prod_{ a \in A} \left(\frac{N^{{\cal L}^{ij}}_n(L_{ij},a)}{N^{{\cal L}^{ij}}_n(L_{ij})}\right)^{N^{{\cal L}^{ij}}_n(L_{ij},a)} \geq \prod_{ a \in A} P(a|L_{ij})^{N^{{\cal L}^{ij}}_n(L_{ij},a)}$$
$BIC({\cal L}, x_1^n)-BIC({{\cal L}^{ij}}, x_1^n)$ is bounded above by
\begin{eqnarray*}
&\ln&\left(\prod_{ a \in A} \left(\frac{N^{{\cal L}}_n(L_{i},a)}{N^{{\cal L}}_n(L_{i})}\right)^{N^{{\cal L}}_n(L_{i},a)}\right) + \ln\left(\prod_{ a \in A} \left(\frac{N^{{\cal L}}_n(L_{j},a)}{N^{{\cal L}}_n(L_{j})}\right)^{N^{{\cal L}}_n(L_{j},a)}\right)\\
&-&\ln\left(\prod_{ a \in A} P(a|L_{ij})^{N^{{\cal L}^{ij}}_n(L_{ij},a)}\right)-\frac{(|{A}|-1)}{2}\ln(n)\\
&=&N^{{\cal L}}_n(L_{i})D\left( \frac{N^{{\cal L}}_n(L_{i},.)}{N^{{\cal L}}_n(L_{i})} \Bigg| \Bigg| P(.|L_{i})\right) +N^{{\cal L}}_n(L_{j})D\left( \frac{N^{{\cal L}}_n(L_{j},.)}{N^{{\cal L}}_n(L_{j})} \Bigg| \Bigg| P(.|L_{j}) \right) - \frac{(|{A}|-1)}{2}\ln(n).
\end{eqnarray*}
Where $D(p||q)$ is the relative entropy, given by definition (\ref{entropy}). The first equality came from (\ref{paipb}) and (\ref{nlij}). Using proposition (\ref{lemaCsiszar06}), proposition (\ref{lema2Csiszar02}), for any $\delta>0$ and $n$ large enough,
\begin{eqnarray}
D\left( \frac{N^{{\cal L}}_n(L,.)}{N^{{\cal L}}_n(L)} \Bigg| \Bigg| P(.|L)\right) &\leq& \sum_{ a \in A} \frac{\left( \frac{N^{{\cal L}}_n(L,a)}{N^{{\cal L}}_n(L)} - P(a|L)\right)^2}{P(a|L)}\\
&\leq& \sum_{ a \in A}\frac{\frac{\delta \ln(n)}{N^{{\cal L}}_n(L)}}{P(a|L)}.\label{RELENT1}
\end{eqnarray}
Then for any $\delta>0$ and $n$ large enough,
\begin{eqnarray*}
BIC({\cal L}, x_1^n)-BIC({{\cal L}^{ij}}, x_1^n) &\leq& \frac{2\delta |A|}{p}\ln(n) - \frac{(|{A}|-1)}{2}\ln(n)\\
&=&\ln(n)\left( \frac{2\delta |A|}{p}-\frac{(|{A}|-1)}{2} \right)\\
\end{eqnarray*}
where $p=\min\{P(a|L): a\in A,L\in\{ L_i,L_j\} \}.$
In particular, taking $\delta<\frac{p(|A|-1)}{4|A|}$, for $n$ large enough,
$$BIC({\cal L}, x_1^n)-BIC({{\cal L}^{ij}}, x_1^n)<0.$$
\section*{Acknowledgements} We thank Antonio Galves, Nancy Garcia, Charlotte Galves and Florencia Leonardi for their useful comments and discussions.
\section{Ordinary text}
The ends of words and sentences are marked
by spaces. It doesn't matter how many
spaces you type; one is as good as 100. The
end of a line counts as a space.
One or more blank lines denote the end
of a paragraph.
Since any number of consecutive spaces are treated like a single
one, the formatting of the input file makes no difference to
\TeX,
but it makes a difference to you.
When you use
\LaTeX,
making your input file as easy to read as possible
will be a great help as you write your document and when you
change it. This sample file shows how you can add comments to
your own input file.
Because printing is different from typewriting, there are a
number of things that you have to do differently when preparing
an input file than if you were just typing the document directly.
Quotation marks like
``this''
have to be handled specially, as do quotes within quotes:
``\,`this'
is what I just
wrote, not `that'\,''.
Dashes come in three sizes: an
intra-word
dash, a medium dash for number ranges like
1--2,
and a punctuation
dash---like
this.
A sentence-ending space should be larger than the space between words
within a sentence. You sometimes have to type special commands in
conjunction with punctuation characters to get this right, as in the
following sentence.
Gnats, gnus, etc.\
all begin with G\@.
You should check the spaces after periods when reading your output to
make sure you haven't forgotten any special cases.
Generating an ellipsis
\ldots\
%
with the right spacing around the periods
requires a special command.
\TeX\ interprets some common characters as commands, so you must type
special commands to generate them. These characters include the
following:
\& \% \# \{ and~\}.
In printing, text is emphasized by using an
{\em italic\/}
type style.
\begin{em}
A long segment of text can also be emphasized in this way. Text within
such a segment given additional emphasis
with\/ {\em Roman}
type. Italic type loses its ability to emphasize and become simply
distracting when used excessively.
\end{em}
It is sometimes necessary to prevent \TeX\ from breaking a line where
it might otherwise do so. This may be at a space, as between the
``Mr.'' and ``Jones'' in
``Mr.~Jones'',
or within a word---especially when the word is a symbol like
\mbox{\em itemnum\/}
that makes little sense when hyphenated across
lines.
\TeX\ is good at typesetting mathematical formulas like
\( x-3y = 7 \)
or
\( a_{1} > x^{2n} / y^{2n} > x' \).
Remember that a letter like
$x$
is a formula when it denotes a mathematical symbol, and should
be treated as one.
\section{Notes}
Footnotes\footnote{This is an example of a footnote.}
pose no problem\footnote{And another one}.
\section{Displayed text}
Text is displayed by indenting it from the left margin.
Quotations are commonly displayed. There are short quotations
\begin{quote}
This is a short a quotation. It consists of a
single paragraph of text. There is no paragraph
indentation.
\end{quote}
and longer ones.
\begin{quotation}
This is a longer quotation. It consists of two paragraphs
of text. The beginning of each paragraph is indicated
by an extra indentation.
This is the second paragraph of the quotation. It is just
as dull as the first paragraph.
\end{quotation}
Another frequently-displayed structure is a list.
The following is an example of an {\em itemized} list, four levels deep.
\begin{itemize}
\item This is the first item of an itemized list. Each item
in the list is marked with a ``tick''. The document
style determines what kind of tick mark is used.
\item This is the second item of the list. It contains another
list nested inside it. The three inner lists are an {\em itemized}
list.
\begin{itemize}
\item This is the first item of an enumerated list that
is nested within the itemized list.
\item This is the second item of the inner list. \LaTeX\
allows you to nest lists deeper than you really should.
\end{itemize}
This is the rest of the second item of the outer list. It
is no more interesting than any other part of the item.
\item This is the third item of the list.
\end{itemize}
The following is an example of an {\em enumerated} list, four levels deep.
\begin{enumerate}
\item This is the first item of an enumerated list. Each item
in the list is marked with a ``tick''. The document
style determines what kind of tick mark is used.
\item This is the second item of the list. It contains another
list nested inside it. The three inner lists are an {\em enumerated}
list.
\begin{enumerate}
\item This is the first item of an enumerated list that
is nested within the enumerated list.
\item This is the second item of the inner list. \LaTeX\
allows you to nest lists deeper than you really should.
\end{enumerate}
This is the rest of the second item of the outer list. It
is no more interesting than any other part of the item.
\item This is the third item of the list.
\end{enumerate}
The following is an example of a {\em description} list.
\begin{description}
\item[Cow] Highly intelligent animal that can produce milk out of grass.
\item[Horse] Less intelligent animal renowned for its legs.
\item[Human being] Not so intelligent animal that thinks that it can think.
\end{description}
You can even display poetry.
\begin{verse}
There is an environment for verse \\
Whose features some poets will curse.
For instead of making\\
Them do {\em all\/} line breaking, \\
It allows them to put too many words on a line when they'd
rather be forced to be terse.
\end{verse}
Mathematical formulas may also be displayed. A displayed formula is
one-line long; multiline formulas require special formatting
instructions.
\[ x' + y^{2} = z_{i}^{2}\]
Don't start a paragraph with a displayed equation, nor make
one a paragraph by itself.
Example of a theorem:
\begin{thm}
All conjectures are interesting, but some conjectures are more
interesting than others.
\end{thm}
\begin{proof}
Obvious.
\end{proof}
\section{Tables and figures}
Cross reference to labelled table: As you can see in Table~\ref{sphericcase} on
page~\pageref{sphericcase} and also in Table~\ref{parset} on page~\pageref{parset}.
\begin{table*}
\caption{The spherical case ($I_1=0$, $I_2=0$).}
\label{sphericcase}
\begin{tabular}{crrrrc}
\hline
Equil. \\
Points & \multicolumn{1}{c}{$x$} & \multicolumn{1}{c}{$y$} & \multicolumn{1}{c}{$z$} & \multicolumn{1}{c}{$C$} &
S \\
\hline
$~~L_1$ & $-$2.485252241 & 0.000000000 & 0.017100631 & 8.230711648 & U \\
$~~L_2$ & 0.000000000 & 0.000000000 & 3.068883732 & 0.000000000 & S \\
$~~L_3$ & 0.009869059 & 0.000000000 & 4.756386544 & $-$0.000057922 & U \\
$~~L_4$ & 0.210589855 & 0.000000000 & $-$0.007021459 & 9.440510897 & U \\
$~~L_5$ & 0.455926604 & 0.000000000 & $-$0.212446624 & 7.586126667 & U \\
$~~L_6$ & 0.667031314 & 0.000000000 & 0.529879957 & 3.497660052 & U \\
$~~L_7$ & 2.164386674 & 0.000000000 & $-$0.169308438 & 6.866562449 & U \\
$~~L_8$ & 0.560414471 & 0.421735658 & $-$0.093667445 & 9.241525367 & U \\
$~~L_9$ & 0.560414471 & $-$0.421735658 & $-$0.093667445 & 9.241525367 & U
\\
$~~L_{10}$ & 1.472523232 & 1.393484549 & $-$0.083801333 & 6.733436505 & U \\
$~~L_{11}$ & 1.472523232 & $-$1.393484549 & $-$0.083801333 & 6.733436505 & U
\\ \hline
\end{tabular}
\end{table*}
A major
point of difference lies in the value of the specific production rate $\pi$ for
large values of the specific growth rate $\mu$.
Already in the early publications \cite{r1,r2,r3}
it appeared that high glucose
concentrations in the production phase are well correlated with a
low penicillin yield (the
`glucose effect'). It has been confirmed recently
\cite{r1,r2,r3,r4}
that
high glucose concentrations inhibit the synthesis of the enzymes of the
penicillin pathway, but not the actual penicillin biosynthesis.
In other words, glucose represses (and not inhibits) the penicillin
biosynthesis.
These findings do not contradict the results of
\cite{r1} and of \cite{r4} which were obtained for
continuous culture fermentations.
Because for high values of the specific
growth rate $\mu$ it is most likely (as shall be discussed below) that
maintenance metabolism occurs, it can be shown that
in steady state continuous culture conditions, and with $\mu$ described by a Monod kinetics
\begin{equation}
C_{s} = K_{M} \frac{\mu/\mu_{x}}{1-\mu/\mu_{x}} \label{cs}
\end{equation}
Pirt \& Rhigelato determined $\pi$ for $\mu$ between
$0.023$ and $0.086$ h$^{-1}$.
They also reported a value $\mu_{x} \approx 0.095$
h$^{-1}$, so that for their experiments $\mu/\mu_{x}$ is in the range
of $0.24$ to $0.9$.
Substituting $K _M$ in Eq. (\ref{cs}) by
the value $K_{M}=1$ g/L as used by \cite{r1}, one finds
with the above equation $0.3 < C_{s} < 9$ g/L. This agrees well with
the work of \cite{r4}, who reported that penicillin biosynthesis
repression only occurs at glucose concentrations from $C_{s}=10$ g/L on.
The conclusion is that the glucose concentrations in the experiments of
Pirt \& Rhigelato probably were too low for glucose repression to be
detected. The experimental data published by Ryu \& Hospodka
are not detailed sufficiently to permit a similar analysis.
\begin{table}
\centering
\caption{Parameter sets used by Bajpai \& Reu\ss\ }\label{parset}
\begin{tabular}{lrll}
\hline
\multicolumn{2}{l}{\it parameter} & {\it Set 1} & {\it Set 2}\\
\hline
$\mu_{x}$ & [h$^{-1}$] & 0.092 & 0.11 \\
$K_{x}$ & [g/g DM] & 0.15 & 0.006 \\
$\mu_{p}$ & [g/g DM h] & 0.005 & 0.004 \\
$K_{p}$ & [g/L] & 0.0002 & 0.0001 \\
$K_{i}$ & [g/L] & 0.1 & 0.1 \\
$Y_{x/s}$ & [g DM/g] & 0.45 & 0.47 \\
$Y_{p/s}$ & [g/g] & 0.9 & 1.2 \\
$k_{h}$ & [h$^{-1}$] & 0.04 & 0.01 \\
$m_{s}$ & [g/g DM h] & 0.014 & 0.029 \\
\hline
\end{tabular}
\end{table}
Bajpai \& Reu\ss\ decided to disregard the
differences between time constants for the two regulation mechanisms
(glucose repression or inhibition) because of the
relatively very long fermentation times, and therefore proposed a Haldane
expression for $\pi$.
It is interesting that simulations with the \cite{r4} model for the
initial conditions given by these authors indicate that, when the
remaining substrate is fed at a constant rate, a considerable and
unrealistic amount of penicillin is
produced when the glucose concentration is still very high \cite{r2,r3,r4}
Simulations with the Bajpai \& Reu\ss\ model correctly predict almost
no penicillin production in similar conditions.
\begin{figure}
\vspace{6pc}
\caption[]{Pathway of the penicillin G biosynthesis.}
\label{penG}
\end{figure}
Sample of cross-reference to figure.
Figure~\ref{penG} shows that is not easy to get something on paper.
\section{Headings}
\subsection{Subsection}
Carr-Goldstein based their model on balancing methods and
biochemical know\-ledge. The original model (1980) contained an equation for the
oxygen dynamics which has been omitted in a second paper
(1981). This simplified model shall be discussed here.
\subsubsection{Subsubsection}
Carr-Goldstein
based their model on balancing methods and
biochemical know\-ledge. The original model (1980) contained an equation for the
oxygen dynamics which has been omitted in a second paper
(1981). This simplified model shall be discussed here.
\section{Equations and the like}
Two equations:
\begin{equation}
C_{s} = K_{M} \frac{\mu/\mu_{x}}{1-\mu/\mu_{x}} \label{ccs}
\end{equation}
and
\begin{equation}
G = \frac{P_{\rm opt} - P_{\rm ref}}{P_{\rm ref}} \mbox{\ }100 \mbox{\ }(\%)
\end{equation}
Two equation arrays:
\begin{eqnarray}
\frac{dS}{dt} & = & - \sigma X + s_{F} F\\
\frac{dX}{dt} & = & \mu X\\
\frac{dP}{dt} & = & \pi X - k_{h} P\\
\frac{dV}{dt} & = & F
\end{eqnarray}
and,
\begin{eqnarray}
\mu_{\rm substr} & = & \mu_{x} \frac{C_{s}}{K_{x}C_{x}+C_{s}} \\
\mu & = & \mu_{\rm substr} - Y_{x/s}(1-H(C_{s}))(m_{s}+\pi /Y_{p/s}) \\
\sigma & = & \mu_{\rm substr}/Y_{x/s}+ H(C_{s}) (m_{s}+ \pi /Y_{p/s})
\end{eqnarray}
|
\section*{Introduction}
In an unpublished work in 1992, Thomas Bier presented a surprisingly simple way to construct a large number of simplicial spheres
by using Alexander duality.
Let $\Delta$ be a simplicial complex on $[n]=\{1,2,\dots,n\}$ which is not the set of all subsets of $[n]$,
and let $\Delta^{\vee}=\{ F \subset [n]: [n] \setminus F \not \in \Delta \}$ be the Alexander dual of $\Delta$.
The \textit{Bier sphere $\bier (\Delta)$ of $\Delta$}
is defined as the deleted join of $\Delta$ and $\Delta^\vee$,
in other words, $\bier (\Delta)$ is the simplicial complex on $\{x_1,\dots,x_n,y_1,\dots,y_n\}$
defined by
$$
\bier (\Delta) =
\{ X_F \cup Y_G:
F \in \Delta,\
G \in \Delta^{\vee},\
F \cap G= \emptyset\},
$$
where $X_F=\{x_i:i \in F\}$ and $Y_G=\{y_i:i \in G\}$.
Bier proved that this simplicial complex is indeed a triangulation of a sphere,
and Bier spheres are of interest in topological combinatorics
in connection with the van Kampen-Flores theorem and non-polytopal triangulations of spheres \cite{BPSZ,Ma}.
The main purpose of this paper is to extend
Bier's construction to finite multicomplexes.
Our construction method is quite different from the original approach given by Bier.
Our approach is similar to the construction of Billera-Lee polytopes \cite{BL} and Kalai's squeezed spheres \cite{Ka}.
We first define a shellable ball $\B(M)$ from a finite multicomplex $M$,
and define the Bier sphere of $M$ as its boundary.
Here we briefly define our generalized Bier spheres.
Fix $\cc=(c_1,\dots,c_n) \in \ZZ^n_{\geq 0}$.
A monomial $x_1^{a_1}x_2^{a_2} \cdots x_n^{a_n}$, where $x_1,x_2,\dots,x_n$ are variables,
is called a \textit{$\cc$-monomial} if $a_i \leq c_i$ for all $i$.
A {\em $\cc$-multicomplex} $M$ is a non-empty set of $\cc$-monomials such that if
$u \in M$ and a monomial $v$ divides $u$ then $v \in M$.
Let
$$\widetilde X_i=\{ x_i^{(0)},x_i^{(1)},\dots,x_i^{(c_i)}\}$$
be sets of new indeterminates for $i=1,2,\dots,n$ and let $\widetilde X=\bigcup_{i=1}^n \widetilde X_i$.
We define the simplicial complex $\B_\cc(M)$ to be the simplicial complex generated by
$$\left\{ \widetilde X \setminus \{ x_1^{(a_1)},x_2^{(a_2)},\dots,x_n^{(a_n)} \}: x_1^{a_1} x_2^{a_2}\cdots x_n^{a_n} \in M \right \}.$$
In Lemma \ref{ball},
we prove that if $M$ is not the set of all $\cc$-monomials
then $\B_\cc(M)$ is a $(|\cc|-1)$-dimensional shellable ball,
where $|\cc|=c_1+ \cdots +c_n$.
We define the
\textit{Bier sphere $\bier_\cc (M)$ of a multicomplex $M$ (w.r.t.\ $\cc$)}
as the boundary of $\B_\cc(M)$.
The above definition looks very different from the construction of classical Bier spheres of simplicial complexes.
However, we show that classical Bier spheres are Bier spheres of multicomplexes w.r.t.\ $\cc=(1,1,\dots,1)$.
Note that another generalization of Bier's construction (to posets) was given by Bj\"orner et al.\ \cite{BPSZ}.
We also
prove the following combinatorial properties of $\bier_\cc(M)$
(the definition of $g$-vectors, shellability and edge decomposability will be given in Sections 1 and 4).
\begin{itemize}
\item $\bier_\cc(M)$ is shellable (Theorem \ref{shellable}).
\item The $g$-vector of $\bier_\cc(M)$ is given by
$$
g_i(\bier_\cc(M))= \# \{ u \in M : \deg u =i\} -\# \{u \in M: \deg u=|\cc|-i\},
$$
where $\# X$ denotes the cardinality of a finite set $X$ (Theorem \ref{hvector}).
\item $\bier_\cc(M)$ is edge decomposable (Theorem \ref{edgedecomposable}).
\end{itemize}
The first and second results generalize
the results of Bj\"oner et al.\ \cite[Theorems 4.1 and 5.2]{BPSZ}
who proved the same statement for classical Bier spheres of simplicial complexes.
Edge decomposability was introduced by Nevo \cite{Ne} in the study of $g$-vectors of simplicial spheres.
This property is important since if a simplicial complex is edge decomposable
then its face vector satisfies McMullen's $g$-condition.
Thus the result yields a new class of simplicial spheres whose face vectors satisfy
McMullen's $g$-condition,
since most Bier spheres are not realizable as polytopes \cite{BPSZ,Ma}.
It would be of interest to find a simple proof of this fact by using the second result
without using edge decomposability.
The construction of $\bier(M)$ is inspired by the study of shellability of multicomplexes and polarization
which are techniques in commutative algebra theory \cite{HP,So}.
Let $S=K[x_1,\dots,x_n]$ be a polynomial ring over a field $K$ with $\deg x_i=1$ for any $i$.
Let $\overline \cc =\cc + (1,1,\dots,1)$.
For a $\overline \cc$-monomial $x_1^{a_1}x_2^{a_2}\cdots x_n^{a_n}$,
its {\em polarization (w.r.t.\ $\overline \cc$)} is the squarefree monomial
$$\pol(x_1^{a_1}x_2^{a_2} \cdots x_n^{a_n})=
\prod_{a_i \ne 0} (x_{i,0} x_{i,1}\cdots x_{i,a_i-1})$$
in a polynomial ring $\widetilde S=K[x_{i,j}: 1 \leq i \leq n,\ 0 \leq j \leq c_i]$.
For a monomial ideal $I \subset S$ minimally generated by $\overline \cc$-monomials $u_1,\dots,u_s$,
the {\em polarization of $I$ (w.r.t.\ $\overline \cc$)} is the squarefree monomial ideal
$$\pol(I)=\big(\pol(u_1),\dots,\pol(u_s)\big) \subset \widetilde S.$$
Polarizations of monomial ideals are used in the study of graded Betti numbers in commutative algebra.
An advantage of polarization is that,
since taking polarizations does not change graded Betti numbers,
polarization sometimes reduce an algebraic problem on graded Betti numbers of monomial ideals
to a combinatorial problem of simplicial complexes.
Our key observation which connects polarization and Bier spheres is the following.
Let $M$ be a $\cc$-multicomplex,
and let $I(M)$ be the ideal generated by all monomials in $S$ which are not in $M$.
In Lemma \ref{jahan},
we show
$$
\pol\big(I(M)\big)=\left(\prod_{x_i^{(j)} \in F} x_{i,j}: F \subset \widetilde X,\ F \not \in \B_\cc(M)\right).$$
Thus we show that the Stanley-Reisner ideal of the Bier ball $\B_\cc(M)$
is the polarization of the monomial ideal $I(M)$.
By using the above fact,
we study algebraic aspects of Bier spheres.
Since Bier spheres of simplicial complexes are defined by using Alexander duality,
it is natural to ask if Bier spheres of multicomplexes are also related to Alexander duality.
In Section 3, we show that Bier spheres of multicomplexes can be defined by using
Alexander duality for monomial ideals, introduced by Miller \cite{Mi,Mi2}.
In Section 5, we discuss a connection between Bier spheres and linkage theory,
and study graded Betti numbers of Stanley-Reisner rings of Bier spheres.
The results of the paper show that Bier spheres and Kalai's squeezed spheres \cite{Ka} can be constructed in a similar way.
Kalai's squeezed spheres also arise from finite multicomplexes by certain operations
(see \cite[p.\ 6]{Ka} and \cite[Proposition 4.1]{Mu1}),
and give many shellable edge decomposable spheres which are not realizable as polytopes (\cite{Ka,Le,Mu2}).
It might be of interest to find a general construction of shellable spheres which includes both Bier spheres and Kalai's squeezed spheres.
\section{Bier sphere of a multicomplex}
We first recall the basics on simplicial complexes and multicomplexes.
Let $V$ be a finite set.
A simplicial complex $\Delta$ on $V$ is a collection of subsets of $V$ such that
if $F \in \Delta$ and $G \subset F$ then $G \in \Delta$
(we do not assume that $\Delta$ contains all $1$-subsets of $V$).
An element $F \in \Delta$ with $\#F=i+1$ is called an {\em ($i$-dimensional) face of $\Delta$},
and maximal faces under inclusion are called {\em facets}.
The \textit{dimension of $\Delta$} is the maximal dimension of its faces.
A simplicial complex is said to be {\em pure} if all its facets have the same cardinality.
For subsets $F_1,F_2,\dots,F_s$ of $V$,
we write $\langle F_1,F_2,\dots,F_s\rangle$ for the simplicial complex generated by $F_1,F_2,\dots,F_s$,
in other words,
$$\langle F_1,F_2,\dots,F_s\rangle=\big\{ G \subset V: G \subset F_i \mbox{ for some }i \in \{1,2,\dots,s\}\big\}.$$
\begin{definition}
A $(d-1)$-dimensional pure simplicial complex $\Delta$ is said to be {\em shellable} if there is an order $F_1,F_2,\dots,F_s$
of the facets of $\Delta$ such that
$$\langle F_1,F_2,\dots,F_{i-1} \rangle \cap \langle F_i \rangle$$
is generated by subsets of $F_i$ of cardinality $d-1$.
The order $F_1,F_2,\dots,F_s$ is called a {\em shelling of $\Delta$}.
\end{definition}
We say that a simplicial complex $\Delta$ is a {\em simplicial $d$-ball} (or {\em $d$-sphere})
if its geometric realization is homeomorphic to a $d$-ball (or $d$-sphere).
It is well-known that if a $(d-1)$-dimensional simplicial complex $\Delta$ is shellable and if any $(d-2)$-dimensional face of $\Delta$ is contained in at most two facets, then $\Delta$ is a simplicial ball or sphere (see \cite[Theorem 11.4]{Bj}).
This implies the following fact.
\begin{lemma}
\label{shellball}
Let $\Delta$ be a simplicial $(d-1)$-sphere and let $\Gamma$ be a $(d-1)$-dimensional subcomplex of $\Delta$ with $\Gamma \ne \Delta$.
If $\Gamma$ is shellable
then $\Gamma$ is a simplicial $(d-1)$-ball.
\end{lemma}
Let $X=\{x_1,x_2,\dots,x_n\}$ be a set of indeterminates.
For $\aaa=(a_1,a_2,\dots,a_n)\in \ZZ_{\geq 0}^n$,
we write
$$x^\aaa = x_1^{a_1}x_2^{a_2}\cdots x_n^{a_n}.$$
Fix $\cc=(c_1,c_2,\dots,c_n) \in \ZZ_{\geq 0}^n$.
A $\cc$-multicomplex is said to be \textit{$\cc$-full}
if it is the set of all $\cc$-monomials.
A $\cc$-multicomplex which is not $\cc$-full is called a \textit{proper $\cc$-multicomplex}.
Let
$$\widetilde X_i=\{ x_i^{(0)},x_i^{(1)},\dots,x_i^{(c_i)}\}$$
be sets of new indeterminates for $i=1,2,\dots,n$, and let
$$\widetilde X= \widetilde X_1 \bigcup \widetilde X_2 \bigcup \cdots \bigcup \widetilde X_n.$$
\begin{definition}
For any $\cc$-monomial $x^\aaa=x_1^{a_1}x_2^{a_2}\cdots x_n^{a_n}$,
let
$$F_\cc(x^\aaa)=
\widetilde X \setminus \{ x_1^{(a_1)},x_2^{(a_2)},\dots,x_n^{(a_n)} \}.
$$
For any $\cc$-multicomplex $M$,
we define the simplicial complex $\B_\cc(M)$ on $\widetilde X$ by
$$\B_\cc(M)= \left\langle F_\cc(x^\aaa): x^\aaa \in M \right \rangle.$$
\end{definition}
For simplicial complexes $\Delta$ and $\Gamma$ with disjoint vertices,
the simplicial complex
$$\Delta * \Gamma =\{F \cup G: F \in \Delta \mbox{ and } G \in \Gamma\}$$
is called the {\em join of $\Delta$ and $\Gamma$}.
Let $\partial \widetilde X_i=\{F \subset \widetilde X_i: F \ne \widetilde X_i\}$ for $i=1,2,\dots,n$,
and let
$$\Lambda_\cc=\partial \widetilde X_1 * \partial \widetilde X_2 * \cdots * \partial \widetilde X_n.$$
Then $\Lambda_\cc$ is a simplicial $(|\cc|-1)$-sphere,
where $|\cc|=c_1+\cdots+c_n$,
since each $\partial \widetilde X_i$ is the boundary of a simplex and since the join of two simplicial spheres is again a simplicial sphere.
Since $F_\cc(x^\aaa)=\widetilde X \setminus \{x_1^{(a_1)},\dots,x_n^{(a_n)}\}$ is a facet of $\Lambda_\cc$,
$\B_\cc(M)$ is a subcomplex of $\Lambda_\cc$ having the same dimension as $\Lambda_\cc$.
\begin{lemma} \label{ball}
Let $M$ be a $\cc$-multicomplex.
\begin{itemize}
\item[(i)] $\mathcal B_\cc(M) = \Lambda_\cc$ if and only if $M$ is $\cc$-full.
\item[(ii)] $\B_\cc(M)$ is shellable.
\item[(iii)] If $M$ is not $\cc$-full then $\B_\cc(M)$ is a simplicial $(|\cc|-1)$-ball.
\end{itemize}
\end{lemma}
\begin{proof}
(i) is straightforward.
(ii) was essentially proved in \cite[Theorem 4.3]{So}.
But we include a proof for the sake of completeness.
Let $x^\aaa \in M$ be a monomial that does not divide any other monomial in $M$.
It is enough to prove that
\begin{align}
\label{1.1}
\B_\cc(M\setminus \{x^\aaa\})\cap \langle F_\cc(x^\aaa) \rangle
=\big \langle F_\cc(x^\aaa) \setminus \{x_i^{(j)}\}: i=1,2,\dots,n,\ 0 \leq j \leq a_i-1 \big\rangle.
\end{align}
We first show that the left-hand side contains the right-hand side.
Observe that $x^\aaa x_i^{-a_i+j} \in M$ if $j \leq a_i -1$.
Then $F_\cc(x^\aaa) \setminus \{x_i^{(j)}\} \subset F_\cc (x^\aaa x_i^{-a_i+j}) \in \B_\cc(M\setminus \{x^\aaa\})$.
Second, we show that the right-hand side contains the left-hand side.
It is enough to prove that
$$G=\{ x_i^{(j)}: i=1,2,\dots,n,\ 0\leq j \leq a_i-1\}$$
is not contained in $\B_\cc(M\setminus\{x^\aaa\})$
since the set $G$ is the (unique) smallest element among the elements in $\langle F_\cc(x^\aaa)\rangle$ which are not contained in the right-hand side of \eqref{1.1}.
Suppose contrary that $G \in \B_\cc(M\setminus\{x^\aaa\})$.
There exists a monomial $x^\bb \in M$ with $x^\bb \ne x^\aaa$ such that $G \subset F_\cc(x^\bb)$.
Then, by the definition of $F_\cc(-)$, we have $b_i \geq a_i$ for all $i$,
which implies $x^\aaa$ divides $x^\bb$.
This contradicts the choice of $x^\aaa$.
(iii)
By (i),
$\B_\cc(M) \subsetneq \Lambda_\cc$.
Then Lemma \ref{shellball} and (ii) say that $\B_\cc(M)$ is a simplicial $(|\cc|-1)$-ball.
\end{proof}
We call $\B_\cc(M)$ the {\em Bier ball of $M$ with respect to $\cc$.}
\begin{remark}\label{r1}
$\B_\cc(M)$ depends not only on $M$ but also on $\cc$.
However, the crucial case will be the case when $x_1^{c_1} \cdots x_n^{c_n}$
is equal to the least common multiple of monomials in $M$
since $\B_{(c_1+1,c_2,\dots,c_n)}(M)=\{x_1^{(c_1+1)}\}*\B_\cc(M)$ is a cone of $\B_\cc(M)$.
\end{remark}
\begin{remark}
Clearly $\B_{(c_1,\dots,c_{n-1},0)}(M)=\B_{(c_1,\dots,c_{n-1})}(M)$.
Thus one can assume $\cc \in \ZZ_{\geq 1}^n$.
However, in this paper we include $0$
since considering $\ZZ_{\geq 0}^n$ is convenient for induction purposes.
\end{remark}
\begin{remark}
In the special case when $M$ is the set of all monomials of degree $\leq k$ for some integer $k\geq 0$,
Lemma \ref{ball}(iii) was proved in \cite[Theorem 3.1]{HP}.
\end{remark}
\begin{remark}
Although we only prove shellability,
$\B_\cc(M)$ is vertex decomposable (see \cite[p.\ 1854]{Bj}).
Indeed, both the link and the deletion of $\B_\cc(M)$ w.r.t.\ $x_1^{(c_1)}$
are Bier balls.
\end{remark}
Now we define Bier spheres of multicomplexes.
Let $\Delta$ be a simplicial $(d-1)$-ball.
Then each $(d-2)$-dimensional face
of $\Delta$ is contained in at most two facets of $\Delta$.
Then its boundary
$$\partial \Delta = \langle F \in \Delta: \#F=d-1,\ \mbox{$F$ is contained in exactly one facet of $\Delta$}\rangle$$
is a simplicial $(d-2)$-sphere.
\begin{definition}
Let $M$ be a proper $\cc$-multicomplex.
We call the boundary $\bier_\cc (M)\\
=\partial \B_\cc(M)$ of the $(|\cc|-1)$-dimensional simplicial ball $\B_\cc(M)$ the
\textit{Bier sphere of a multicomplex $M$ with respect to $\cc$}.
\end{definition}
In the rest of this section,
we study some easy combinatorial properties of Bier spheres.
First,
we describe the facets of $\bier_\cc (M)$.
For any monomial $x^\aaa$ and for any pure power of a variable $x_i^j$,
we define
$$x^\aaa \diamond x_i^j= x^\aaa x_i^{-a_i+j} = x_1^{a_1} \cdots x_{i-1}^{a_{i-1}} x_i^j x_{i+1}^{a_{i+1}} \cdots x_n^{a_n}$$
(we assume $x_i^0 \ne 1$ when we consider $x^\aaa \diamond x_i^0$).
The facets of $\bier_\cc (M)$ are given as follows.
\begin{proposition} \label{facets}
Let $M$ be a proper $\cc$-multicomplex. Then
$$\bier_\cc(M)=\big\langle
F_\cc(x^\aaa)\setminus \{x_i^{(j)} \}: x^\aaa \in M,\ x^\aaa \diamond x_i^j \not \in M,\
a_i<j \leq c_i \big \rangle.
$$
\end{proposition}
\begin{proof}
Let $x^\aaa \in M$.
Note that $x^\aaa \diamond x_i^j \not \in M$ implies $a_i<j$.
$\bier_\cc(M)$ is generated by all codimension $1$ faces $F$ of $\B_\cc(M)$ such that $F$ is contained in exactly one facet of $\B_\cc(M)$.
It is easy to see that if $F_\cc(x^\aaa) \setminus \{x_i^{(j)}\}$,
where $x_i^{(j)} \in F_\cc(x^\aaa)$,
is contained in
$F_\cc(x^\bb)$ for some $\cc$-monomial $x^\bb$, then $x^\bb$ must be either
$x^\aaa$ or $x^\aaa \diamond x_i^j$.
This implies the desired formula.
\end{proof}
\begin{example}
Let $\cc=(2,2)$ and $M=\{1,x,y,x^2,y^2\}$. Then
\begin{align*}
\B_{(2,2)}(M)=
\left \langle
\begin{array}{l}
x^{(1)}x^{(2)}y^{(1)}y^{(2)},\
x^{(0)}x^{(2)}y^{(1)}y^{(2)},\
x^{(1)}x^{(2)}y^{(0)}y^{(2)},\\
x^{(0)}x^{(1)}y^{(1)}y^{(2)},\
x^{(1)}x^{(2)}y^{(0)}y^{(1)}
\end{array}
\right \rangle,
\end{align*}
where we identify $x^{(i)}x^{(j)}y^{(k)}y^{(\ell)}$ with $\{x^{(i)},x^{(j)},y^{(k)},y^{(\ell)}\}$ for simplicity, and
\begin{align*}
\bier_{(2,2)}(M)=
\left \langle
\begin{array}{l}
x^{(0)}x^{(2)}y^{(1)},\
x^{(0)}x^{(2)}y^{(2)},\
x^{(1)}y^{(0)}y^{(2)},\
x^{(2)}y^{(0)}y^{(2)},\\
x^{(0)}x^{(1)}y^{(1)},\
x^{(0)}x^{(1)}y^{(2)},\
x^{(1)}y^{(0)}y^{(1)},\
x^{(2)}y^{(0)}y^{(1)}
\end{array}
\right \rangle.
\end{align*}
Then $\bier_{(2,2)}(M)$ is the boundary complex of the octahedron, that is,
$\bier_{(2,2)}(M)=\partial \langle \{x^{(0)},y^{(0)}\} \rangle*\partial \langle \{x^{(1)},x^{(2)}\}\rangle*\partial \langle \{y^{(1)},y^{(2)}\}\rangle$.
\end{example}
\begin{example}
\label{codim1}
Here we classify Bier spheres of multicomplexes with one variable.
Let $\cc=(c)$ and $M=\{1,x,x^2,\dots,x^b\}$, where $0 \leq b < c$.
Then
$$
\B_{(c)}(M)=
\left \langle
\{x^{(0)},x^{(1)}, \cdots, x^{(c)}\} \setminus \{x^{(i)}\}: i=0,1,\dots,b
\right \rangle.
$$
and
$$
\ \bier_{(c)}(M)=
\left \langle
\{x^{(0)},x^{(1)}, \cdots, x^{(c)}\} \setminus \{x^{(i)},x^{(j)}\}: i=0,1,\dots,b,\ j=b+1,\dots,c
\right \rangle.
$$
Hence $\bier_{(c)}(M)=\partial \langle \{x^{(0)},x^{(1)}, \cdots, x^{(b)}\}\rangle * \partial \langle \{x^{(b+1)},x^{(b+2)}, \cdots, x^{(c)}\}\rangle$.
\end{example}
If $\cc=(1,1,\dots,1)$ then $\cc$-multicomplexes can be identified with simplicial complexes.
We show that $\bier_{(1,1,\dots,1)}(M)$ are classical Bier spheres.
Let $\Delta$ and $\Gamma$ be simplicial complexes on the vertex sets $V$ and $W$, respectively.
We say that $\Delta$ is isomorphic to $\Gamma$ if there is a bijection $\phi: V \to W$
such that $\Gamma=\{ \phi(F):F \in \Delta\}$.
\begin{theorem} \label{bier}
Let $\cc=(1,1,\dots,1)$ and $M$ a proper $\cc$-multicomplex on $X$.
Let $\Delta$ be the simplicial complex defined by $\Delta=\{\{i_1,\dots,i_k\} \subset [n]: x_{i_1} \cdots x_{i_k} \in M\}$.
Then $\bier_\cc(M)$ is combinatorially isomorphic to $\bier(\Delta)$.
\end{theorem}
\begin{proof}
A straightforward computation shows (see \cite[Lemma 5.6.4]{Ma})
$$\bier (\Delta) =
\left\langle X_F \cup Y_{[n]\setminus (F\cup\{j\})}:
F \in \Delta,\ F\cup \{j\} \not \in \Delta \right \rangle.
$$
For $F \subset [n]$, let $x^F = \prod_{i \in F} x_i$.
Observe that $F_\cc(x^F)=\{x_i^{(0)}:i \in F\}\cup \{x_i^{(1)}:i \not\in F\}$.
Proposition \ref{facets} shows
$$
\bier_\cc(M)=
\left\langle
F_\cc(x^F)\setminus\{x_j^{(1)}\}:F \in \Delta,\ F\cup\{j\} \not \in \Delta
\right \rangle,
$$
which proves the desired statement.
\end{proof}
Lemma \ref{ball} and Theorem \ref{bier}
provide a new proof of Bier's theorem.
The known proofs of Bier's theorem uses subdivisions or Bistellar flips.
See \cite{BPSZ,de,Ma}.
Finally, we compute face vectors of Bier spheres.
Let $\Delta$ be a $(d-1)$-dimensional simplicial complex.
Let $f_i=f_i(\Delta)$ be the number of faces of $\Delta$ of cardinality $i$.
The \textit{$f$-vector} of $\Delta$ is the vector $f(\Delta)=(f_0,f_1,\dots,f_d)$ where $f_0=1$,
and the \textit{$h$-vector} $h(\Delta)=(h_0,h_1,\dots,h_d) \in \ZZ^{d+1}$ of $\Delta$ is defined by the relation
$$\sum_{i=0}^d h_i t^{d-i} = \sum_{i=0}^d f_i (t-1)^{d-i}.$$
Also, for $0 \leq i \leq \lfloor \frac d 2 \rfloor$, let $g_i= h_i -h_{i-1}$ where $g_0=1$.
The vector $g(\Delta)=(g_0,g_1,\dots,g_{\lfloor \frac d 2 \rfloor})$ is called the \textit{$g$-vector of $\Delta$}.
For a multicomplex $M$,
let $f_i(M)$ be the number of monomials in $M$ of degree $i$.
It is easy to see that knowing the $f$-vector of $\Delta$ is equivalent to knowing the $h$-vector of $\Delta$.
Also, if $\Delta$ is a simplicial sphere then knowing the $g$-vector of $\Delta$ is equivalent to knowing the $h$-vector of $\Delta$
by the Dehn-Sommerville equations $h_i=h_{d-i}$.
The $g$-vectors of Bier spheres are given by the following formula.
\begin{theorem}
\label{hvector}
Let $M$ be a proper $\cc$-multicomplex.
\begin{itemize}
\item[(i)] $h_i (\B_\cc (M))= f_i(M)$ for all $i$.
\item[(ii)] $g_i(\bier_\cc (M))= f_i(M) - f_{|\cc|-i}(M)$ for $i=0,1,\dots, \lfloor \frac {|\cc|-1} 2 \rfloor$.
\end{itemize}
\end{theorem}
\begin{proof}
(i) follows from the shelling \eqref{1.1} and the following fact:
If $\Delta$ is shellable with a shelling $F_1,\dots,F_s$ then, for $i \geq 1$,
\begin{align}
\label{shell}
h_i(\Delta)=\#\{k: k\geq 2,\ \langle F_1,\dots,F_{k-1}\rangle \cap \langle F_k \rangle
\mbox{ is generated by $i$ facets}\}.
\end{align}
See \cite[III, Proposition 2.3]{St}.
(ii) follows from the well-known fact that
if $B$ is a $(d-1)$-dimensional simplicial ball then $g_i(\partial B) = h_i(B)-h_{d-i}(B)$
for $i=0,1,\dots,\lfloor \frac {d-1} 2 \rfloor$.
See \cite[p.\ 137]{St}.
\end{proof}
\section{Shellability}
In the previous section, we prove that $\B_\cc(M)$ is a simplicial ball by using shellability.
Shellability is an important property in combinatorial topology,
which induces several important topological and enumerative properties
like Lemma \ref{shellball} and formula \eqref{shell}.
Thus if one obtains a construction of simplicial spheres
it is a fundamental question to ask whether they are shellable.
Bj\"orner et al.\ \cite{BPSZ} proved that Bier spheres of simplicial complexes are shellable.
In this section,
we extend this result for Bier spheres of multicomplexes.
Let $\cc=(c_1,\dots,c_n) \in \ZZ_{\geq 0}^n$ and let $M$ be a proper $\cc$-multicomplex.
For any $\cc$-monomial $x^\aaa$ and for any pure power of a variable $x_i^j$ with $0 < j \leq c_i$,
let
$$G(x^\aaa;x_i^j)=F_\cc(x^\aaa) \setminus \{x_i^{(j)}\},$$
and let
$$
\facet (M) =
\big\{ G(x^\aaa;x_i^j) : x^\aaa \in M,\ x^\aaa \diamond x_i^j \not \in M,\ a_i < j \leq c_i \big\}.
$$
By Proposition \ref{facets},
$\facet (M)$ is the set of the facets of $\bier_\cc (M)$.
We introduce the order $\succ$ on $\facet (M)$ as follows:
Let $>_\lex$ be the lexicographic order on $\ZZ^n$.
Thus, for $\aaa, \bb \in \ZZ^n$,
$\aaa >_\lex \bb$ if and only if there exists an $i$ such that $a_i > b_i$ and $a_k = b_k$ for all $k<i$.
Let $G(x^\aaa;x_p^s), G(x^\bb;x_q^t) \in \facet (M)$.
We define $G(x^\aaa;x_p^s) \succ G(x^\bb;x_q^t)$ if one of the following conditions hold
\begin{itemize}
\item[(i)]
$(a_1,a_2,\dots,a_{p-1},s,-a_{p+1},\dots,-a_n) >_\lex (b_1,b_2,\dots,b_{q-1},t,-b_{q+1},\dots,-b_n),$
\item[(ii)] $(a_1,a_2,\dots,a_{p-1},s,-a_{p+1},\dots,-a_n) = (b_1,b_2,\dots,b_{q-1},t,-b_{q+1},\dots,-b_n)$ and $x_p^{a_p+1}>_\lex x_q^{b_q+1}$,
\end{itemize}
where we define $x_p^{a_p+1}>_\lex x_q^{b_q+1}$ if $p<q$ or $p=q$ and $a_p>b_q$.
Clearly, $\succ$ is a total order on $\facet (M)$.
\begin{theorem} \label{shellable}
For any proper $\cc$-multicomplex $M$,
$\bier_\cc (M)$ is shellable.
\end{theorem}
\begin{proof}
We show that the order $\succ$ on $\facet (M)$ induces a shelling of $\bier_\cc(M)$.
Fix $G(x^\aaa;x_p^s) \in \facet (M)$.
Let
$$\Gamma = \big\langle G(x^\bb;x_q^t) \in \facet (M): G(x^\bb;x_q^t) \succ G(x^\aaa;x_p^s)\big\rangle,$$
and let
\begin{align*}
H=&\ \{x_i^{(j)}\in \widetilde X: i<p \mbox{ and } j>a_i\}\\
&\ \bigcup \{x_i^{(j)}\in \widetilde X: i>p \mbox{ and } j<a_i\}\\
&\ \bigcup \left(\{x_p^{(j)}\in \widetilde X: j>s\} \cup \{x_p^{(j)}\in \widetilde X: x^\aaa \diamond x_p^j \in M \mbox{ and } j> a_p\} \right).
\end{align*}
Since $s > a_p$, we have
$$H \subset G(x^\aaa;x_p^s)=\widetilde X \setminus \{ x_1^{(a_1)}, \cdots,x_n^{(a_n)},x_p^{(s)}\}.$$
To prove the statement,
it is enough to prove that
\begin{align}
\label{two-one}
\Gamma \cap \big\langle G(x^\aaa;x_p^s) \big\rangle = \big\langle G(x^\aaa;x_p^s) \setminus \{ x_i^{(j)}\} : x_i^{(j)} \in H \big\rangle.
\end{align}
Let $u=x^\aaa \diamond x_p^s$.
The inclusion `$\supset$' follows from the following case analysis.\smallskip
[\textit{Case 1.1}]
Suppose $i<p$ and $x^\aaa \diamond x_i^j \not \in M$ with $j > a_i$.
Then $G(x^\aaa;x_i^j)\in \facet(M)$ satisfies
$G(x^\aaa;x_i^j) \succ G(x^\aaa;x_p^s)$ and
$G(x^\aaa;x_i^j) \supset G(x^\aaa;x_p^s) \setminus \{x_i^{(j)}\}$.
[\textit{Case 1.2}]
Suppose $i<p$ and $x^\aaa \diamond x_i^j \in M$ with $j > a_i$.
Then $G(x^\aaa\diamond x_i^j ;x_p^s)\in \facet(M)$ satisfies
$G(x^\aaa\diamond x_i^j ;x_p^s) \succ G(x^\aaa;x_p^s)$ and
$G(x^\aaa\diamond x_i^j ;x_p^s) \supset G(x^\aaa;x_p^s) \setminus \{x_i^{(j)}\}$.
[\textit{Case 2.1}]
Suppose $i>p$ and $u \diamond x_i^j \not \in M$ with $j < a_i$.
Then $G(x^\aaa\diamond x_i^j;x_p^s)\in \facet(M)$ satisfies
$G(x^\aaa\diamond x_i^j;x_p^s) \succ G(x^\aaa;x_p^s)$ and
$G(x^\aaa\diamond x_i^j;x_p^s) \supset G(x^\aaa;x_p^s) \setminus \{x_i^{(j)}\}$.
[\textit{Case 2.2}]
Suppose $i>p$ and $u \diamond x_i^j \in M$ with $j < a_i$.
Then $G(u \diamond x_i^j ;x_i^{a_i})\in \facet(M)$ satisfies
$G(u \diamond x_i^j ;x_i^{a_i}) \succ G(x^\aaa;x_p^s)$ and
$G(u \diamond x_i^j ;x_i^{a_i}) \supset G(x^\aaa;x_p^s) \setminus \{x_i^{(j)}\}$.
[\textit{Case 3.1}]
Suppose $i=p$ and $j > s$.
Then $G(x^\aaa;x_p^j)\in \facet(M)$ satisfies
$G(x^\aaa;x_p^j) \succ G(x^\aaa;x_p^s)$ and
$G(x^\aaa;x_p^j) \supset G(x^\aaa;x_p^s) \setminus \{x_i^{(j)}\}$.
[\textit{Case 3.2}]
Suppose $i=p$ and $x^\aaa \diamond x_p^j \in M$ with $j>a_p$.
Then $G(x^\aaa\diamond x_p^j;x_p^s)\in \facet(M)$ satisfies
$G(x^\aaa\diamond x_p^j;x_p^s) \succ G(x^\aaa;x_p^s)$ and
$G(x^\aaa\diamond x_p^j;x_p^s) \supset G(x^\aaa;x_p^s) \setminus \{x_i^{(j)}\}$.\medskip
Next, we prove that the right-hand side contains the left-hand side in \eqref{two-one}.
In the same way as in the proof of Lemma \ref{ball}(ii),
what we must prove is $H \not \in \Gamma$.
Suppose $G(x^\bb;x_q^t) \in \facet(M)$ contains $H$ and satisfies $G(x^\bb;x_q^t) \succeq G(x^\aaa;x_p^s)$.
We prove that $G(x^\bb;x_q^t) = G(x^\aaa;x_p^s)$.
Note that $x^\bb \in M$ and $x^\aaa\diamond x_p^s \not \in M$.
Let
$$\aaa'=(a_1,\dots,a_{p-1},s,-a_{p+1},\dots,-a_n)$$
and
$$\mathbf d = (d_1,\dots,d_n) =(b_1,\dots,b_{q-1},t,-b_{q+1},\dots,-b_n).$$
Since $G(x^\bb;x_q^t) \supset H$,
by the definition of $H$, we have
\begin{align}
\label{2.2}
d_1 \leq a_1, \dots, d_{p-1} \leq a_{p-1}, d_p \leq s \mbox{ and } b_{p+1} \geq a_{p+1}, \dots, b_n \geq a_n.
\end{align}
Observe $\mathbf d \geq_\lex \aaa'$ since $G(x^\bb;x_q^t) \succeq G(x^\aaa;x_p^s)$.
If $d_i <a_i$ for some $1 \leq i \leq p-1$ or $d_p <s$ then $\mathbf d <_\lex \aaa'$.
Hence we have
$$
d_1=a_1,\dots,d_{p-1}=a_{p-1} \mbox{ and } d_p =s.
$$
In particular, we have $q \geq p$ since if $q<p$ then $d_p \leq 0$ but $s>a_p \geq 0$ is positive.
We show $q=p$.
If $q>p$ then $b_p=d_p=s$ and $x^\aaa \diamond x_p^s \not \in M$ divides $x^\bb=x_1^{a_1} \cdots x_{p-1}^{a_{p-1}}x_p^s x_{p+1}^{b_{p+1}} \cdots x_n^{b_n}$ by \eqref{2.2}, which contradicts the fact that $x^\bb \in M$.
Now we know $b_1=a_1,\dots,b_{p-1}=a_{p-1}$, $p=q$
and $t=d_p=s$.
By \eqref{2.2} if $b_i > a_i$ for some $p+1 \leq i \leq n$ then $\mathbf d <_\lex \aaa'$.
Thus we have $b_{p+1}=a_{p+1}, \dots,b_n=a_n$,
and therefore $\aaa'=\mathbf d$.
It remains to prove $b_p=a_p$.
Since $G(x^\bb;x_q^t) \succ G(x^\aaa;x_p^s)$,
by the definition (ii) of the order $\succ$ we have $b_p \geq a_p$.
On the other hand, the definition of $H$ says that, for any $x_p^{(j)} \not \in H$ with $j>a_p$, one has $x^\aaa \diamond x_p^j \not \in M$.
This shows that if $b_p>a_p$ then $x^\bb=x^\aaa \diamond x_p^{b_p} \not \in M$ since $G(x^\bb;x_q^t)$ contains $H$,
which contradicts $x^\bb \in M$.
Hence $b_p=a_p$.
\end{proof}
\begin{example}
Let $\cc=(2,2)$ and $M=\{1,x,y,y^2\}$.
Then $\bier_\cc(M)$ is generated by
\begin{align*}
&\big\{
G(1;x^2),G(x;x^2),G(x;y),G(x;y^2),G(y;x),G(y;x^2),G(y^2;x),G(y^2;x^2)\big\}
\\
&=
\left\{
\begin{array}{ll}
x^{(1)}y^{(1)}y^{(2)},x^{(0)}y^{(1)}y^{(2)},x^{(0)}x^{(2)}y^{(2)},
x^{(0)}x^{(2)}y^{(1)},\\
x^{(2)}y^{(0)}y^{(2)},x^{(1)}y^{(0)}y^{(2)},
x^{(2)}y^{(0)}y^{(1)},x^{(1)}y^{(0)}y^{(1)}
\end{array}
\right\}.
\end{align*}
The proof of Theorem \ref{shellable} shows that
$$
G(x;x^2)\! \succ \!
G(1;x^2)\! \succ \!
G(y;x^2)\! \succ \!
G(y^2;x^2)\! \succ \!
G(x;y^2)\! \succ \!
G(x;y)\! \succ \!
G(y;x)\! \succ \!
G(y^2;x)
$$
is a shelling of $\bier_\cc(M)$.
More precisely, the above shelling is
\begin{align*}
& x^{(0)}y^{(1)} y^{(2)} \succ
\dot x^{(1)}y^{(1)}y^{(2)} \succ
x^{(1)}\dot y^{(0)}y^{(2)} \succ
x^{(1)}\dot y^{(0)} \dot y^{(1)}\! \succ
x^{(0)}\dot x^{(2)}y^{(1)}
\succ
x^{(0)}\dot x^{(2)} \dot y^{(2)} \\
&\succ
\dot x^{(2)} \dot y^{(0)}y^{(2)} \succ
\dot x^{(2)} \dot y^{(0)} \dot y^{(1)},
\end{align*}
where variables with $\cdot$ correspond to variables in $H$.
\end{example}
\section{Bier spheres and polarization}
The construction of $\bier_\cc (M)$ is inspired by the study of polarizations
of monomial ideals in commutative algebra.
In this section, we study connections between Bier spheres and polarization.
We recall some basics on commutative algebra.
Let $S=K[x_1,\dots,x_n]$ be the polynomial ring over a field $K$ with $\deg x_i=1$ for any $i$.
For a simplicial complex $\Delta$ on $X=\{x_1,\dots,x_n\}$,
the ideal
$$I_\Delta=(x_{i_1}\cdots x_{i_k}: \{x_{i_1},\dots,x_{i_k}\} \subset X,\ \{x_{i_1},\dots,x_{i_k}\} \not \in \Delta)$$
is called the \textit{Stanley-Reisner ideal of $\Delta$}.
The ring
$$K[\Delta]=S/I_\Delta$$
is called the \textit{Stanley-Reisner ring of $\Delta$}.
Note that the correspondence $\Delta \leftrightarrow I_\Delta$ gives a one-to-one correspondence between
simplicial complexes on $X$ and squarefree monomial ideals in $S$.
Let $I \subset S$ be a homogeneous ideal and $R=S/I$.
The \textit{(Krull) dimension} of $R$ is the maximal number of homogeneous elements of $R$ which are algebraically independent over $K$.
The \textit{Hilbert series of $R$} is the formal power series $H_R(t)=\sum_{k \geq 0} (\dim_K R_k) t^k$,
where $R_k$ is the graded component of $R$ of degree $k$.
It is known that $H_R(t)$ is a rational function of the form
$H_R(t)=(h_0+h_1t+\cdots+h_s t^s)/ {(1-t)^d},$
where $h_s \ne 0$ and where $d$ is the Krull dimension of $R$
\cite[Corollary 4.1.8]{BH}.
We write $h_i(R)=h_i$ for all $i \geq 0$,
where $h_i=0$ if $i >s$.
The vector $(h_0,h_1,\dots,h_s)$ is called the \textit{$h$-vector of $R$}.
Note that if $R=K[\Delta]$ then $s \leq d$, $h_i(\Delta)=h_i(R)$ for all $i \leq d$
and the Krull dimension of $R$ is equal to $\dim \Delta+1$.
Fix $\cc=(c_1,c_2,\dots,c_n) \in \ZZ_{\geq 0}^n$.
Let $\overline \cc=(c_1+1,c_2+1,\dots,c_n+1)$.
For a monomial ideal $I \subset S$,
we write $G(I)$ for the unique minimal set of monomial generators of $I$.
A monomial ideal $I \subset S$ is called a \textit{$\overline \cc$-ideal}
if $I$ is generated by $\overline \cc$-monomials.
\begin{definition}
Let $I \subset S$ be a $\overline \cc$-ideal
and let $\widetilde S=K[x_{i,j}:1 \leq i \leq n, 0 \leq j \leq c_i]$.
The {\em polarization of a $\overline \cc$-monomial $x^\aaa=x_1^{a_1}x_2^{a_2}\cdots x_n^{a_n}$ (with respect to $\overline \cc$)} is the squarefree monomial
$$\pol(x^\aaa)=
\prod_{a_i \ne 0} (x_{i,0} x_{i,1}\cdots x_{i,a_i-1}) \in \widetilde S.$$
The \textit{polarization of $I$ (with respect to $\overline \cc$)} is the squarefree monomial ideal
$$\pol(I)=\big(\pol(x^\aaa): x^\aaa \in G(I)\big) \subset \widetilde S.$$
\end{definition}
The ideal $\pol(I)$ has following properties (see e.g., \cite[pp.\ 59--60]{MS}):
\begin{itemize}
\item $S/I$ and $\widetilde S/\pol(I)$ have the same graded Betti numbers,
in particular, have the same $h$-vector.
\item $\dim \widetilde S/\pol(I)= \dim S/I+ |\cc|$.
\end{itemize}
In the rest of this section,
we identify $x_{i,j}$ and $x_i^{(j)}$,
and regard $\B_\cc(M)$ and $\bier_\cc(M)$ as simplicial complexes on $V=\{x_{i,j}:1\leq i\leq n, 0 \leq j \leq c_i\}$.
For a $\cc$-multicomplex $M$,
let $I(M) \subset S$ be the ideal generated by all monomials which are not in $M$.
If $M$ is a $\cc$-multicomplex then $I(M)$ is a $\overline \cc$-ideal since $I(M)$ contains $x_1^{c_1+1},\dots,x_n^{c_n+1}$.
\begin{lemma} \label{jahan}
For any $\cc$-multicomplex $M$,
$I_{\B_\cc(M)}=\pol(I(M)).$
\end{lemma}
\begin{proof}
Let $\Delta$ be the simplicial complex such that $I_\Delta=\pol(I(M))$.
We claim $\Delta = \B_\cc(M)$.
Since taking polarization does not change $h$-vectors,
by Lemma \ref{hvector}(i),
$$h_i(\Delta)=h_i(\widetilde S/I_\Delta)=h_i(S/I(M))=
\#\{x^\aaa \in M : \deg x^\aaa=i\}=h_i(\B_\cc(M)).$$
Since $\dim \Delta=\dim(\widetilde S/\pol(I(M)))-1=|\cc|-1$,
the above equations show that $\Delta$ and $\B_\cc(M)$ have the same $f$-vector.
Let $F=F_\cc(x^\aaa)=\widetilde X\setminus \{ x_1^{(a_1)},\dots,x_n^{(a_n)}\}$,
where $x^\aaa \in M$,
be a facet of $\B_\cc(M)$.
Since $\Delta$ and $\B_\cc(M)$ have the same $f$-vector,
to prove $\Delta=\B_\cc(M)$,
it is enough to prove that $\prod_{x_i^{(j)} \in F} x_{i,j}$ is not contained in $I_\Delta=\pol(I(M))$.
Suppose contrary that $\prod_{x_i^{(j)} \in F} x_{i,j} \in \pol(I(M))$.
Then there is a monomial $x^\bb \in I(M)$
such that $\pol(x^\bb)$ divides $\prod_{x_i^{(j)} \in F} x_{i,j}$.
By the definition of polarization, we have $b_1 \leq a_1,\dots,b_n \leq a_n$.
Then $x^\bb$ divides $x^\aaa$, which contradicts $x^\aaa \not \in I(M)$.
\end{proof}
\begin{remark}
Jahan \cite[Lemma~3.7]{So} computed the facets of the simplicial complex whose Stanley-Reisner ideal is the polarization of a monomial ideal.
This will give an alternative proof of Lemma \ref{jahan}.
See also \cite[Theorem 3.1]{HP}.
\end{remark}
Lemma \ref{jahan} allows us to study the structure of $\bier_\cc(M)$ from an algebraic viewpoint.
We discuss some algebraic aspects of Bier spheres later in Section 5.
Next we study generators of $I_{\bier_\cc(M)}$.
Let $\Delta$ be a simplicial complex on $[n]$.
Recall that the Alexander dual of $\Delta$
is the simplicial complex $\Delta^\vee=\{F \subset [n]: [n] \setminus F \not \in \Delta\}$.
Let $T=K[x_1,\dots,x_n,y_1,\dots,y_n]$,
$X_\Delta =(x^F: F \not \in \Delta) \subset T$
and $Y_{\Delta^\vee} =(y^F: F \not \in \Delta^\vee) \subset T$,
where $x^F=\prod_{i \in F} x_i$ and $y^F=\prod_{i \in F} y_i$.
Then it is not hard to see that
$$I_{\bier(\Delta)}
=X_\Delta + Y_{\Delta^\vee} +(x_1y_1,x_2y_2,\dots,x_ny_n).$$
Indeed,
one has $x^F y^G \not \in I_{\bier(\Delta)}$
if and only if $F \cap G = \emptyset$,
$x^F \not \in X_\Delta$ and $y^G \not \in Y_{\Delta^\vee}$,
which implies the above equation.
We show that a similar formula holds for
Bier spheres of multicomplexes.
We first recall Alexander duality of monomial ideals
introduced by Miller \cite[Definition 1.5]{Mi2}.
\begin{definition}
\label{alex}
Let $M$ be a $\cc$-multicomplex.
The \textit{Alexander dual of $M$ with respect to $\cc$} is the multicomplex defined by
$$
M^{\vee}=\{ x_1^{c_1-a_1}\cdots x_n^{c_n-a_n}: x^\aaa \mbox{ is a $\cc $-monomial such that } x^\aaa \not \in M\}.
$$
Note that $M^\vee$ depends not only on $M$ but also on $\cc$.
We also write an ideal-theoretic definition of Alexander duality.
For a $\cc$-ideal $I \subset S$,
we call the ideal
$$I^\vee=(x^{c_1-a_1}\cdots x_n^{c_n-a_n}: x^\aaa \mbox{ is a $\cc $-monomial such that } x^\aaa \not \in I) \subset S$$
the \textit{Alexander dual of $I$ with respect to $\cc$.}
\end{definition}
Let $I_\cc(M) \subset S$ be the monomial ideal generated by all $\cc$-monomials which are not in $M$.
Then the above two definitions are related by $I_\cc(M)^\vee = I_\cc(M^\vee)$.
For any $\overline \cc$-monomial $x^\aaa \in S$,
let
$$\pol^{*}(x^\aaa)= \prod_{a_i \ne 0} (x_{i,c_i}x_{i,c_i-1}\cdots x_{i,c_i-a_i+1}).$$
Similarly, for a $\overline \cc$-ideal $I \subset S$, let $\pol^* (I) \subset \widetilde S$ be the ideal generated by $\{\pol^* (x^\aaa): x^\aaa \in G(I)\}$.
(Thus $\pol^*$(-) is the polarization by using an opposite ordering of the variables $x_{i,c_i},x_{i,c_i-1},\dots,x_{i,1}$.)
\begin{lemma}
\label{complement}
Let $M$ be a proper $\cc$-multicomplex and let
$$\mathcal B^*_\cc(M)= \langle F_\cc(x^\aaa): x^\aaa\mbox{ is a $\cc$-monomial such that } x^\aaa \not \in M\rangle$$
be the complementary ball of $\B_\cc(M)$ in $\Lambda_\cc$.
Then
$I_{\mathcal B^*_\cc(M)}= \pol^*(I(M^\vee))$.
\end{lemma}
\begin{proof}
Let $\pi$ be the permutation on the vertex set $V$ defined by $\pi(x_{i,j})=x_{i,c_i-j}$.
Then $\pi(F_\cc(x^\aaa))=F_\cc(x_1^{c_1-a_1}\cdots x_n^{c_n-a_n})$ and
$\pi(\B_\cc(M^\vee))=\B_\cc^*(M)$.
Hence
$$I_{\mathcal B^*_\cc(M)}= \pi (I_{\B_\cc(M^\vee)})= \pi \circ \pol(I(M^\vee))=\pol^*(I(M^\vee)),$$
as desired.
\end{proof}
\begin{theorem}
\label{generator}
Let $M$ be a proper $\cc$-multicomplex.
$$ I_{\bier_\cc (M)}=
\pol(I_\cc(M))+ \pol^*(I_\cc(M^{\vee})) +\pol(x_1^{c_1+1},\dots,x_n^{c_n+1}).$$
\end{theorem}
\begin{proof}
Since $\B_\cc^*(M)$ is the complementary ball of $\B_\cc(M)$ in $\Lambda_\cc$,
we have $\B_\cc(M) \cap \B_\cc^*(M)=\partial \B_\cc(M)=\bier_\cc(M)$.
Since $I(M)=I_\cc(M)+(x_1^{c_1+1},\dots,x_n^{c_n+1})$,
by Lemma \ref{complement} we have
$$I_{\bier_\cc(M)} = I_{\B_\cc(M)}+I_{\mathcal B^*_\cc(M)}=
\pol(I_\cc(M))+ \pol^{*}(I_\cc(M^{\vee})) +\pol(x_1^{c_1+1},\dots,x_n^{c_n+1}),$$
as desired.
\end{proof}
\begin{example}
Let $\cc=(2,2,2)$ and $M=\{1,x,y,z,xz,yz,z^2,yz^2\}$.
Then $I_\cc(M)=(x^2,y^2,xy,xz^2)$
and
$I_\cc(M^\vee)=(x y^2 z, x^2 y)$.
Thus
$$
I_{\bier_\cc(M)}=
(x_0x_1,y_0y_1,x_0y_0,x_0z_0z_1)
+(x_2y_2y_1z_2,x_2x_1y_2)+(x_0x_1x_2,y_0y_1y_2,z_0z_1z_2).$$
\end{example}
\begin{example} \label{3..4}
Theorem \ref{generator} may not give minimal generators.
For example, if $\Delta=\langle \{1,2\},\{3\}\rangle$ then
$$I_{\bier(\Delta)}=(x_1x_3,x_2x_3)+(y_3,y_1y_2)+(x_1y_1,x_2y_2,x_3y_3).$$
However, the set of generators $\{x_1x_3,x_2x_3,y_3,y_1y_2,x_1y_1,x_2y_2,x_3y_3\}$
is not minimal.
\end{example}
Finally we note the following result which immediately follows from the fact $(M^\vee)^\vee =M$.
\begin{corollary}
\label{dual}
Let $M$ be a proper $\cc$-multicomplex.
Then $\bier_\cc(M)$ and $\bier_\cc(M^\vee)$ are combinatorially isomorphic.
The isomorphism is given by the permutation of the vertices $x_{i,j} \to x_{i,c_i-j}$ for all $i,j$.
\end{corollary}
\section{Edge decomposability}
In this section, we show that Bier spheres are edge decomposable.
We first recall the definition of edge decomposability.
Let $\Delta$ be a simplicial complex on $V$.
The \textit{link of $\Delta$ with respect to a face $F \in \Delta$}
is the simplicial complex
$$
\lk_\Delta(F) = \{ G \subset V \setminus F : G \cup F \in \Delta\}.
$$
The \textit{contraction $\mathcal C_\Delta (i,j)$ of $\Delta$ with respect to an edge $\{i,j\} \in \Delta$}
is the simplicial complex which is obtained
from $\Delta$ by identifying the vertices $i$ and $j$,
in other words,
$$\mathcal C_\Delta(i,j) =
\{ F \in \Delta: i \not \in F\} \cup \{ (F \setminus \{i\}) \cup \{j\}:
i \in F \in \Delta\}.
$$
(We consider that $\mathcal C_\Delta(i,j)$ is a simplicial complex on $V \setminus \{i\}$.)
We say that $\Delta$ satisfies the \textit{Link condition
with respect to $\{i,j\} \subset V$} if
\begin{align*}
\lk_\Delta(\{i\}) \cap \lk_\Delta(\{j\}) = \lk_\Delta(\{i,j\}).
\end{align*}
\begin{definition}
\label{edgedecomp}
The boundary complex of a simplex and $\{\emptyset\}$ are \textit{edge decomposable},
and, recursively, a pure simplicial complex $\Delta$ is said to be \textit{edge decomposable}
if there exists an edge $\{i,j\} \in \Delta$ such that $\Delta$ satisfies the Link condition w.r.t.\
$\{i,j\}$ and both $\lk_\Delta(\{i,j\})$ and $\mathcal C_\Delta(i,j)$ are edge decomposable.
\end{definition}
Edge decomposability was introduced by Nevo \cite{Ne} in the study of the $g$-conjecture for spheres,
which states that the $g$-vector of a simplicial sphere is the $f$-vector of a multicomplex.
He proved that the $g$-vector of an edge decomposable sphere is non-negative in \cite{Ne}.
Later, it was proved in \cite{BN,Mu2} that the $g$-vector of an edge decomposable complex
is the $f$-vector of a multicomplex.
Unfortunately, not all spheres are edge decomposable.
For example, the boundary complex of
Lockeberg's non-vertex decomposable $4$-polytope (see \cite{Ha}) does not satisfy the Link condition w.r.t.\ any edge,
and therefore is not edge decomposable.
On the other hand, it was proved in \cite[Proposition 5.4]{Mu2} that Kalai's squeezed spheres are edge decomposable.
This shows that there are many edge decomposable spheres which are not realizable as polytopes.
In the rest of this section,
we prove that Bier spheres are edge decomposable.
\begin{lemma} \label{join}
If $\Delta$ and $\Gamma$ are edge decomposable then $\Delta*\Gamma$ is edge decomposable.
\end{lemma}
\begin{proof}
By the definition of edge decomposability,
we may assume that $\Delta$ and $\Gamma$ are the boundaries of simplexes of dimension at least $1$.
Suppose $\Delta*\Gamma=\partial F * \partial G$, where $F$ and $G$ are simplexes.
Then, for any pair $\{u,v\}$ of vertices, where $u \in F$ and $v \in G$,
it is easy to see that $\Delta*\Gamma$ satisfies the Link condition w.r.t.\ $\{u,v\}$,
$\lk_{\Delta*\Gamma}(\{u,v\})= \partial (F\setminus \{u\}) * \partial (G\setminus \{v\})$ and
$\mathcal C_{\Delta*\Gamma}(u,v)=\partial (F \cup G \setminus \{u\})$.
By using this fact,
the statement follows inductively.
\end{proof}
The above lemma and Example \ref{codim1} show that if $n=1$ then
$\bier_\cc (M)$ is edge decomposable.
We study the case when $n \geq 2$.
\begin{lemma}
\label{reduction}
Let $n \geq 2$ and let $M$ be a proper $\cc$-multicomplex.
If either $\{x_n^{(0)}\}$ or $\{x_n^{(c_n)}\}$ is not in $\bier_\cc(M)$ then there exist $\cc' \in \ZZ_{\geq 0}^{n-1}$
and a $\cc'$-multicomplex $M'$ on $X \setminus \{x_n\}$ such that $\bier_{\cc'}(M')$ is combinatorially isomorphic to $\bier_\cc(M)$.
\end{lemma}
\begin{proof}
By Corollary \ref{dual}, it is enough to consider the case when $\{x_n^{(0)}\} \not \in \bier_\cc(M)$.
Since $x_n^{(0)} \in G(I_{\bier_\cc(M)})$, by Theorem \ref{generator}, $x_n \in I_\cc(M)$,
and therefore $M$ contains no monomials which are divisible by $x_n$.
Then $M$ is a $(c_1,c_2,\dots,c_{n-1})$-multicomplex on $\{x_1,\dots,x_{n-1}\}$
and
\begin{align*}
\B_\cc(M)=&\ \B_{(c_1,\dots,c_{n-1},0)}(M)* \langle \{x_n^{(1)},\dots,x_n^{(c_n)}\}\rangle \\
\cong&\ \B_{(c_1,\dots,c_{n-1},0)}(M)*\langle \{x_1^{(c_1+1)},\dots,x_1^{(c_1+c_n)}\}\rangle\\
=&\ \B_{(c_1+c_n,\dots,c_{n-1},0)}(M)\\
=&\ \B_{(c_1+c_n,\dots,c_{n-1})}(M),
\end{align*}
as desired.
\end{proof}
\begin{lemma}
\label{linkcondition}
Let $n \geq 2$ and let $M$ be a proper $\cc$-multicomplex.
Suppose that both $\{x_1^{(c_1)}\}$ and $\{x_n^{(0)}\}$ are in $\bier_\cc(M)$.
Then $\{x_1^{(c_1)},x_n^{(0)}\} \in \bier_\cc(M)$ and
$\bier_\cc (M)$ satisfies the Link condition with respect to $\{x_1^{(c_1)},x_n^{(0)}\}$.
\end{lemma}
\begin{proof}
Note that $c_1>0$ and $c_n>0$ since $\{x_1^{(c_1)}\}$ and $\{x_n^{(0)}\}$ are in $\bier_\cc(M)$.
It is known that a simplicial complex $\Delta$ on $X$ satisfies the Link condition w.r.t.\ $\{x_i,x_j \} \in \Delta$
if and only if $G(I_\Delta)$ has no monomials which are divisible by $x_i x_j$ \cite[Lemma 2.1]{Mu2}.
By Theorem \ref{generator}, $G(I_{\bier_\cc(M)})$ cannot have monomials which are divisible by $x_1^{(c_1)}x_n^{(0)}$
since $G(\pol(I_\cc(M)))$ contains no monomials which are divisible by $x_1^{(c_1)}$
and $G(\pol^*(I_\cc(M^\vee)))$ contains no monomials which are divisible by $x_n^{(0)}$.
In particular, since $x_1^{(c_1)}, x_n^{(0)}$ and $x_1^{(c_1)}x_n^{(0)}$ are not in $G(I_{\bier_\cc(M)})$, we have $\{x_1^{(c_1)},x_n^{(0)}\} \in \bier_\cc(M)$.
\end{proof}
\begin{lemma}
\label{keylemma}
With the same notation as in Lemma \ref{linkcondition},
\begin{itemize}
\item[(i)]
the link of $\bier_\cc(M)$ with respect to $\{x_1^{(c_1)},x_n^{(0)}\}$
is combinatorially isomorphic to some Bier sphere $\bier_{\cc'} (M')$;
\item[(ii)]
the contraction of $\bier_\cc(M)$ with respect to $\{x_1^{(c_1)},x_n^{(0)}\}$
is combinatorially isomorphic to some Bier sphere $\bier_{\cc'} (M')$ with $\cc' \in \ZZ_{\geq 0}^{n-1}$.
\end{itemize}
\end{lemma}
\begin{proof}
(i)
Let $\cc'=(c'_1,c_2',\dots,c_n')=(c_1-1,c_2,\dots,c_n)$
and $M' = \{ x^\aaa \in M: x^\aaa \mbox{ is a $\cc'$-monomial}\}$.
If $M'$ is $\cc'$-full then by Proposition \ref{facets} all facets of $\bier_\cc(M)$ do not contain $x_1^{(c_1)}$,
which contradicts the assumption $\{x_1^{(c_1)} \} \in \bier_\cc(M)$.
Thus $M$ is not $\cc'$-full.
We claim that
\begin{align}
\label{4.1}
\bier_{\cc'}(M')= \lk_{\bier_\cc(M)}(\{x_1^{(c_1)}\}).
\end{align}
Since both complexes are simplicial spheres having the same dimension,
it is enough to prove that $\bier_{\cc'}(M')\subset \lk_{\bier_\cc(M)}(\{x_1^{(c_1)}\}).$
Let $F$ be a facet of $\bier_{\cc'}(M')$.
Then by Proposition \ref{facets}
there exist $x^\aaa \in M'$ and $x_i^j$ with $j \leq c_i'$ such that
$x^\aaa \diamond x_i^j \not \in M'$ and $F=F_{\cc'}(x^\aaa) \setminus \{x_i^{(j)}\}$.
By the definition of $M'$,
$x^\aaa \in M$ and $x^\aaa \diamond x_i^j \not \in M$.
Hence by Proposition \ref{facets}
$$F_\cc(x^\aaa) \setminus \{x_i^{(j)}\} = F \cup \{x_1^{(c_1)}\}$$
is a facet of $\bier_\cc(M)$, and therefore $F \in \lk_{\bier_\cc(M)}(\{x_1^{(c_1)}\})$.
Let $M''=(M')^{\vee}$ be the Alexander dual of $M'$ with respect to $\cc'$.
Then, by Corollary \ref{dual},
$\lk_{\bier_{\cc'}(M'')}(\{x_n^{(c_n)}\})$
is combinatorially isomorphic to $\lk_{\bier_{\cc'}(M')}(\{x_n^{(0)}\})=\lk_{\bier_\cc(M)}(\{x_1^{(c_1)},x_n^{(0)}\})$.
On the other hand,
\eqref{4.1} says that $\lk_{\bier_{\cc'}(M'')}(\{x_n^{(c_n)}\})$
is equal to the Bier sphere
$$\bier_{(c_1-1,c_2,\dots,c_{n-1},c_n-1)}\big(\big\{x^\aaa\in M'': x^\aaa \mbox{ is a $(c_1-1,c_2,\dots,c_{n-1},c_n-1)$-monomial}\big\}\big),$$
as desired.
\medskip
(ii)
Let $\cc'=(c'_1,\dots,c'_{n-1})=(c_1+c_n,c_2,\dots,c_{n-1})$.
For any monomial $x_1^sx_n^tu \in M$, where $u$ is a monomial on $x_2,\dots,x_{n-1}$,
let
\begin{align*}
\sigma(x_1^sx_n^tu)=
\begin{cases}
x_1^{s+t} u, & \mbox{ if $s=c_1$ or $t=0,$}\\
0, & \mbox{ otherwise},
\end{cases}
\end{align*}
and
$$M'=\{\sigma(x^\aaa):x^\aaa \in M\mbox{ and } \sigma(x^\aaa) \ne 0\}.$$
We will show that $\bier_{\cc'}(M')$ is combinatorially isomorphic to $\mathcal C_{\bier_\cc (M)}(x_1^{(c_1)},x_n^{(0)})$.
We first prove that $M'$ is indeed a proper $\cc'$-multicomplex.
Since $x^\cc \not \in M$, $x^{\cc'} \not \in M'$.
It is enough to prove that $M'$ is a multicomplex.
If $x_1^\ell u \in M'$ then one has $x_1^\ell u \in M$ if $\ell \leq c_1$
and $x_1^{c_1} x_n^{\ell-c_1} u \in M$ if $\ell >c_1.$
In the former case
$x_1^\ell u/x_i=\sigma(x_1^\ell u/x_i) \in M'$ for any $x_i$ which divides $x_1^\ell u$.
In the latter case
$x_1^\ell u/x_i=\sigma(x_1^{c_1}x_n^{\ell-c_1} u/x_i) \in M'$ for any $x_i$ with $i \geq 2$ which divides $x_1^\ell u$
and
$x_1^\ell u/x_1=\sigma(x_1^{c_1}x_n^{\ell-c_1} u/x_n) \in M'$.
Hence $M'$ is a multicomplex.
Let $\rho$ be the map from the set of subsets of $\{x_i^{(j)}: 1 \leq i \leq n,\ 0 \leq j \leq c_i\}$
to the set of subsets of $\{x_i^{(j)}: 1 \leq i \leq n-1,\ 0 \leq j \leq c'_i\}$
induced by
\begin{align*}
\rho(x_i^{(j)})=
\begin{cases}
x_i^{(j)}, & \mbox{ if $i \ne n$,}\\
x_1^{(c_1+j)}, & \mbox{ if $i =n$.}
\end{cases}
\end{align*}
Then, since $\rho(x_1^{(c_1)})=\rho(x_n^{(0)})$,
$\rho(\bier_\cc(M))$ is combinatorially isomorphic to
the contraction $\mathcal C_{\bier_\cc(M)}(x_1^{(c_1)},x_n^{(0)})$.
We claim that
$$\rho(\bier_\cc(M))
=\bier_{\cc'}(M').$$
It follows from \cite[Theorem 1.4]{Ne} and Lemma \ref{linkcondition} that
$\rho(\bier_\cc(M))$ is a simplicial sphere.
Thus it is enough to prove $\rho(\bier_\cc(M))
\supset \bier_{\cc'}(M')$.
Let $x_1^\ell u \in M'$, where $u$ is a monomial on $x_2,\dots,x_{n-1}$,
and let $x_i^j$ with $j \leq c'_i$ be such that $x_1^\ell u \diamond x_i^j \not \in M'$.
By Proposition \ref{facets},
what we must prove is $F_{\cc'}(x_1^\ell u )\setminus \{x_i^{(j)}\} \in \rho(\bier_\cc(M))$.
This follows from the following case analysis.
\medskip
[\textit{Case 1.1}]
Suppose $i \ne 1$ and $\ell \leq c_1$.
Then $x_1^\ell u \in M$ and $x_1^\ell u \diamond x_i^j \not \in M$.
$F_{\cc'}(x_1^\ell u )\setminus \{x_i^{(j)}\} = \rho(F_\cc(x_1^\ell u)\setminus \{x_i^{(j)}\}) \in \rho(\bier_\cc(M)).$
[\textit{Case 1.2}]
Suppose $i \ne 1$ and $\ell > c_1$.
Then $x_1^{c_1}x_n^{\ell-c_1} u \in M$ and $x_1^{c_1}x_n^{\ell-c_1} u \diamond x_i^j \not \in M$.
$F_{\cc'}(x_1^\ell u )\setminus \{x_i^{(j)}\} = \rho(F_\cc(x_1^{c_1}x_n^{\ell-c_1} u)\setminus \{x_i^{(j)}\}) \in \rho(\bier_\cc(M)).$
[\textit{Case 2.1}]
Suppose $i=1$ and $\ell <j \leq c_1$.
Then $x_1^\ell u \in M$ and $x_1^\ell u \diamond x_1^j \not \in M$.
$F_{\cc'}(x_1^\ell u )\setminus \{x_1^{(j)}\} = \rho(F_\cc(x_1^\ell u)\setminus \{x_1^{(j)}\}) \in \rho(\bier_\cc(M)).$
[\textit{Case 2.2}]
Suppose $i=1$ and $\ell \leq c_1 <j$.
Then $x_1^\ell u \in M$ and $x_1^{c_1} x_n^{j-c_1} u \not \in M$.
If $x_1^\ell x_n^{j-c_1} u \not \in M$ then
$F_{\cc'}(x_1^\ell u )\setminus \{x_1^{(j)}\} = \rho(F_\cc(x_1^\ell u)\setminus \{x_n^{(j-c_1)}\}) \in \rho(\bier_\cc(M)).$
If $x_1^\ell x_n^{j-c_1} u \in M$ then
$F_{\cc'}(x_1^\ell u )\setminus \{x_1^{(j)}\} = \rho(F_\cc(x_1^\ell x_n^{j-c_1} u)\setminus \{x_1^{(c_1)}\}) \in \rho(\bier_\cc(M)).$
[\textit{Case 2.3}]
Suppose $i=1$ and $ c_1 <\ell <j$.
Then $x_1^{c_1}x_n^{\ell-c_1} u \in M$ and $x_1^{c_1} x_n^{j-c_1} u \not \in M$.
$F_{\cc'}(x_1^\ell u )\setminus \{x_i^{(j)}\} = \rho(F_\cc(x_1^{c_1}x_n^{\ell-c_1} u)\setminus \{x_n^{(j-c_1)}\}) \in \rho(\bier_\cc(M)).$
\end{proof}
\begin{theorem}
\label{edgedecomposable}
For any proper $\cc$-multicomplex $M$,
$\bier_\cc(M)$ is edge decomposable.
\end{theorem}
\begin{proof}
We use induction on dimension and $n$.
If $n=1$ then the statement follows from Lemma \ref{join}.
Also, any simplicial $(d-1)$-sphere is edge decomposable if $d \leq 2$.
Suppose $n>1$.
By Lemma \ref{reduction} we may assume that $\{x_1^{(c_1)}\}$ and $\{x_n^{(0)}\}$ are in $\bier_\cc(M)$.
Then the statement follows from Lemmas \ref{linkcondition} and \ref{keylemma} together with the induction hypothesis.
\end{proof}
By \cite[Corollary 3.5]{Mu2},
Theorem \ref{edgedecomposable} implies the following fact.
\begin{corollary} \label{g}
The $g$-vector of $\bier_\cc(M)$ is the $f$-vector of a multicomplex.
\end{corollary}
We proved the above corollary by using edge decomposability.
On the other hand,
since we obtain an explicit formula of the $g$-vector of $\bier_\cc(M)$ in Theorem \ref{hvector}(ii),
it would be desirable to find a simple combinatorial proof of the above corollary
by using that formula.
For Bier spheres of simplicial complexes, Bj\"orner et al.\ \cite[Corollary 5.5]{BPSZ} gave such a proof,
and their proof can be extended to the case when $c_1=\dots=c_n$ by using Clements-Lindstr\"om theorem \cite{CL}.
However, their method seems not to be applicable when $c_1,c_2,\dots,c_n$ are not equal.
Recall that
any Bier sphere $\bier_\cc(M)$ is the boundary of a simplicial ball $\B_\cc(M)$ which is a subcomplex of the simplicial sphere $\Lambda_\cc
=\partial \widetilde X_1 * \cdots * \partial \widetilde X_n$ with the same dimension.
Corollary \ref{g} can be extended as follows.
\begin{corollary}
If $B \subset \Lambda_\cc$ is a simplicial ball with the same dimension as $\Lambda_\cc$,
then the $g$-vector of $\partial B$ is the $f$-vector of a multicomplex.
\end{corollary}
\begin{proof}
Any subcomplex of $\Lambda_\cc$ with the same dimension as $\Lambda_\cc$ is a balanced complex of type $(c_1-1,\dots,c_n-1)$ \cite{St2}.
Since $B$ is Cohen-Macaulay,
it follows from \cite[Theorem 4.4]{St2} that the $h$-vector of $B$ is equal to the $h$-vector of a $\cc$-multicomplex $M$.
Then $h(B)=h(\B_\cc(M))$ and $g(\partial B)=g(\bier_\cc(M))$.
\end{proof}
\section{An algebraic study of Bier spheres}
In this section, we study algebraic aspects of Bier spheres.
Fix $\cc=(c_1,\dots,c_n) \in \ZZ_{\geq 1}^n$.
Let $S=K[x_1,\dots,x_n]$ and $\widetilde S=K[x_{i,j}:1\leq i \leq n,\ 0 \leq j \leq c_i]$.
\subsection*{An algebraic proof of Bier's theorem}
We first introduce some basic notations on commutative algebra.
Let $I \subset S$ be a homogeneous ideal and $R=S/I$.
A sequence $\theta_1,\dots,\theta_r \in R$ is said to be an $R$-sequence if $(\theta_1,\dots,\theta_r)R \ne R$ and
$\theta_i$ is not a zero divisor of $R/(\theta_1,\dots,\theta_{i-1})$ for all $i$.
The ring $R$ is said to be \textit{Cohen-Macaulay} if there is an $R$-sequence $\theta_1,\dots,\theta_d$ of length $d$,
where $d$ is the Krull dimension of $R$.
If $R$ is Cohen-Macaulay then the number $\dim_K \{m \in R/(\theta_1,\dots,\theta_d): (x_1,\dots,x_n)m=0\}$ is independent of a choice of an $R$-sequence
$\theta_1,\dots,\theta_d$ and this number is called the \textit{type} of $R$ \cite[I ,Theorem 12.4]{St}.
We say that $R$ is \textit{Gorenstein} if it is a Cohen-Macaulay ring of type $1$.
The Gorenstein property is important in commutative algebra since it implies many nice symmetry.
The Gorenstein property also appears in several combinatorial situations in which symmetry appears.
See e.g., \cite{St3} and \cite[II \S 5]{St}.
For simplicity, we say that an ideal $I \subset S$ is Gorenstein if $S/I$ is Gorenstein.
When we consider the correspondence $\Delta \leftrightarrow K[\Delta]$,
the property that $\Delta$ is a sphere is close to the Gorenstein property of $K[\Delta]$.
Indeed, $K[\Delta]$ is Gorenstein if and only if $\Delta$ is the join of a simplex and a homology sphere
\cite[II Theorem 5.1]{St}.
In particular,
Bier's theorem shows that the ideal in $K[x_1,\dots,x_n,y_1,\dots,y_n]$ generated by
$$\{x^F : F \not \in \Delta\} \cup \{y^F: F \not \in \Delta^\vee\} \cup \{x_1y_1,\dots,x_ny_n\}$$
is Gorenstein.
Although Bier's proof is simple (see \cite[pp.\ 112--116]{Ma}),
it is natural to ask if there is an algebraic way to prove that the above ideal is Gorenstein.
Linkage theory gives such a proof.
We will not explain details on linkage theory,
since we only need the following fact:
Let $Q \subset S$ be a Gorenstein ideal,
$L \subset S$ a Cohen-Macaulay ideal which contains $Q$
and $L'=Q:L=\{f \in S :f L \subset Q\}$.
Suppose that $Q$ and $L$ are radical ideals and $\dim S/L=\dim S/Q$.
Then $S/(L+L')$ is Gorenstein.
(This fact is an immediate consequence of \cite[Theorem 4.2.1 and Proposition 5.2.2(c)]{Mig}.)
The following simple result due to Miller
gives a connection between linkage theory and Alexander duality for monomial ideals defined in Definition \ref{alex}.
\begin{lemma}[{Miller \cite[Theorem 2.1]{Mi2}}]
\label{mlinkage}
Let $P=(x_1^{c_1+1},\dots,x_n^{c_n+1})$
and let $I \subset S$ be a $\cc$-ideal.
Then $P:(I+P)=I^\vee+P$.
\end{lemma}
\begin{lemma} \label{linkage}
For a $\cc$-ideal $I \subset S$, one has
$$\pol(P):\pol(I+P)=\pol^*(I^\vee+P).$$
\end{lemma}
\begin{proof}
For any monomial $w \in \widetilde S$,
$w \in \pol(P):\pol(I+P)$ if and only if, for any $x^\aaa \in G(I+P)$,
$\pol(x^\aaa) w$ is divisible by $x_{i,0}x_{i,1}\cdots x_{i,c_i}$ for some $i$.
Then, since $\pol(x^\aaa)$ is a monomial of the form $ \prod_{a_i \ne 0} x_{i,0}x_{i,1} \cdots x_{i,a_i-1}$,
any generator $w \in G(\pol(P):\pol(I+P))$ must be of the form $w=\pol^*(x^\bb)$ for some $x^\bb \in S$.
On the other hand, for $\cc$-monomials $x^\aaa, x^\bb \in S$,
$\pol(x^\aaa)\pol^*(x^\bb) \in \pol(P)$ if and only if
$\pol(x^\aaa)\pol^*(x^\bb)$ is divisible by $x_{i,0}x_{i,1}\cdots x_{i,c_i}$ for some $i$,
equivalently, $x^\aaa x^\bb$ is divisible by $x_i^{c_i+1}$.
Thus $\pol(x^\aaa)\pol^*(x^\bb) \in \pol(P)$
if and only if $x^\aaa x^\bb \in P$.
Hence $\pol^*(x^\bb) \in G(\pol(P):\pol(I+P))$ if and only if $x^\bb \in G(P:(I+P))$.
Then the statement follows from Lemma \ref{mlinkage}.
\end{proof}
Since $\pol(P)$
is generated by an $S$-sequence,
$\widetilde S/\pol(P)$ is a Gorenstein ring of dimension $|\cc|$.
Since the dimension of $S/(I+P)$ is $0$ and since taking polarizations preserves the Cohen-Macaulay property,
$\widetilde S/\pol(I+P)$ is a Cohen-Macaulay ring of dimension $|\cc|$.
Then the standard fact in linkage theory mentioned before Lemma \ref{mlinkage}
gives a purely algebraic proof of the following statement (apply the case when $Q=\pol(P)$ and $L=\pol(I+P)$).
\begin{theorem}
For any $\cc$-ideal $I \subset S$,
the ideal $\pol(I)+\pol^*(I^\vee)+\pol(P) \subset \widetilde S$ is Gorenstein.
\end{theorem}
\begin{remark}
By Lemmas \ref{complement} and \ref{linkage},
the Stanley-Reisner complex of $I_{\Lambda_\cc}:I_{\B_\cc(M)}=\pol(P): I_{\B_\cc(M)}$
is the complementary ball of $\B_\cc(M)$ in $\Lambda_\cc$.
This is the special case of the following general fact:
Let $\Gamma$ be a $d$-dimensional simplicial sphere on $[n]$,
$B \subset \Gamma$ a $d$-dimensional ball and
$B^c=\langle F \subset [n] : F \in \Gamma \setminus B \rangle$.
Then $I_\Gamma: I_B = I_{B^c}$.
\end{remark}
\subsection*{Graded Betti numbers and Bier spheres}
An important application of polarization appears in the study of graded Betti numbers.
Computations of the (multi) graded Betti numbers are one of current trends in combinatorial commutative algebra.
In the rest of this section, we study graded Betti numbers of Bier spheres.
We refer the readers to \cite{MS} for basics on multigraded commutative algebra.
Let $I \subset S$ be a monomial ideal.
Then $\Tor_i(S/I,K)$ is $\ZZ_{\geq 0}^n$-graded.
The integers
$$\beta_{i,(a_1,\dots,a_n)}=\dim_K \Tor_i(S/I,K)_{(a_1,\dots,a_n)}$$
are called the \textit{multigraded Betti numbers of $S/I$},
where $M_{(a_1,\dots,a_n)}$ denotes the homogeneous component of a $\ZZ_{\geq 0}^n$-graded $S$-module $M$ of degree $(a_1,\dots,a_n)$.
In the rest of this paper,
we identify $\aaa\in \ZZ_{\geq 0}^n$ with $x^\aaa$ for convenience.
The numbers $\beta_{i,j}(S/I)=\sum_{\deg x^\aaa=j} \beta_{i,x^\aaa}(S/I)$ and $\beta_i(S/I)=\sum_{j} \beta_{i,j}(S/I)$
are called \textit{graded Betti numbers} and \textit{total Betti numbers} of $S/I$ respectively.
The next result is useful to study graded Betti numbers of $K[\bier_\cc (M)]$.
\begin{lemma} \label{cone}
Let $B$ be a $(d-1)$-dimensional simplicial ball on $X=\{x_1,\dots,x_n\}$,
and let $\Delta =\partial B$ be the boundary complex of $B$.
If $B$ is a cone
then for any $i$ and for any squarefree monomial $x^F = \prod_{j \in F} x_j\in S$,
$$\beta_{i,x^F} (S/I_\Delta) = \beta_{i,x^F}(S/I_B) + \beta_{n+1-d-i, x^{[n] \setminus F}}(S/I_B).$$
In particular, $\beta_{i,j} (S/I_{\Delta}) = \beta_{i,j}(S/I_B) + \beta_{n+1-d-i, n-j}(S/I_B)$
for all $i,j$.
\end{lemma}
\begin{proof}
Since multigraded Betti numbers of Stanley-Reisner rings are concentrated in squarefree degrees (see \cite[Corollary 5.12]{MS}),
it is enough to prove the first statement.
The short exact sequence
$$0 \longrightarrow I_{\Delta}/I_B \longrightarrow S/I_B \longrightarrow S/I_{\Delta} \longrightarrow 0$$
yields the long exact sequence
\begin{align*}
&\cdots \longrightarrow \Tor_i(I_\Delta/I_B,K)_{x^F}
\longrightarrow \Tor_i(S/I_B,K)_{x^F}
\longrightarrow \Tor_i(S/I_\Delta,K)_{x^F} \\
&\hspace{18pt}
\longrightarrow \Tor_{i-1}(I_\Delta/I_B,K)_{x^F}
\longrightarrow \Tor_{i-1}(S/I_B,K)_{x^F}
\longrightarrow \cdots.
\end{align*}
If $B$ is a cone w.r.t. the vertex $x_k$ then all facets of $B$ contain $x_k$.
Then $G(I_B)$ has no monomials which are divisible by $x_k$.
Thus
\begin{align}
\label{3.1}
\Tor_i(S/I_B,K)_{x^F}=0 \mbox{ if } k \in F.
\end{align}
On the other hand,
since $I_\Delta/I_B$ is the canonical module of $S/I_B$ \cite[II Theorem 7.3]{St},
we have
\begin{align} \label{3.2}
\Tor_i(I_\Delta/I_B,K)_{x^F}=\Tor_{n-d-i}(S/I_B,K)_{x^{[n]\setminus F}}
\end{align}
(see \cite[Theorem 13.37]{MS}) and
\begin{align} \label{3.3}
\Tor_i(I_\Delta/I_B,K)_{x^F}=0 \mbox{ if } k \not \in F.
\end{align}
Then by applying \eqref{3.1}, \eqref{3.2} and \eqref{3.3} to the long exact sequence we have
$$
\beta_{i,x^F}(S/I_\Delta)= \beta_{i,x^F}(S/I_B)
=\beta_{i,x^F}(S/I_B) + \beta_{n+1-d-i, x^{[n] \setminus F}}(S/I_B)
\ \mbox{ if } k \not \in F$$
and
$$
\beta_{i,x^F}(S/I_\Delta)= \beta_{n+1-d-i, x^{[n] \setminus F}}(S/I_B)
=\beta_{i,x^F}(S/I_B) + \beta_{n+1-d-i, x^{[n] \setminus F}}(S/I_B)
\ \mbox{ if } k \in F,$$
as desired.
\end{proof}
Let $M$ be a $\cc$-multicomplex and $\mathrm{Lcm}(M)$ the least common multiple of monomials in $M$.
As we see in Remark \ref{r1}, if $\mathrm{Lcm}(M) \ne x^\cc$ then $\B_\cc(M)$ is a cone.
Since taking polarizations does not change graded Betti numbers,
Lemma \ref{cone} yields the next corollary.
\begin{corollary} \label{BettiNumber}
Let $M$ be a proper $\cc$-multicomplex on $X$.
If $x^\cc \ne \mathrm{Lcm}(M)$
then
$$\beta^{\widetilde S}_{i,j}(\widetilde S/I_{\bier_\cc (M)}) = \beta_{i,j}^S(S/I(M)) + \beta^S_{n+1-i,|\overline \cc|-j}(S/I(M))$$
for all $i$ and $j$, where $\beta^{\widetilde S}_{i,j}(\widetilde S/I_{\bier_\cc (M)})$ are the graded Betti numbers over $\widetilde S$.
\end{corollary}
Note that the above formula does not always hold if $\mathrm{Lcm}(M) = x^\cc$.
For example, Example \ref{3..4} does not satisfy the formula.
Although we need an assumption on $\mathrm{Lcm}(M)$,
Corollary \ref{BettiNumber} will be useful to find many examples of graded Betti numbers of Gorenstein Stanley-Reisner rings.
For example, the above corollary implies the following non-trivial result.
\begin{corollary}
Let $I \subset S$ be a monomial ideal such that $S/I$ has finite length,
and let $b_i=\beta_i(S/I)$ for all $i$.
Then there exists a simplicial complex $\Delta$ such that
$K[\Delta]$ is Gorenstein and $\beta_i(K[\Delta])=b_i+b_{n+1-i}$ for all $i$.
\end{corollary}
\begin{proof}
If $S/I$ has finite length then there exists a finite multicomplex $M$ such that
$I=I(M)$. Choose a sufficiently large $\cc \in \ZZ_{\geq 0}^n$.
Corollary \ref{BettiNumber} says $\beta_i(K[\bier_\cc(M)])\!=b_i+b_{n+1-i}$ for all $i$.
\end{proof}
\begin{example} \label{BettiTable}
To understand the meaning of Corollary \ref{BettiNumber},
it is convenient to consider Betti tables (tables whose $i,j$-th entry is $\beta_{i,i+j}$).
Let $\cc=(2,2,2)$ and $M=\{1,x,y,z,xz,yz,z^2,yz^2\}$.
Then we have $I(M)=(x^2,y^2,z^3,xy,xz^2)$ and the Betti table of $K[x,y,z]/I(M)$
computed by the computer algebra system Macaulay 2 \cite{GS} is
\begin{verbatim}
total: 1 5 6 2
0: 1 . . .
1: . 3 2 .
2: . 2 3 1
3: . . 1 1
\end{verbatim}
Corollary \ref{BettiNumber} says that
the Betti table of $K[\bier_\cc(M)]$ is the sum of the Betti table
of $K[x,y,z]/I(M)$ and the table which is obtained by transposing this table.
Thus the Betti table of $K[\bier_\cc(M)]$ is computed as follows:\medskip
\begin{verbatim}
total: 1 5 6 2 . total: . 2 6 5 1 total: 1 7 12 7 1
0: 1 . . . . 0: . . . . . 0: 1 . . . .
1: . 3 2 . . 1: . . . . . 1: . 3 2 . .
2: . 2 3 1 . + 2: . 1 1 . . = 2: . 3 4 1 .
3: . . 1 1 . 3: . 1 3 2 . 3: . 1 4 3 .
4: . . . . . 4: . . 2 3 . 4: . . 2 3 .
5: . . . . . 5: . . . . 1 5: . . . . 1
\end{verbatim}
\end{example}
\bigskip
\noindent
\textbf{Acknowledgments}:
This work was supported by KAKENHI 09J00756.
I would like to thank the referees for careful readings
and many helpful comments.
|
\section{Introduction and results}
The dynamics of the D-branes of type II superstring theories is well-approximated by the effective world-volume field theories which consist of the sum of Dirac-Born-Infeld (DBI) and Chern-Simons (CS) actions.
The DBI action describes the dynamics of the brane in the presence of the NSNS background fields. For constant background fields it can be found by requiring the consistency with nonlinear T-duality \cite{Leigh:1989jq,Bachas:1995kx}
\begin{eqnarray}
S_{DBI}&=&-T_p\int d^{p+1}x\,e^{-\phi}\sqrt{-\det\left(G_{ab}+B_{ab}+2\pi\alpha'F_{ab}\right)}
\end{eqnarray}
where $G_{ab}$ and $B_{ab}$ are the pulled back of the bulk fields $G_{\mu\nu}$ and $B_{\mu\nu}$ onto the world-volume of D-brane\footnote{Our index conversion is that the Greek letters $(\mu,\nu,\cdots)$ are the indices of the space-time coordinates, the Latin letters $(a,d,c,\cdots)$ are the world-volume indices and the letters $(i,j,k,\cdots)$ are the normal bundle indices.}. The curvature corrections to this action has been found in \cite{Bachas:1999um} by requiring consistency of the effective action with the $O(\alpha'^2)$ terms of the corresponding disk-level scattering amplitude \cite{Garousi:1996ad,Hashimoto:1996kf}. The on-shell ambiguity of these couplings has been removed in \cite{Garousi:2009dj} by requiring the consistency of the couplings with linear T-duality. Moreover, this consistency fixes the couplings of non-constant dilaton and B-field at the order $O(\alpha'^2)$ in the action which are reproduced by the corresponding disk level scattering amplitude. In particular, it has been found in \cite{Garousi:2009dj} that the consistency with T-duality/S-matrix requires the non-constant dilaton appears in the string frame action only as the overall factor of $e^{-\phi}$.
The CS part on the other hand describes the coupling of D-branes to the RR fields. For constant background fields it is given by \cite{Polchinski:1995mt,Douglas:1995bn}
\begin{eqnarray}
S_{CS}&=&T_{p}\int_{M^{p+1}}e^{B}C\labell{CS2}
\end{eqnarray}
where $M^{p+1}$ represents the world volume of the D$_p$-brane, $C$ is meant to represent a sum over all appropriate RR potential forms and the multiplication rule is the wedge product. The abelian gauge field can be added to the action as $B\rightarrow B+2\pi\alpha'F$. The curvature corrections to this action has been found
by requiring that the chiral anomaly on the world volume of intersecting D-branes (I-brane) cancels with the anomalous variation of the CS action \cite{Green:1996dd,Cheung:1997az,Minasian:1997mm}. This correction is
\begin{eqnarray}
S_{CS}&=&T_{p}\int_{M^{p+1}}e^BC\left(\frac{{\cal A}(4\pi^2\alpha'R_T)}{{\cal A}(4\pi^2\alpha'R_N)}\right)^{1/2}\labell{CS}
\end{eqnarray}
where ${\cal A}(R_{T,N})$ is the Dirac roof genus of the tangent and normal bundle curvatures respectively,
\begin{eqnarray}
\sqrt{\frac{{\cal A}(4\pi^2\alpha'R_T)}{{\cal A}(4\pi^2\alpha'R_N)}}&=&1+\frac{(4\pi^2\alpha')^2}{384\pi^2}(\tr R_T^2-\tr R_N^2)+\cdots \labell{roof}
\end{eqnarray}
For totally-geodesic embeddings of world-volume in the ambient spacetime, ${\rm R}_{T,N}$ are the pulled back curvature 2-forms of the tangent and normal bundles respectively (see the appendix in ref.
\cite{Bachas:1999um} for more details).
It has been pointed out in \cite{Myers:1999ps} that the anomalous CS couplings \reef{CS} must be incomplete for non-constant B-field as they are not compatible with T-duality. T-duality exchanges the components of the metric and the B-field whereas the couplings \reef{CS} includes only the curvature terms. Compatibility of this action with T-duality should give a bunch of new couplings \cite{Ga, kg}.
In this paper we would like to show that for non-constant RR and NSNS fields there are other contribution to the action \reef{CS} at order $O(\alpha'^2)$ which may not arise from requiring the consistency of the action \reef{CS} with T-duality. These terms which involve linear NSNS field can be found by studying the S-matrix element of one RR and one NSNS vertex operators \cite{Garousi:1996ad} and by requiring them to be consistent with linear T-duality. We will find the following string frame couplings at order $O(\alpha'^2)$:
\begin{eqnarray}
S_{}&\sim&T_p\int d^{p+1}x\,\epsilon^{a_0\cdots a_p}\left(\frac{1}{2!(p-1)!}[F^{(p)}_{ia_2\cdots a_p,a}H_{a_0a_1}{}^{a,i}-\frac{1}{p}F^{(p)}_{a_1a_2\cdots a_p,i}(H_{a_0a}{}^{i,a}-H_{a_0j}{}^{i,j})]\right.\nonumber\\
&&\left.\qquad\qquad\qquad\qquad+\frac{2}{p!}[\frac{1}{2!}F^{(p+2)}_{ia_1\cdots a_pj,a}\cR^a{}_{a_0}{}^{ij}-\frac{1}{p+1}F^{(p+2)}_{a_0\cdots a_pj,i}\hat{\cR}^{ij}]\right.\nonumber\\
&&\left.\qquad\qquad\qquad\qquad-\frac{1}{3!(p+1)!}F^{(p+4)}_{ia_0\cdots a_pjk,a}H^{ijk,a}\right)\labell{LTdual}
\end{eqnarray}
where as usual commas denote partial differentiation.
It has been shown in \cite{Garousi:2009dj} that the compatibility of the curvature corrections to the DBI action with linear T-duality transformations
requires the non-constant dilaton appears in the string frame action only through the overall factor of $e^{-\phi}$. This factor has been absorbed in the RR field so one expects that the dilaton appears in the above action only through the string frame metric. We will show that the coupling of one $F^{(p+2)}$ and one dilaton in the Einstein frame which can be calculated by the S-matrix element, is reproduced exactly by transforming the couplings in the second line above to the Einstein frame.
The couplings in \reef{LTdual} have been found by the S-matrix element of one RR and one NSNS vertex operators and by T-duality. The S-matrix method produces the on-shell couplings and consistency of the couplings with linear T-duality then fixes the on-shell ambiguity of the couplings. Correction to this action can also be found by requiring it to be consistent with nonlinear T-duality transformations. We will consider one particular nonlinear term in the T-duality transformation of the RR field and then examine the consistency of \reef{LTdual} with it to find new couplings. The new couplings are given by the above action in which $F^{(n)}$ is replaced by ${\cal F}^{(n)}$ where
\begin{eqnarray}
{\cal F}^{(n)}&=&F^{(n)}+B\wedge F^{(n-2)}+H\wedge C^{(n-3)}\labell{newF}\\
&&+\frac{1}{2!}B\wedge B\wedge F^{(n-4)}+\frac{1}{2!}B\wedge H\wedge C^{(n-5)}+\frac{1}{2!}H\wedge B\wedge C^{(n-5)}+\cdots\nonumber\\
&=&d{\cal C}^{(n)}\nonumber
\end{eqnarray}
where ${\cal C}=e^{B}C$, is the RR potential in the CS action \reef{CS2}.
An outline of the paper is as follows: We begin the section 2 by writing the S-matrix element of one RR and one NSNS vertex operators. From the contact terms of this amplitude at order $(\alpha')^2$, we will find the on-shell couplings of one massless RR and two NSNS fields. In section 3, we review the T-duality transformations and the strategy for checking the consistency of a D-brane action with T-duality. In section 3.1, we check the consistency of the couplings found in section 2 with linear T-duality which fixes the on-shell ambiguity of the couplings. After fixing the on-shell ambiguity of the gravity couplings, we show that the dilaton appears in the action only through the string frame metric. In section 3.2, we check the consistency of the couplings \reef{LTdual} with nonlinear T-duality and show that the field strength in the action \reef{LTdual} should be given by \reef{newF}.
\section{Scattering amplitudes}
A method for finding the couplings in effective field theory is the S-matrix method. The standard CS coupling \reef{CS2} has been confirmed by the S-matrix method in {\it e.g.,}\ \cite{Garousi:1998bj,Dymarsky:2002kk}. The couplings of NSNS and RR fluxes to various types of D-branes have been found in \cite{Billo':2008sp} by evaluating disk amplitudes among two open string and one closed string vertex operators. To find the couplings of one RR and one NSNS states to D$_p$-brane, one needs
the scattering amplitude of their corresponding vertex operators which is given by \cite{Garousi:1996ad}
\begin{eqnarray}
A(\veps_1,p_1;\veps_2,p_2)&=&-\frac{1}{8}T_p\alpha'^2K(1,2)\frac{\Gamma(-\alpha' t/4)\Gamma(\alpha' q^2)}{\Gamma(1-\alpha't/4+\alpha' q^2)}\nonumber\\
&=&\frac{1}{2}T_pK(1,2)\left(\frac{1}{q^2t}+\frac{\pi^2\alpha'^2}{24}+O(\alpha'^4)\right)\labell{Amp}
\end{eqnarray}
where $q^2=p_1^ap_1^b\eta_{ab}$ is the momentum flowing along the world-volume of D-brane, and $t=-(p_1+p_2)^2$ is the momentum transfer in the transverse direction. The kinematic factor is
\begin{eqnarray}
K(1,2)&=&i\frac{q^2}{\sqrt{2}}\Tr(P_-\Gamma_{1(n)}M_p\gamma^{\nu}\gamma\!\cdot\!(p_1+p_2)\gamma^{\mu})(\veps_2\!\cdot\! D)_{\mu\nu}\labell{kin}\\
&&-i\frac{t}{2\sqrt{2}}[\Tr(P_-\Gamma_{1(n)}M_p\gamma\!\cdot\! D\!\cdot\!\veps_2^T\!\cdot\! D\!\cdot\! p_2)-\Tr(P_-\Gamma_{1(n)}M_p\gamma\!\cdot\!\veps_2\!\cdot\! D\!\cdot\! p_2)\nonumber\\
&&\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad-\Tr(P_-\Gamma_{1(n)}M_p\gamma\!\cdot\! D\!\cdot\! p_2)\Tr(\veps_2\!\cdot\! D)]\nonumber
\end{eqnarray}
where the matrix $D^{\mu}_{\nu}$ is diagonal with +1 in the world volume directions and -1 in the transverse directions, and
\begin{eqnarray}
\Gamma_{1(n)}&=&\frac{1}{n!}F_{1\nu_1\cdots\nu_n}\gamma^{\nu_1}\cdots\gamma^{\nu_n}\nonumber\\
M_p&=&\frac{\pm 1}{(p+1)!}\epsilon_{a_0\cdots a_p}\gamma^{a_0}\cdots\gamma^{a_p}
\end{eqnarray}
where $F_1$ is the linearized RR field strength $n$-form and $\epsilon$ is the volume $p+1$-form of the $D_p$-brane. In equation \reef{kin}, $P_-=\frac{1}{2}(1-\gamma_{11})$ is the chiral projection operator and $\veps_2$ is the NS-NS polarization. The $\gamma_{11}$ in the chiral projection gives the magnetic couplings and $1$ gives the electric couplings. The first term in \reef{Amp} produces the massless poles resulting from the $(\alpha')^0$ order of the DBI and CS couplings on the D-brane, and the supergravity couplings in the bulk. The second term in \reef{Amp} should produce $(\alpha')^2$ couplings of one RR and one NSNS on the D-brane in which we are interested.
Using the identity $M_p\gamma^{\mu}=D^{\mu}{}_{\nu}\gamma^{\nu}M_p$ \cite{Garousi:1996ad} and the algebra $\{\gamma^{\mu},\gamma^{\nu}\}=-2\eta^{\mu\nu}$, one can write the above kinematic factor for the electric couplings as
\begin{eqnarray}
K(1,2)&\!\!\!=\!\!\!&i\frac{q^2}{2\sqrt{2}}\Tr(\Gamma_{1(n)}\gamma^{\nu}M_p\gamma\!\cdot\!(p_1+p_2)\gamma^{\mu})(\veps_2)_{\mu\nu}\labell{kin2}\\
&&-i\frac{t}{4\sqrt{2}}
-\Tr(\Gamma_{1(n)}\gamma\!\cdot\! D\!\cdot\!\veps_2\!\cdot\! D\!\cdot\! p_2M_p)\nonumber\\
&&-\frac{1}{2}\Tr(\Gamma_{1(n)}\gamma^{\mu}M_p\gamma\!\cdot\! D\!\cdot\! p_2\gamma^{\nu})(D\!\cdot\!\veps_2\!\cdot\! D+\veps^T_2)_{\mu\nu}]\nonumber
\end{eqnarray}
One can easily check that the kinematic factor is zero for $n\le p-2$ and for $n> p+4$. This factor is non-zero for $n=p,\, n=p+2$ and for $n=p+4$. Let us consider each case separately.
\subsection{$n=p$ case}
For $n=p$ case, one needs to perform the following traces:
\begin{eqnarray}
\Tr(\gamma^{\mu_1}\cdots\gamma^{\mu_p}\gamma^{\mu}\gamma^{a_0}\cdots\gamma^{a_p}\gamma^{\alpha}\gamma^{\nu})&{\rm and}& \Tr(\gamma^{\mu_1}\cdots\gamma^{\mu_p}\gamma^{\alpha}\gamma^{a_0}\cdots\gamma^{a_p})
\end{eqnarray}
They make various contraction of the indices. The first one simplifies to
\begin{eqnarray}
&&(-1)^p[\eta^{\mu\nu}\eta^{\alpha a_0}\eta^{\mu_1a_1}+\eta^{\alpha\nu}\eta^{a_0\mu}\eta^{a_1\mu_1}-\eta^{\alpha\mu}\eta^{a_0\nu}\eta^{a_1\mu_1}-p\eta^{\alpha\mu_1}\eta^{a_0\mu}\eta^{a_1\nu}\nonumber\\
&&-p\eta^{\alpha a_0}\eta^{\mu\mu_1}\eta^{a_1\nu}-p\eta^{\alpha a_0}\eta^{a_1\mu}\eta^{\nu\mu_1}]
p(p+1)\Tr(\gamma^{\mu_2}\cdots\gamma^{\mu_p}\gamma^{a_2}\cdots\gamma^{a_p})
\end{eqnarray}
The second trace simplifies to
\begin{eqnarray}
\Tr(\gamma^{\mu_1}\cdots\gamma^{\mu_p}\gamma^{\alpha}\gamma^{a_0}\cdots\gamma^{a_p})&=&(-1)^p\eta^{\alpha a_0}\eta^{\mu_1 a_1}p(p+1)\Tr(\gamma^{\mu_2}\cdots\gamma^{\mu_p}\gamma^{a_2}\cdots\gamma^{a_p})
\end{eqnarray}
and the trace $\Tr(\gamma^{\mu_2}\cdots\gamma^{\mu_p}\gamma^{a_2}\cdots\gamma^{a_p})$ causes the RR field strength to contract with the volume form $\epsilon$, {\it i.e.,}\ $F_{1\mu_1a_2\cdots a_p}\epsilon^{a_0\cdots a_p}$.
Using the above traces, one finds that the kinematic factor \reef{kin2} for the graviton is
\begin{eqnarray}
K(1,2)&\sim&-i\frac{t}{2\sqrt{2}}[F_{1a_1\cdots a_p}(\veps_2)_{a_0}{}^ap_{2a}+pF_{1aa_2\cdots a_p}(\veps_2)_{a_1}{}^ap_{2a_0}]\epsilon^{a_0\cdots a_p}\labell{K12}
\end{eqnarray}
Using the fact that the indices $a_0,\cdots, a_p$ contracted with the totally antisymmetric $\epsilon^{a_0\cdots a_p}$ tenser and the conservation of the momentum $p_{1a}+p_{2a}=0$, one can write
\begin{eqnarray}
pF_{1aa_2\cdots a_p}p_{2a_0}&=&p\,p_{1a}C_{1a_2\cdots a_p}p_{2a_0} \nonumber\\
&=&F_{1a_0a_2\cdots a_p}p_{2a}\labell{iden1}
\end{eqnarray}
which makes the kinematic factor \reef{K12} to be zero. For B-field, one finds the following non-zero result for the kinematic factor \reef{kin2}:
\begin{eqnarray}
K(1,2)&\sim &-\left(i\frac{q^2}{2\sqrt{2}}[2F_{1a_1a_2\cdots a_p}(\veps_2)_{a_0i}(p_1+p_2)^i-pF_{1ia_2\cdots a_p}(\veps_2)_{a_0a_1}(p_1+p_2)^i]\right.\nonumber\\
&&\left.-i\frac{t}{\sqrt{2}}F_{1a_1a_2\cdots a_p}(\veps_2)_{a_0i}p_2^i\right)\epsilon^{a_0a_1\cdots a_p}
\end{eqnarray}
As a check of the calculation, if one replaces the B-field polarization with $(\veps_2)_{\mu\nu}\rightarrow \zeta_{\mu}(p_2)_\nu-\zeta_{\nu}(p_2)_\mu$ the kinematic factor vanishes, as expected from the Ward identity.
To find the field theory couplings corresponding to the above momentum space contact terms, we use the following identities:
\begin{eqnarray}
F_{1a_1\cdots a_p}(\veps_2)_{a_0a}&=&\frac{p}{2}F_{1aa_2\cdots a_p}(\veps_2)_{a_0a_1}\nonumber\\
F_{1a_1\cdots a_p}(p_1)_i&=&pF_{1ia_2\cdots a_p}(p_1)_{a_1}
\end{eqnarray}
where we have used the fact that the indices $a_0,\cdots, a_p$ contracted with the totally antisymmetric $\epsilon^{a_0\cdots a_p}$ tenser. Using these identities one can write the kinematic factor as
\begin{eqnarray}
K(1,2)&\sim &-\frac{ip}{\sqrt{2}}\left((\veps_2)_{a_0a_1}[-\frac{1}{2}p_1\!\cdot\! V\!\cdot\! p_1F_{1ia_2\cdots a_p}(p_2)^i-p_1\!\cdot\! N\!\cdot\! p_2F_{1aa_2\cdots a_p}(p_2)^a]\right.\nonumber\\
&&\left.+p_1\!\cdot\! V\!\cdot\! p_1F_{1ia_2\cdots a_p}(p_1)_{a_1}(\veps_2)_{a_0}{}^i\right)\epsilon^{a_0a_1\cdots a_p}
\end{eqnarray}
which satisfies the Ward identity. The couplings corresponding to the above terms are:
\begin{eqnarray}
\frac{T_p}{2!(p-1)!}\int d^{p+1}x\,\epsilon^{a_0a_1\cdots a_p}\left(F^{(p)}_{ia_2\cdots a_p,a}H_{a_0a_1}{}^{a,i}- F^{(p)}_{aa_2\cdots a_p,i}H_{a_0a_1}{}^{i,a}\right)\labell{first}
\end{eqnarray}
where
\begin{eqnarray}
H_{\mu\nu\alpha}&=&B_{\mu\nu,\alpha}+B_{\alpha\mu,\nu}+B_{\nu\alpha,\mu}
\end{eqnarray}
The other terms in \reef{Amp} correspond to the higher derivative of the couplings \reef{first} in which we are not interested in this paper.
The last coupling in \reef{first} has on-shell ambiguity. To see this we note that the index $a$ in this term can be either $a_0$ or $a_1$. If $a=a_0$, it can be written as $-2F_{a_0a_2\cdots a_p,i}H_{aa_1}{}^{i,a}/p$, and if $a=a_1$, it can be written as $-2F_{a_1a_2\cdots a_p,i}H_{a_0a}{}^{i,a}/p$. Interchanging $a_1\leftrightarrow a_0$ in the latter expression and using the fact that it has the overall factor of the volume form, one can write it as the former expression. Hence, the last term in \reef{first} can be written as $-2F_{a_1a_2\cdots a_p,i}H_{a_0a}{}^{i,a}/p$. Moreover, using the on-shell condition $H_{\nu\rho\alpha,\mu}{}^{\mu}=0$, one can write it as $2F_{a_1a_2\cdots a_p,i}H_{a_0j}{}^{i,j}/p$ or as
\begin{eqnarray}
-\frac{1}{p}F_{a_1\cdots a_p,i}(H_{a_0a}{}^{i,a}-H_{a_0j}{}^{i,j})
\end{eqnarray}
We will fix the above on-shell ambiguity in section 3 by requiring the consistency of the coupling with the T-duality transformations.
\subsection{$n=p+2$ case}
For $n=p+2$ case, the traces in \reef{kin2} simplify to
\begin{eqnarray}
\Tr(\gamma^{\mu_1}\cdots\gamma^{\mu_{p+2}}\gamma^{\mu}\gamma^{a_0}\cdots\gamma^{a_p}\gamma^{\alpha}\gamma^{\nu})&\!\!\!=\!\!\!&(p+1)(p+2)\Tr(\gamma^{\mu_3}\cdots\gamma^{\mu_{p+2}}\gamma^{a_1}\cdots\gamma^{a_p})\times\nonumber\\
&&[-\eta^{\mu\nu}\eta^{\alpha\mu_1}\eta^{a_0\mu_2}+\eta^{\alpha\nu}\eta^{a_0\mu_2}\eta^{\mu\mu_1}+\eta^{\alpha\mu}\eta^{\mu_1\nu}\eta^{a_0\mu_2}+\nonumber\\
&&(p+1)(\eta^{\alpha\mu_1}\eta^{\mu_2\mu}\eta^{a_0\nu}+\eta^{\alpha \mu_1}\eta^{\mu a_0}\eta^{\mu_2\nu}+\eta^{\alpha a_0}\eta^{\mu_1\mu}\eta^{\nu\mu_2})]\nonumber\\
\Tr(\gamma^{\mu_1}\cdots\gamma^{\mu_{p+2}}\gamma^{\alpha}\gamma^{a_0}\cdots\gamma^{a_p})&=&-\eta^{\alpha \mu_1}\eta^{\mu_2 a_0}(p+1)(p+2)\Tr(\gamma^{\mu_3}\cdots\gamma^{\mu_{p+2}}\gamma^{a_1}\cdots\gamma^{a_p})\nonumber
\end{eqnarray}
The trace $\Tr(\gamma^{\mu_3}\cdots\gamma^{\mu_{p+2}}\gamma^{a_1}\cdots\gamma^{a_p})$ causes the RR field strength to contract with the volume form as $F_{i\mu_1\mu_2a_1\cdots a_p}\epsilon^{a_0\cdots a_p}$. Inserting these traces in \reef{kin2}, one finds
the kinematic factor for B-field becomes
\begin{eqnarray}
K(1,2)&\sim & -i\frac{t}{2\sqrt{2}}[-F_{1ia_0\cdots a_p}(\veps_2)^i{}_a p_{2a}+(p+1)F_{1iaa_1\cdots a_p}(\veps_2)^i{}_ap_{2a_0}]\epsilon^{a_0\cdots a_p}\labell{K122}
\end{eqnarray}
Using the fact that the indices $a_0,\cdots, a_p$ contract with the totally antisymmetric tensor $\epsilon^{a_0\cdots a_p}$ and the conservation of the momentum $p_{1a}+p_{2a}=0$, one can write
\begin{eqnarray}
(p+1)F_{1iaa_1\cdots a_p}p_{2a_0}\ell^a
&=&F_{1ia_0a_1\cdots a_p}p_{2a}\ell^a\labell{iden2}
\end{eqnarray}
for any vector $\ell^a$. This makes the kinematic factor \reef{K122} to be zero. For the symmetric polarization, graviton or dilaton, one finds the kinematic factor \reef{kin2} to be
\begin{eqnarray}
K(1,2)&\sim &\left(i\frac{q^2}{\sqrt{2}}[F_{1ja_0\cdots a_p}(\veps_2)^j{}_i(p_1+p_2)^i+(p+1)F_{1i\mu a_1\cdots a_p}(\veps_2)^{\mu}{}_{a_0}(p_1+p_2)^i]\right.\nonumber\\
&&+i\frac{t}{2\sqrt{2}}[2F_{1ia_0\cdots a_p}(\veps_2)^{ia}p_{2a}-(p+1)F_{1aa_1\cdots a_p}(\veps_2)^a{}_{a_0}p_2^i]\nonumber\\
&&\left.-\frac{i}{2\sqrt{2}}F_{1ia_0\cdots a_p}p_2^i(q^2-\frac{t}{2})\Tr(\veps_2)\right)\epsilon^{a_0a_1\cdots a_p}\labell{kin3}
\end{eqnarray}
The last term is zero for graviton, but is has contribution to the dilaton amplitude. Using the identity \reef{iden2}, one can write the above equation for the graviton as
\begin{eqnarray}
K(1,2)&\sim &\frac{i(p+1)}{\sqrt{2}}\left(F_{1jaa_1\cdots a_p}p_{2a_0}p_2^a(\veps_2)^j{}_ip_1^i-p_1\!\cdot\! V\!\cdot\! p_2F_{1ija_1\cdots a_p}(\veps_2)^{j}{}_{a_0}p_2^i\right.\nonumber\\
&&\left.+p_1\!\cdot\! N\!\cdot\! p_2[F_{1iaa_1\cdots a_p}(\veps_2)^{a}{}_{a_0}p_{2}^i-2F_{1iaa_1\cdots a_p}(\veps_2)^{ai}p_{2a_0}]\right)\epsilon^{a_0a_1\cdots a_p}
\end{eqnarray}
It satisfies the Ward identity. Using the identity
\begin{eqnarray}
F_{1jaa_1\cdots a_p}p_{1i}p_{2a_0}&=&F_{1iaa_1\cdots a_p}p_{1j}p_{2a_0}-F_{1ija_1\cdots a_p}p_{1a}p_{2a_0}\,,
\end{eqnarray}
one finds that the field theory corresponding to the above amplitude is
\begin{eqnarray}
\frac{T_p}{p!}\int d^{p+1}x\,\epsilon^{a_0a_1\cdots a_p}\left(\frac{1}{2}F^{(p+2)}_{ija_1\cdots a_p,a}\cR^a{}_{a_0}{}^{ij}+F^{(p+2)}_{jaa_1\cdots a_p,i}\cR^{i}{}_{a_0}{}^{aj}\right)\labell{second}
\end{eqnarray}
The Riemann tensor at the linear order in the graviton is
\begin{eqnarray}
\cR_{\mu\nu\rho\lambda}&=&\frac{1}{2}(h_{\mu\lambda,\nu\rho}+h_{\nu\rho,\mu\lambda}-h_{\mu\rho,\nu\lambda}-h_{\nu\lambda,\mu\rho})
\end{eqnarray}
where we have considered perturbation around the flat space where the metric takes the form $G_{\mu\nu}=\eta_{\mu\nu}+h_{\mu\nu}$. The last term in the above amplitude has again on-shell ambiguity. We will show in section 3 that this term in the present form is not consistent with T-duality. However, it can be written in a T-dual invariant form using the on-shell conditions.
The dilaton amplitude can be found from the amplitude \reef{kin3} by using the following polarization:
\begin{eqnarray}
\veps_{\mu\nu}&=&\frac{1}{\sqrt{8}}(\eta_{\mu\nu}-\ell_{\mu}p_{\nu}-\ell_{\nu}p_{\mu})\,;\qquad \ell\!\cdot\! p=1\labell{pol}
\end{eqnarray}
where the auxiliary vector $\ell_{\mu}$ insures that the polarization satisfies the on-shell condition $p\!\cdot\!\veps_{\nu}=0$. One finds the dilaton amplitude to be
\begin{eqnarray}
K(1,2)&\sim &\frac{i(p-3)}{\sqrt{2}}\left(p_1\!\cdot\! N\!\cdot\! p_2 F_{1ia_0\cdots a_p}p_2^i\right)\epsilon^{a_0a_1\cdots a_p}
\end{eqnarray}
The field theory corresponding to the above amplitude is
\begin{eqnarray}
\frac{(p-3)T_p}{(p+1)!}\int d^{p+1}x\,\epsilon^{a_0a_1\cdots a_p}\left(F^{(p+2)}_{ia_0\cdots a_p,j}\phi\,^{,ij}\right)\labell{dil}
\end{eqnarray}
This coupling is zero for D$_{3}$-brane which is consistent with the fact that the world volume theory of D$_{3}$-brane is a conformal field theory.
\subsection{$n=p+4$ case}
For $n=p+4$ case, one needs only the following trace:
\begin{eqnarray}
&&\Tr(\gamma^{\mu_1}\cdots\gamma^{\mu_{p+4}}\gamma^{\mu}\gamma^{a_0}\cdots\gamma^{a_p}\gamma^{\alpha}\gamma^{\nu})\nonumber\\
&&=(-1)^p\eta^{\alpha \mu_1}\eta^{\mu\mu_2 }\eta^{\nu\mu_3}(p+2)(p+3)(p+4)\Tr(\gamma^{\mu_4}\cdots\gamma^{\mu_{p+4}}\gamma^{a_0}\cdots\gamma^{a_p})
\end{eqnarray}
One can easily check that the kinematic factor \reef{kin2} is zero for graviton, and for B-field it is
\begin{eqnarray}
K(1,2)&\sim & -i\frac{q^2}{2\sqrt{2}}F_{1ijka_0\cdots a_p}(\veps_2)^{jk}p_2^i\epsilon^{a_0\cdots a_p}\labell{K123}
\end{eqnarray}
which satisfies the Ward identity. The coupling corresponding to the above amplitude is
\begin{eqnarray}
\frac{T_p}{3!(p+1)!}\int d^{p+1}x\,\epsilon^{a_0\cdots a_p}F^{(p+4)}_{ijka_0\cdots a_p,a}H^{ijk,a}\labell{third}
\end{eqnarray}
Note that $H^{ijk,a}=H^{ija,k}$ when the indices $i,j,k$ are totally antisymmetric, as in above equation. In the next section, we will examine that the consistency of the above couplings with T-duality.
\section{T-duality}
In this section we would like to study the transformation of the couplings that we have found in the previous section under the T-duality. We denote $y$ the Killing direction along which the T-duality is going to be implemented.
The full set of T-duality transformations has been found in \cite{TB,Meessen:1998qm}
\begin{eqnarray}
e^{2\tphi}&=&\frac{e^{2\phi}}{G_{yy}}\nonumber\\
\tG_{yy}&=&\frac{1}{G_{yy}}\nonumber\\
\tG_{\mu y}&=&\frac{B_{\mu y}}{G_{yy}}\nonumber\\
\tG_{\mu\nu}&=&G_{\mu\nu}-\frac{G_{\mu y}G_{\nu y}-B_{\mu y}B_{\nu y}}{G_{yy}}\nonumber\\
\tB_{\mu y}&=&\frac{G_{\mu y}}{G_{yy}}\nonumber\\
\tB_{\mu\nu}&=&B_{\mu\nu}-\frac{B_{\mu y}G_{\nu y}-G_{\mu y}B_{\nu y}}{G_{yy}}\nonumber\\
\tC^{(n)}_{\mu\cdots \nu\alpha y}&=&C^{(n-1)}_{\mu\cdots \nu\alpha }-(n-1)\frac{C^{(n-1)}_{[\mu\cdots\nu|y}G^{}_{|\alpha]y}}{G_{yy}}\nonumber\\
\tC^{(n)}_{\mu\cdots\nu\alpha\beta}&=&C^{(n+1)}_{\mu\cdots\nu\alpha\beta y}+nC^{(n-1)}_{[\mu\cdots\nu\alpha}B^{}_{\beta]y}+n(n-1)\frac{C^{(n-1)}_{[\mu\cdots\nu|y}B^{}_{|\alpha|y}G^{}_{|\beta]y}}{G_{yy}}\labell{Cy}
\end{eqnarray}
where $\mu,\nu,\alpha,\beta\ne y$. In above transformation the metric is the string frame metric. If $y$ is identified on a circle of radius $R$, {\it i.e.,}\ $y\sim y+2\pi R$, then after T-duality the radius becomes $\tilde{R}=\alpha'/R$. The string coupling is also shifted as $\tilde{g}=g\sqrt{\alpha'}/R$.
The strategy for finding T-duality invariant couplings is given in \cite{Garousi:2009dj}. Let us review it here.
Suppose we are implementing T-duality along a world volume direction of a D$_p$-brane denoted $y$. One should first separate the world-volume indices into $y$ index and the world-volume indices which do not include $y$, and then apply the above T-duality transformations. The latter indices are complete world-volume indices of the T-dual D$_{p-1}$-brane. However, the $y$ index in the T-dual theory which is a normal bundle index is not a complete index.
On the other hand, the normal bundle indices of the original theory are not complete in the T-dual D$_{p-1}$-brane. They are not include $y$. In a T-duality invariant theory, the $y$ indices must be combined with the incomplete normal bundle indices to give the complete normal bundle indices in the T-dual D$_{p-1}$-brane. If a theory is not invariant under the T-duality, one should then add new terms to it to have the complete indices in the T-dual theory. In this way one makes the theory to be T-duality invariant by adding new couplings.
One may also implement T-duality along a transverse direction of a D$_p$-brane denoted $y$. In this case one must first separate the transverse indices to $y$ and the transverse indices which do not include $y$, and then apply the above T-duality transformations. The latter indices are the complete transverse indices of the T-dual D$_{p+1}$-brane. However, the complete world-volume indices of the original D$_p$-brane are not the complete indices of the T-dual D$_{p+1}$-brane. They must include the $y$ index to be complete. In a T-duality invariant theory, the $y$ index which is a world-volume index in the T-dual theory must be combined with the incomplete world-volume indices of the T-dual D$_{p+1}$-brane to give the complete world-volume indices.
\subsection{Linear T-duality }
In this subsection we would like to study the consistency of the couplings with linear T-duality transformations. Assuming the NSNS and RR fields are small perturbations around the flat space, the T-duality transformations take the following linear form:
\begin{eqnarray}
&&\tilde{\phi}=\phi-\frac{1}{2}{h}_{yy},\,
\tilde{h}_{yy}=-h_{yy},\, \tilde{h}_{\mu y}=B_{\mu y},\, \tilde{B}_{\mu y}=h_{\mu y},\,\tilde{h}_{\mu\nu}=h_{\mu\nu},\,\tilde{B}_{\mu\nu}=B_{\mu\nu}\nonumber\\
&&\tC^{(n)}_{\mu\cdots \nu\alpha y}=C^{(n-1)}_{\mu\cdots \nu\alpha },\,\tC^{(n)}_{\mu\cdots\nu\alpha\beta}=C^{(n+1)}_{\mu\cdots\nu\alpha\beta y}\labell{linear}
\end{eqnarray}
Consistency of the curvature squared corrections to the DBI action under the above linear T-duality transformations has been examined in \cite{Garousi:2009dj}. The consistency requires adding some $H$-squared terms to the DBI action which are also consistent with the corresponding S-matrix element. We are going to do similar calculation for the couplings that we have found in the previous section.
We begin by studying the T-duality of the couplings in \reef{first}. Consider implementing T-duality along a world volume direction of the D$_p$-brane\footnote{The couplings \reef{first} are consistent with the linear T-duality transformations \reef{linear} when implementing the T-duality along a transverse direction.}. From the contraction with the world volume form, one of the indices $a_2,\cdots, a_{p}$ of the RR field strength or the indices $a_0,a_1$ of the NSNS field strength in \reef{first} must include $y$, and so there are two cases to consider: First when $y$ appears as an index on the RR field strength and second when $y$ is an index in the NSNS field strength. In the former case, we note that there are $p-1$ indices in the RR field strength which are contracted with the volume form. Each of these indices can be $y$. However, because of the totally antisymmetric property of the volume form and the RR field strength, they all are identical. So one can write \reef{first} as
\begin{eqnarray}
\frac{T_p}{2!(p-2)!}\int d^{p+1}x\,\epsilon^{a_0a_1\cdots a_{p-1}y}\left(F^{(p)}_{ia_2\cdots a_{p-1}y,a}H_{a_0a_1}{}^{a,i}- F^{(p)}_{aa_2\cdots a_{p-1}y,i}H_{a_0a_1}{}^{i,a}\right)\labell{first1}
\end{eqnarray}
Note that the indices $i,a$ appear as the derivative indices so nigher of them can be $y$. Moreover, because of the world volume form, none of the indices $a_0,\cdots a_{p-1}$ can be $y$. The transform of the above couplings under the linear T-duality \reef{linear} gives the following couplings for D$_{p-1}$-brane:
\begin{eqnarray}
2\pi\sqrt{\alpha'}\frac{T_p}{2!(p-2)!}\int d^{p}x\,\epsilon^{a_0a_1\cdots a_{p-1}}\left(F^{(p-1)}_{ia_2\cdots a_{p-1},a}H_{a_0a_1}{}^{a,i}- F^{(p-1)}_{aa_2\cdots a_{p-1},i}H_{a_0a_1}{}^{i,a}\right)\nonumber
\end{eqnarray}
where we have also used the fact that $T_p\sim 1/g_s$. Using the relation $2\pi\sqrt{\alpha'}T_p=T_{p-1}$, one observes that the above couplings are exactly the couplings \reef{first} for D$_{p-1}$-brane.
We will now check the case that the T-dual coordinate $y$ is carried by the NSNS field strength.
There are two possibilities for the NSNS field strength in \reef{first} to carry the T-dual coordinate $y$, {\it i.e.,}\ either $a_0$ or $a_1$ carries the index $y$. Since the two possibilities are identical, one can write \reef{first} as
\begin{eqnarray}
\frac{T_p}{(p-1)!}\int d^{p+1}x\,\epsilon^{ya_1\cdots a_{p}}\left(F^{(p)}_{ia_2\cdots a_{p},a}H_{ya_1}{}^{a,i}- F^{(p)}_{aa_2\cdots a_{p},i}H_{ya_1}{}^{i,a}\right)\labell{first2}
\end{eqnarray}
Note again that the indices $i,a$ and $a_1,\cdots , a_p$ can not be $y$. The above couplings transform under linear T-duality to the following couplings of D$_{p-1}$-brane:
\begin{eqnarray}
\frac{T_{p-1}}{(p-1)!}\int d^{p}x\,\epsilon^{a_1\cdots a_{p}}\left(2F^{(p+1)}_{ia_2\cdots a_{p}y,a}\cR^a{}_{a_1}{}^{iy}- 2F^{(p+1)}_{aa_2\cdots a_{p}y,i}\cR^i{}_{a_1}{}^{ay}\right)\labell{first22}
\end{eqnarray}
where we have used the assumption in T-duality that all field are independent of the Killing direction $y$. The coordinate $y$ in the T-dual theory is a transverse coordinate. Inspired by the above couplings, one may guess that the correct form of the couplings for D$_{p-1}$-brane are in fact,
\begin{eqnarray}
\frac{T_{p-1}}{(p-1)!}\int d^{p}x\,\epsilon^{a_1\cdots a_{p}}\left(F^{(p+1)}_{ia_2\cdots a_{p}j,a}\cR^a{}_{a_1}{}^{ij}- 2F^{(p+1)}_{aa_2\cdots a_{p}j,i}\cR^i{}_{a_1}{}^{aj}\right)\labell{first222}
\end{eqnarray}
This is consistent with the couplings \reef{second} that we have found from the S-matrix.
The last term above is not consistent with the T-duality along the world volume direction. To see this, consider the case that $a_1$ carries the index $y$. The world volume index $a$ should be separated into $y$ and $\ta$, which does not include the coordinate $y$. So the second term in \reef{first222} can be written as
\begin{eqnarray}
-\frac{2T_{p-1}}{(p-1)!}\int d^px\,\epsilon^{ya_2\cdots a_p}\left(F^{(p+1)}_{\ta a_2\cdots a_pj,i}\cR^i{}_y{}^{\ta j}+ F^{(p+1)}_{y a_2\cdots a_pj,i}\cR^i{}_y{}^{yj}\right)
\end{eqnarray}
Under the linear T-duality it transforms to
\begin{eqnarray}
-\frac{T_{p-1}}{(p-1)!}\int d^{p-1}x\,\epsilon^{a_2\cdots a_p}\left(F^{(p+2)}_{\ta a_2\cdots a_pjy,i}H^{\ta yj,i}-(-1)^pF^{(p)}_{a_2\cdots a_pj,i}h_{yy}{}^{,ij}\right)
\end{eqnarray}
The first term above, in particular, indicates that there must be the following coupling:
\begin{eqnarray}
-\frac{T_{p}}{2!(p+1)!}\int d^{p+1}x\,\epsilon^{a_0\cdots a_p}F^{(p+4)}_{aa_0\cdots a_pjk,i}H^{akj,i}
\end{eqnarray}
However, this coupling is not produced by the S-matrix element \reef{K123}. To fix this inconsistency, we use the on-shell conditions to rewrite the second term in \reef{first222} in a T-dual form. The index $a$ in this term can be only $a_1$ so we can rewrite it as $-2F^{(p+1)}_{a_1a_2\cdots a_{p}j,i}\cR^i{}_{a}{}^{aj}/p$. Moreover, using the on-shell conditions, one can write the curvature as $\hat{\cR}^{ij}$ where
\begin{eqnarray}
\hat{\cR}_{ij}&\equiv &\frac{1}{2}(\cR_{ia}{}^a{}_j-\cR_{ik}{}^k{}_j)
\end{eqnarray}
It does not have $a_1$ index anymore to produce inconsistency with T-duality. It has been shown in \cite{Garousi:2009dj} that it is invariant under linear T-duality transformations \reef{linear} when $i,j\ne y$. Hence, the couplings \reef{first222} can be written for D$_p$-brane as
\begin{eqnarray}
2T_{p}\int d^{p+1}x\,\epsilon^{a_0\cdots a_{p}}\left(\frac{1}{2!p!}F^{(p+2)}_{ia_1\cdots a_pj,a}\cR^a{}_{a_0}{}^{ij}-\frac{1}{(p+1)!}F^{(p+2)}_{a_0\cdots a_pj,i}\hat{\cR}^{ij}\right)\labell{second1}
\end{eqnarray}
which are equivalent to the couplings \reef{second} using on-shell conditions. These are the couplings in the second line of \reef{LTdual}.
It has been speculated in \cite{Garousi:2009dj} that the non-constant dilaton appears in the string frame D-brane action in the same way that the constant dilaton appears in the action, {\it e.g.,}\ the non-constant dilaton appears only through the overall factor of $e^{-\phi}$ in the string frame DBI action. This proposal has been verified for DBI action by explicit calculation at order $\alpha'^2$ in \cite{Garousi:2009dj}. We now check the proposal for the couplings that we have found. According to this proposal the dilaton appears only through the string frame metric in \reef{second1}.
In other words, the dilaton couplings in the Einstein frame should be given by transforming the string frame couplings \reef{second1} to the Einstein frame, {\it i.e.,}\ replacing $h_{\mu\nu}\rightarrow \phi\eta_{\mu\nu}/2$. This replacement gives $\cR^a{}_{a_0}{}^{ij}\rightarrow 0$, $\cR^i{}_{a}{}^{aj}\rightarrow (\eta^{ij}\phi_{,a}{}^{a}+\eta_{a}{}^a\phi^{,ij})/4$ and $\cR^i{}_{k}{}^{kj}\rightarrow [\eta^{ij}\phi_{,k}{}^{k}+(\eta_{k}{}^k-2)\phi^{,ij}]/4$. Using the on-shell condition that $F_{a_0\cdots a_pi}{}^{,i}=-F_{a_0\cdots a_pa}{}^{,a}=0$, one finds the following dilaton coupling:
\begin{eqnarray}
-\frac{T_p(p-3)}{2(p+1)!}\int d^{p+1}x\,\epsilon^{a_0\cdots a_p}F^{(p+2)}_{a_0\cdots a_pj,i}\phi\,^{,ij}
\end{eqnarray}
which is exactly the coupling \reef{dil}. Note that if one uses the replacement $h_{\mu\nu}\rightarrow \phi\eta_{\mu\nu}/2$ in the couplings \reef{second} which is consistent with S-matrix but not with T-duality, one would not find the correct dilaton coupling in the Einstein frame.
Having fixed the on-shell ambiguity of the last term in \reef{second} by requiring the consistency with linear T-duality, we now fix the on-shell ambiguity of the last term in \reef{first} by requiring that the T-duality along a transverse direction of the equation \reef{second1} should produce the $F^{(p)}H$ couplings. Reversing the steps to find the first term in \reef{second1}, one finds the first term in \reef{first}. To find the T-dual of the last term in \reef{second1}, we write it as
\begin{eqnarray}
-\frac{2T_p}{(p+1)!}\int d^{p+1}x\,\epsilon^{a_0\cdots a_p}\left(F^{(p+2)}_{a_0\cdots a_p\tj,i}\tilde{\cR}^{i\tj}+ F^{(p+2)}_{a_0\cdots a_py,i}\tilde{\cR}^{iy}\right)
\end{eqnarray}
where the transverse index $\tj$ does not include $y$. The T-duality transformation of the first term above gives a term which is reproduced by the second term in \reef{second1}, and the T-duality of the second term gives the following terms:
\begin{eqnarray}
\frac{T_{p+1}}{2(p+1)!}\int d^{p+1}xdy\,\epsilon^{ya_0\cdots a_p}F^{(p+1)}_{a_0\cdots a_p,i}\left(H^{ia}{}_{y,a}-H^{ij}{}_{y,j}\right)
\end{eqnarray}
Inspired by these terms, one guesses that there must be the following couplings:
\begin{eqnarray}
\frac{T_{p}}{2p!}\int d^{p+1}x\,\epsilon^{a_1\cdots a_p}F^{(p+1)}_{a_1\cdots a_p,i}\left(H^{ia}{}_{a_0,a}-H^{ij}{}_{a_0,j}\right)
\end{eqnarray}
This fixes the on-shell ambiguity in the second term in \reef{first}. Hence, the couplings which are consistent with the S-matrix element and with the linear T-duality are those that appear in the first line of \reef{LTdual}.
Now we consider the transformation of the couplings \reef{second1} under linear T-duality transformations along a world volume direction. From the contraction with the world volume form, one of the indices of the RR field strength or the index of the curvature in \reef{second1} must include $y$, and so again there are two cases to consider: First when $y$ appears as an index on the RR field strength and second when $y$ is an index in the curvature. In the former case, one can easily check that the T-dual couplings are consistent with \reef{second1}. In the latter case, we write \reef{second1} as
\begin{eqnarray}
2T_{p}\int d^{p+1}x\,\epsilon^{ya_1\cdots a_{p}}\left(\frac{1}{2!p!}F^{(p+2)}_{ia_1\cdots a_{p}j,a}\cR^a{}_{y}{}^{ij}\right)\labell{second2}
\end{eqnarray}
It transforms under linear T-duality to the following coupling of D$_{p-1}$-brane:
\begin{eqnarray}
T_{p-1}\int d^{p}x\,\epsilon^{a_1\cdots a_{p}}\left(\frac{1}{2!p!}F^{(p+3)}_{ia_1\cdots a_{p}jy,a}H^{iyj,a}\right)\labell{second3}
\end{eqnarray}
Note that after T-duality the transverse indices $i,j$ should be written as $\ti,\tj$ which do not include $y$, however, because $H^{\ti y\ti}$ is totally antisymmetric we wrote it as $H^{iyj}$. Inspired by this couplings, one may guess that the correct form of the coupling for D$_{p}$-brane is in fact,
\begin{eqnarray}
T_{p}\int d^{p+1}x\,\epsilon^{a_0a_1\cdots a_{p}}\left(\frac{1}{3!(p+1)!}F^{(p+4)}_{ia_0\cdots a_{p}jk,a}H^{ikj,a}\right)\labell{second4}
\end{eqnarray}
which is the coupling \reef{third} that we have found from the S-matrix element. This is the coupling in the last line of \reef{LTdual}. There is no index in the NSNS field strength in above coupling that contracts with the world volume form. So continuing the T-duality along a world volume direction, one would find no new term involving $F^{p+6}$. This is consistent with the S-matrix calculation in the previous section that indicates there is no coupling for $F^{p+6}$.
\subsection{Non-linear T-duality }
We have seen in the previous section that the couplings \reef{LTdual} are consistent with the linear T-duality transformations \reef{linear}. However, they are not consistent with nonlinear T-duality transformations. In this paper, we would like to study the effect of the second term in the T-duality transformation of the RR potential in \reef{Cy}. So we consider the following T-duality transformations:
\begin{eqnarray}
&& \tilde{h}_{\mu y}=B_{\mu y},\, \tilde{B}_{\mu y}=h_{\mu y},\,\tilde{h}_{\mu\nu}=h_{\mu\nu},\,\tilde{B}_{\mu\nu}=B_{\mu\nu},\,\tC^{(n)}_{\mu\cdots\nu\alpha\beta}=nC^{(n-1)}_{[\mu\cdots\nu\alpha}B^{}_{\beta]y}\labell{nonlinear}
\end{eqnarray}
The reason for choosing only the nonlinear term above is that the consistency of the RR potential $C$ with this term makes it to be $e^BC$. To see this, consider the linear coupling of the CS action \reef{CS2}, {\it i.e.,}\
\begin{eqnarray}
T_p\int C^{(p+1)}&=&\frac{T_p}{(p+1)!}\int d^{p+1}x\,\epsilon^{a_0\cdots a_p}C^{(p+1)}_{a_0\cdots a_p}
\end{eqnarray}
The transformation of the above coupling of D$_p$-brane under T-duality along a transverse direction gives the following coupling for D$_{p+1}$-brane:
\begin{eqnarray}
\frac{T_{p+1}}{p!}\int d^{p+1}xdy\,\epsilon^{a_0\cdots a_py}C^{(p)}_{a_0\cdots a_{p-1}}B_{a_{p}y}
\end{eqnarray}
This dictates that there must be the following coupling:
\begin{eqnarray}
\frac{T_{p}}{2!(p-1)!}\int d^{p+1}x\,\epsilon^{a_0\cdots a_p}C^{(p-1)}_{a_0\cdots a_{p-2}}B_{a_{p-1}a_p}&=&T_p\int C^{(p-1)}\wedge B
\end{eqnarray}
which is a standard term in the CS action \reef{CS2}.
We now study the consistency of the couplings in the first line of \reef{LTdual} under the T-duality transformation \reef{nonlinear}. To study the effect of last term in \reef{nonlinear}, we have to consider the couplings in which the RR field carries no index $y$. We begin by implementing T-duality along a transverse direction of the D$_p$-brane. The B-fields in the first line of \reef{LTdual} are invariant under \reef{nonlinear}, so
these couplings transform under the T-duality to the following couplings for D$_{p+1}$-brane:
\begin{eqnarray}
\frac{T_{p+1}}{2!(p-1)!}\int d^{p+1}xdy\,\epsilon^{a_0a_1\cdots a_{p}y}\left(\tilde{F}^{(p)}_{ia_2\cdots a_{p},a}{H}_{a_0a_1}{}^{a,i}- \frac{1}{p}\tilde{F}^{(p)}_{a_1\cdots a_p,i}(H_{a_0a}{}^{i,a}-H_{a_0j}{}^{i,j})\right)\labell{first11}
\end{eqnarray}
Let us first consider the second term above. The transformation of the RR field strength $\tilde{F}^{(p)}_{a_1\cdots a_{p}}$ under the T-duality \reef{nonlinear} is
\begin{eqnarray}
\tilde{F}^{(p)}_{a_1\cdots a_{p}}&=&p(\tilde{C}^{(p-1)}_{a_2\cdots a_{p}})_{,a_1}\nonumber\\
&=&pF^{(p-1)}_{a_1\cdots a_{p-1}}B_{a_py}+\frac{p(p-1)}{2}C^{(p-2)}_{a_2\cdots a_{p-1}}H_{a_pya_1}\labell{tF}
\end{eqnarray}
where here and in the subsequent identities we have used the fact that the world volume indices $a_0,a_1,\cdots a_p$ are contracted with the totally antisymmetric world volume tensor. Inserting this in the second term in \reef{first11}, one finds that there must be the following couplings:
\begin{eqnarray}
&&-\frac{T_{p+1}}{2!}\int d^{p+1}xdy\,\epsilon^{a_0a_1\cdots a_{p+1}}\left(\frac{1}{2!(p-1)!}{F}^{(p-1)}_{a_1\cdots a_{p-1}}B_{a_pa_{p+1}}+ \frac{1}{3!(p-2)!}{C}^{(p-2)}_{a_2\cdots a_{p-1}}H_{a_pa_{p+1}a_1}\right)_{,i}\nonumber\\
&&\qquad\qquad\qquad\qquad\qquad\qquad\times (H_{a_0a}{}^{i,a}-H_{a_0j}{}^{i,j})\labell{first110}
\end{eqnarray}
The terms in the bracket in the first line is
\begin{eqnarray}
\frac{1}{(p+1)!}(B\wedge F^{(p-1)})_{a_1\cdots a_{p+1}}+\frac{1}{(p+1)!}(H\wedge C^{(p-2)})_{a_1\cdots a_{p+1}}
\end{eqnarray}
Hence, the couplings \reef{first110} are given by the second term in \reef{LTdual} in which the RR field strength is \reef{newF}.
The transformation of the RR field strength $\tilde{F}^{(p)}_{ia_2\cdots a_{p}}$ in the first term of \reef{first11} is
\begin{eqnarray}
\tilde{F}^{(p)}_{ia_2\cdots a_{p}}&=&\tC^{(p-1)}_{a_2\cdots a_p,i}-(p-1)\tC^{(p-1)}_{ia_3\cdots a_p,a_2}\nonumber\\
&=&(p-1)F^{(p-1)}_{ia_2\cdots a_{p-1}}B_{a_py}+(-1)^{p-2}F^{(p-1)}_{a_2\cdots a_p}B_{iy}\labell{tF2}\\
&&+(-1)^{p-2}\frac{(p-1)(p-2)}{2}C^{(p-2)}_{ia_2\cdots a_{p-2}}H_{a_{p-1}a_py}+(p-1)C^{(p-2)}_{a_2\cdots a_{p-1}}H_{a_pyi}\nonumber
\end{eqnarray}
Inserting this in equation \reef{first11}, one finds that there must be the following new terms:
\begin{eqnarray}
&&\frac{T_{p+1}}{2!}\int d^{p+1}xdy\,\epsilon^{a_0a_1\cdots a_{p}a_{p+1}}\,H_{a_0a_1}{}^{a,i}\left(\frac{1}{2!(p-2)!}F^{(p-1)}_{ia_2\cdots a_{p-1}}B_{a_pa_{p+1}}+\frac{1}{(p-1)!}F^{(p-1)}_{a_3\cdots a_{p+1}}B_{ia_2}\right.\nonumber\\
&&\left.-\frac{1}{3!(p-3)!}C^{(p-2)}_{a_2\cdots a_{p-2}i}H_{a_{p-1}a_pa_{p+1}}+\frac{1}{2!(p-2)!}C^{(p-2)}_{a_2\cdots a_{p-1}}H_{a_pa_{p+1}i}\right)_{,a}\labell{first111}
\end{eqnarray}
Consider the following identities:
\begin{eqnarray}
\frac{1}{p!}(B\wedge F^{(p-1)})_{ia_2\cdots a_{p+1}}&=&\frac{1}{(p-1)!}B_{ia_2}F^{(p-1)}_{a_3\cdots a_{p+1}}-\frac{1}{2!(p-2)!}B_{a_3a_2}F^{(p-2)}_{ia_4\cdots a_{p+1}}\nonumber\\
\frac{1}{p!}(H\wedge C^{(p-2)})_{ia_2\cdots a_{p+1}}&=&\frac{1}{2!(p-1)!}H_{ia_2a_3}C^{(p-2)}_{a_4\cdots a_{p+1}}-\frac{1}{3!(p-3)!}H_{a_2a_3a_4}C^{(p-2)}_{ia_5\cdots a_{p+1}}
\end{eqnarray}
The sum of the above terms gives exactly the terms in the bracket in \reef{first111}. Hence the new terms \reef{first111} are given by the coupling in the first term of \reef{LTdual} in which the RR field strength is given by \reef{newF}.
Now consider T-duality of the couplings in the first line of \reef{LTdual} along a world volume direction. From the contraction with the world volume form, one of the indices $a_0,a_1$ of the NSNS field strength must include $y$. Note that the RR field has no $y$ index in the nonlinear T-duality transformation \reef{nonlinear}. The T-duality on the B-field is the same as in the previous section, so the couplings transform under the T-duality \reef{nonlinear} to the following couplings for D$_{p-1}$-brane:
\begin{eqnarray}
2\frac{T_{p-1}}{(p-1)!}\int d^{p}x\,\epsilon^{a_1\cdots a_{p}}\left(\tilde{F}^{(p)}_{ia_2\cdots a_{p},a}\cR^a{}_{a_1}{}^{iy}- \frac{1}{p}\tilde{F}^{(p)}_{a_1\cdots a_{p},i}\hat{\cR}^{iy}\right)\labell{nfirst22}
\end{eqnarray}
The transformation of the RR field strengths are given by \reef{tF2} and \reef{tF}. However, the $y$ coordinate is now a transverse coordinate, so unlike the previous case the new terms inspired by the above couplings can not be written as $B\wedge F+H\wedge C$. This indicates that there must be some other contributions as well. The other contributions are coming from the T-duality transformation of the couplings in the second line of \reef{LTdual} along a transverse direction. The transformation of these terms under \reef{nonlinear} gives the following couplings for D$_{p+1}$-brane:
\begin{eqnarray}
2\frac{T_{p+1}}{p!}\int d^{p}xdy\,\epsilon^{a_0\cdots a_{p}y}\left(\frac{1}{2!}\tilde{F}^{(p+2)}_{ia_1\cdots a_{p}j,a}\cR^a{}_{a_0}{}^{ij}- \frac{1}{p+1}\tilde{F}^{(p+2)}_{a_0\cdots a_{p}j,i}\hat{\cR}^{ij}\right)\labell{nfirst222}
\end{eqnarray}
The new couplings inspired by the couplings \reef{nfirst22} and \reef{nfirst222} should be given by the couplings in the second line of \reef{LTdual} in which the RR field strength is \reef{newF}.
Similarly, the T-duality of the couplings in the second line of \reef{LTdual} along a world volume direction is given by the following coupling for D$_{p-1}$-brane:
\begin{eqnarray}
T_{p-1}\int d^{p}x\,\epsilon^{a_1\cdots a_{p}}\left(\frac{1}{2!p!}\tilde{F}^{(p+2)}_{ia_1\cdots a_pj,a}H^{iyj,a}\right)\labell{second11}
\end{eqnarray}
and the T-duality of the coupling in the last line of \reef{LTdual} along a transverse direction is given by the following coupling for D$_{p+1}$-brane:
\begin{eqnarray}
T_{p+1}\int d^{p+1}xdy\,\epsilon^{a_0\cdots a_{p}y}\left(\frac{1}{3!(p+1)!}\tilde{F}^{(p+4)}_{i a_0\cdots a_{p}jk,a}H^{ikj,a}\right)\labell{second42}
\end{eqnarray}
The new couplings inspired by the couplings \reef{second11} and \reef{second42} should be given by the couplings in the last line of \reef{LTdual} in which the RR field strength is \reef{newF}. Let us check this case explicitly.
Shifting $p\rightarrow p+1$ in \reef{second11}, one can write this equation as
\begin{eqnarray}
T_{p}\int d^{p+1}x\,\epsilon^{a_0\cdots a_{p}}\left(\frac{1}{2!(p+1)!}\tilde{F}^{(p+3)}_{ia_0\cdots a_pj,a}H^{iyj,a}\right)\labell{second111}
\end{eqnarray}
The transformation of the RR field strength $\tilde{F}^{(p+3)}_{ia_0\cdots a_pj}$ under \reef{nonlinear} is
\begin{eqnarray}
\tilde{F}^{(p+3)}_{ia_0\cdots a_pj}&=&2\tC^{(p+2)}_{a_0\cdots a_pj,i}-(p+1)\tC^{(p+2)}_{ia_1\cdots a_pj,a_0}\nonumber\\
&=&2F^{(p+2)}_{ia_0\cdots a_p}B_{jy}-(p+1)F^{(p+2)}_{ia_0\cdots a_{p-1}j}B_{a_py}\nonumber\\
&&-2(p+1)C^{(p+1)}_{a_0\cdots a_{p-1}j}H_{a_pyi}+C^{(p+1)}_{a_0\cdots a_p}H_{jyi}+\frac{p(p+1)}{2}C^{(p+1)}_{a_0\cdots a_{p-2}ij}H_{a_pya_{p-1}}
\end{eqnarray}
Inserting this in equation \reef{second111}, one finds that there should be the following new terms:
\begin{eqnarray}
&&T_{p}\int d^{p+1}x\,\epsilon^{a_0\cdots a_{p}}\left([\frac{1}{2!(p+1)!}F^{(p+2)}_{ia_0\cdots a_p}B_{jk}-\frac{1}{2!p!}F^{(p+2)}_{ia_0\cdots a_{p-1}j}B_{a_pk}\right.\labell{one}\\
&&\left.-\frac{1}{2!p!}C^{(p+1)}_{a_0\cdots a_{p-1}j}H_{a_pki}+\frac{1}{3!(p+1)!}C^{(p+1)}_{a_0\cdots a_p}H_{jki}+\frac{1}{2!2!(p-1)!}C^{(p+1)}_{a_0\cdots a_{p-2}ij}H_{a_{p-1}a_pk}]_{,a}H^{ikj,a}\right)\nonumber
\end{eqnarray}
Checking the indices, one realizes that the above terms are not given by the last term of \reef{LTdual} in which $F$ is replaced by \reef{newF}. Now shifting $p\rightarrow p-1$ in equation \reef{second42}, one can write it as
\begin{eqnarray}
T_{p}\int d^{p}xdy\,\epsilon^{a_0\cdots a_{p-1}y}\left(\frac{1}{3!(p)!}\tilde{F}^{(p+3)}_{i a_0\cdots a_{p-1}jk,a}H^{ikj,a}\right)\labell{second421}
\end{eqnarray}
The transformation of the RR field strength $\tilde{F}^{(p+3)}_{i a_0\cdots a_{p-1}jk}$ under \reef{nonlinear} is
\begin{eqnarray}
\tilde{F}^{(p+3)}_{i a_0\cdots a_{p-1}jk}&=&3\tC^{(p+2)}_{a_0\cdots a_{p-1}jk,i}-p\tC^{(p+2)}_{ia_1\cdots a_{p-1}jk,a_0}\nonumber\\
&=&3F^{(p+2)}_{ia_0\cdots a_{p-1}j}B_{ky}+pF^{(p+2)}_{ia_0\cdots a_{p-2}jk}B_{a_{p-1}y}\nonumber\\
&&+3pC^{(p+1)}_{a_0\cdots a_{p-2}jk}H_{a_{p-1}yi}+3C^{(p+1)}_{a_0\cdots a_{p-1}j}H_{kyi}+\frac{p(p-1)}{2}C^{(p+1)}_{ia_1\cdots a_{p-3}a_0jk}H_{a_{p-2}a_{p-1}y}\nonumber
\end{eqnarray}
Inserting this in equation \reef{second421}, one finds that the first , third and fourth terms above are reproduced by the new couplings \reef{one} when one chooses $a_p=y$ in \reef{one}. The other two terms led us to guess that there should be the following new terms:
\begin{eqnarray}
&&T_{p}\int d^{p+1}x\,\epsilon^{a_0\cdots a_{p}}\left([\frac{1}{2!3!(p-1)!}F^{(p+2)}_{ia_0\cdots a_{p-2}jk}B_{a_{p-1}a_p}\right.\labell{two}\\
&&\left.\qquad\qquad\qquad\qquad-\frac{1}{3!3!(p-2)!}C^{(p+1)}_{a_0\cdots a_{p-3}ijk}H_{a_{p-2}a_{p-1}a_p}]_{,a}H^{ikj,a}\right)\nonumber
\end{eqnarray}
One can easily check that the new couplings \reef{one} and \reef{two} are reproduced by the last term in \reef{LTdual} in which the RR field strength is given by $B\wedge F^{(p+2)}+H\wedge C^{(p+1)}$.
Now consider the new couplings in \reef{LTdual} which have one RR and two NSNS states. If one implements the T-duality on these terms and use the nonlinear T-duality transformations \reef{nonlinear}, one should find new terms which are given by \reef{LTdual} in which the RR field strength is given by the terms in the second line of \reef{newF}. Let us check this for the second term in \reef{LTdual} which is simple to analyze. Implementing the T-duality along a transverse direction, one finds the following couplings for D$_{p+1}$-brane:
\begin{eqnarray}
&&-\frac{T_{p+1}}{2}\int d^{p+1}xdy\,\epsilon^{a_0\cdots a_py}(H_{a_0a}{}^{i,a}-H_{a_0j}{}^{i,j})\labell{final}\\
&&\times\left(\frac{1}{2!(p-2)!}B_{a_1a_2}\tilde{F}^{(p-2)} _{a_3\cdots a_p}+\frac{1}{3!(p-4)!}H_{a_1a_2a_3}C^{(p-4)}_{a_4\cdots a_{p-1}}B_{a_py}\right)_{,i}\nonumber
\end{eqnarray}
where the T-duality of the RR field strength is given in \reef{tF}. Inspired by the above equation, one guesses that there must be the following couplings:
\begin{eqnarray}
&&-\frac{T_{p+1}}{2}\int d^{p+1}xdy\,\epsilon^{a_0\cdots a_py}(H_{a_0a}{}^{i,a}-H_{a_0j}{}^{i,j})\left(\frac{1}{2!2!2!(p-3)!}B_{a_1a_2}F^{(p-3)}_{a_3\cdots a_{p-1}}B_{a_pa_{p+1}}\right.\nonumber\\
&&\left.+\frac{1}{2!2!3!(p-4)!}B_{a_1a_2}C^{(p-4)}_{a_4\cdots a_{p-1}}H_{a_pa_{p+1}a_3}+\frac{1}{2!2!3!(p-4)!}H_{a_1a_2a_3}C^{(p-4)}_{a_4\cdots a_{p-1}}B_{a_pa_{p+1}}\right)_{,i}\nonumber
\end{eqnarray}
Shifting $p+1\rightarrow p$, one finds that the above couplings are exactly given by the second term in \reef{LTdual} in which the RR field strength is given by the second line of \reef{newF}. This ends our illustration of the consistency between the couplings \reef{LTdual} and the T-duality.
We have seen that the consistency with a particular term of the nonlinear T-duality guides us to write the RR field strength in \reef{LTdual} as ${\cal F}=d{\cal C}$. The resulting couplings however are consistent with all nonlinear terms of the RR potential. In fact the nonlinear T-duality transformation \reef{Cy} for the RR potential can be written as \cite{Taylor:1999pr}
\begin{eqnarray}
\tilde{\cal C}^{(n)}_{\mu\cdots \nu\alpha y}={\cal C}^{(n-1)}_{\mu\cdots \nu\alpha }&,&\tilde{\cal C}^{(n)}_{\mu\cdots\nu\alpha\beta}={\cal C}^{(n+1)}_{\mu\cdots\nu\alpha\beta y}\labell{nonlinear2}
\end{eqnarray}
The calculations in section 3.1 then show that the couplings \reef{LTdual} with the RR field strength ${\cal F}=d{\cal C}$ are consistent with the full nonlinear T-duality transformation for the RR field. It would be interesting to confirm the couplings of one RR and two NSNS fields in \reef{LTdual} by the disk level scattering amplitude.
{\bf Acknowledgments}: I would like to thank Katrin Becker and Rob Myers for useful discussions.
This work is supported by Ferdowsi University of Mashhad under grant p/964(1388/12/4).
\bibliographystyle{/Users/Nick/utphys}
\bibliographystyle{utphys} |
\section{Blaschke products}\label{section.blaschke}
\begin{definition} A (finite) Blaschke product is a function of the form
\begin{equation}\label{blaschkeformula}
b(z) = \lambda \prod_{j=1}^n \frac{z-a_j}{1 - \overline{a_j} z}
\end{equation}
for some nonnegative integer $n$, some $\lambda \in \mathbb{T}$, and some $a_1, \dots, a_n \in \mathbb{D}$. The nonnegative integer $n$ is called the \emph{order} of the Blaschke product.
\end{definition}
If $n=0$ we interpret the empty product as $1$. The domain of a Blaschke product is either $\mathbb{T}$, $\mathbb{D}$, or the closure $\overline{\mathbb{D}}$ of $\mathbb{D}$, depending on context. A Blaschke product is evidently a rational function that maps $\mathbb{T}$ to itself and has no poles in $\mathbb{D}$ (it suffices to check the case $n=1$). It is well known that these properties characterize the Blaschke products.
\begin{proposition}\label{wuk} If a rational function $r$ maps $\mathbb{T}$ to itself and has no poles in $\mathbb{D}$, then it is a Blaschke product of order equal to the number $n$ of zeros of $r$ in $\mathbb{D}$, counted according to multiplicity.
\end{proposition}
\begin{proof}
We induct on $n$. If $n=0$, then $r = q^{-1}$ for some polynomial $q$; write $q(z) = \sum_{k=0}^m q_k z^k$ with $q_m \neq 0$. As $q(\mathbb{T}) \subseteq \mathbb{T}$ we have
$$
q(z)^{-1} = \overline{q(z)} = \overline{q((\overline{z})^{-1})} = \sum_{k=0}^m \overline{q_k} z^{-k} = \frac{\sum_{k=0}^m \overline{q_k} z^{m-k}}{z^m}, \qquad z \in \mathbb{T},
$$
so this holds for all nonzero $z \in \mathbb{D}$. As $q$ has no zeros in $\mathbb{D}$, the extreme right hand side has no pole at $0$; thus $m=0$ and $q$ is constant as desired.
If $r$ has $n+1$ zeros in $\mathbb{D}$, choose one, $a$, and note that $r(z) \cdot (\frac{z-a}{1-\overline{a}z})^{-1}$ has $n$ zeros in $\mathbb{D}$ and maps $\mathbb{T}$ to itself.
\end{proof}
\begin{definition}
If $b$ is a Blaschke product, we let $U_b = \{z \in \mathbb{T}: \Im z \geq 0\}$.
\end{definition}
If the zeros of a Blaschke product are $a_1, \dots, a_n$, we calculate from \eqref{blaschkeformula}
$$
\frac{z b'(z)}{b(z)} = \sum_{j=1}^n \frac{1 - |a_j|^2}{|z - a_j|^2} > 0, \qquad z \in \mathbb{T},
$$
so the argument of $b(e^{it})$ is strictly increasing in $t$. The argument principle implies that $b(e^{it})$ travels $n$ times counterclockwise around $\mathbb{T}$ as $t$ runs from $0$ to $2\pi$.
\begin{corollary}\label{mapping}
A Blaschke product $b$ has order $n$ if and only if $U_b$ is a disjoint union of $n$ arcs.
\end{corollary}
This is the main reason we include $\mathbb{T}$ as a ``union of $0$ arcs.''
\section{Blaschke products from unions of arcs}\label{section.const}
Let $S = \{z \in \mathbb{C}: 0 \leq 2 \Re z \leq 1\}$ and let $\phi$ denote the function
$$
\phi(z) = \frac{\exp(2 \pi i (z-1/4)) - 1}{\exp(2 \pi i (z - 1/4)) + 1}.
$$
It is easy to show (see e.g. \cite[\S III.3]{conwaybook}) that $\phi$ maps $S$ bijectively onto $\overline{\mathbb{D}} \setminus \{\pm 1\}$, that $\phi$ restricts to an analytic bijection of the interior of $S$ with $\mathbb{D}$, that $\phi$ maps the right boundary line of $S$ onto $\{z \in \mathbb{T}: \Im z > 0\}$, and that $\phi$ maps the left boundary line of $S$ onto $\{z \in \mathbb{T}: \Im z < 0\}$.
\begin{proposition}\label{hprop}
If $E$ is a disjoint union of $n \geq 0$ arcs and $h_E$ is given by
\begin{equation}\label{definition.he}
h_E(z) = \frac{1}{2} \hat{E}(0) + \sum_{k=1}^{\infty} \hat{E}(k) z^k, \qquad z \in \mathbb{D},
\end{equation}
then $h_E$ is an analytic map of $\mathbb{D}$ into $S$, and the function $\mathbb{D} \to \overline{\mathbb{D}}$ given by
$$
b_E = \phi \circ h_E
$$
extends uniquely to a Blaschke product $\overline{\mathbb{D}} \to \overline{\mathbb{D}}$ of order $n$ satisfying $U_{b_E} = E$.
\end{proposition}
Using the formulas for $\phi$ and $h_E$ one can show without much work that $b_E$ is a rational function; the work in proving Proposition~\ref{hprop} is to establish that $b_E$ has the mapping properties of Proposition~\ref{wuk}, and hence is a Blaschke product, and to prove that $U_{b_E} = E$.
To motivate the argument, let us work nonrigorously for a moment. Formally we have the series expansion
\begin{equation}\label{chie}
\chi_E(z) = \sum_{k \in \mathbb{Z}} \hat{E}(k) z^k, \qquad z \in \mathbb{T},
\end{equation}
and formal manipulation of the series \eqref{definition.he} with $z \in \mathbb{T}$ then shows that
$$
\chi_E(z) = h_E(z) + \overline{h_E(z)} = 2 \Re h_E(z), \qquad z \in \mathbb{T}.
$$
As $\chi_E$ is $\{0, 1\}$ valued on $\mathbb{T}$, the maximum principle for harmonic functions then implies that $h_E$ maps $\mathbb{D}$ into $S$, so $b_E = \phi \circ h_E$ maps $\overline{\mathbb{D}}$ into $\overline{\mathbb{D}}$ and sends the circle to itself. By Proposition~\ref{wuk} it follows that $b_E$ is a Blaschke product; the equality $U_{b_E} = E$ comes from the mapping properties of $\phi$ on the boundary of $S$.
What makes this argument nonrigorous is that the series \eqref{chie} does not converge for all $z \in \mathbb{T}$, and to equate $\chi_E$ with $2 \Re h_E$ is to ignore the distinction between a discontinuous real valued function on $\mathbb{T}$ and a harmonic function on $\mathbb{D}$. To fill in these gaps, we need to use the actual connection between $2 \Re h_E$ and $\chi_E$--- the former is the Poisson integral of the latter.
\begin{proof}
It is easily checked that \eqref{definition.he} does define an analytic function on $\mathbb{D}$, e.g. because $\sum_{k=1}^{\infty}|\hat{E}(k)|^2$ is convergent. One can then verify the identity
$$
2 h_E(z) = \frac{1}{2\pi} \int_0^{2\pi} \frac{1 + z e^{-is}}{1 - z e^{-is}} \chi_E(e^{is}) \, ds, \qquad z \in \mathbb{D}.
$$
(Fix $z$, expand $\frac{1}{1 - z e^{-is}}$ as a power series in $z$ and interchange the sum and the integral.) Taking real parts it follows that for any $r \in [0,1)$ and any $t$
\begin{equation}\label{realpart}
2 \Re h_E(r e^{it}) = \frac{1}{2\pi} \int_0^{2\pi} P_r(t-s) \chi_E(e^{is}) \, ds,
\end{equation}
where
$$
P_r(t) = \Re \left(\frac{1 + r e^{it}}{1 - r e^{it}}\right)
$$
is the \emph{Poisson kernel}. It is elementary (see e.g. \cite[\S X.2]{conwaybook}) that for $r \in [0,1)$ the function $P_r$ is nonnegative and satisfies $\frac{1}{2\pi} \int_0^{2\pi} P_r(\theta) \, d\theta = 1$; thus \eqref{realpart} implies that $2 \Re h_E(z) \in [0,1]$ for all $z \in \mathbb{D}$, and $h_E$ maps $\mathbb{D}$ into $S$.
As $r$ increases to $1$, the $P_r$ converge uniformly to the zero function on the complement of any neighborhood of $0$ (see e.g. \cite[\S X.2]{conwaybook}). From \eqref{realpart} we conclude
\begin{equation}\label{boundary}
\lim_{r \uparrow 1} 2 \Re h_E(rz) = \chi_E(z)
\end{equation}
at any $z \in \mathbb{T}$ at which $\chi_E$ is continuous. We conclude that for any such $z$ the limit $\lim_{r \uparrow 1} (\phi \circ h_E)(rz)$ exists and is in $\mathbb{T}$.
We claim that $\phi \circ h_E$ is a rational function. In the case $n=0$ this is clear. Otherwise, from the definition of $\phi$ it suffices to show that $\exp(2 \pi i h_E)$ is a rational function, and for this it suffices to treat the case $n=1$. In this case there are real numbers $a < b$ with $b-a<2\pi$ satisfying $E = \{e^{it}: t \in [a,b]\}$, and $\hat{E}(k) = \frac{\exp(-ikb) - \exp(-ika)}{-2 \pi ik}$ for all $k > 0$. Let $\log$ denote the analytic logarithm defined on $\mathbb{C} \setminus \{z \in \mathbb{C}: z \leq 0\}$ that is real on the positive real axis and recall that $\log(1 - z) = -\sum_{k=1}^{\infty} \frac{z^k}{k}$ for all $z \in \mathbb{D}$. A comparison of power series shows
$$
h_E(z) = \frac{b-a}{4\pi} + \frac{1}{2\pi i} \left(\log(1 - e^{-ib} z) - \log(1 - e^{-ia} z)\right), \qquad z \in \mathbb{D},
$$
so $\exp(2 \pi i h_E) = \exp(i\frac{b-a}{2}) \frac{1 - e^{-ib} z}{1 - e^{-ia} z}$ is rational.
At this point we know that $b_E = \phi \circ h_E$ is a rational function mapping $\mathbb{D}$ into itself. From \eqref{boundary} we deduce that $b_E$ maps $\mathbb{T}$ into itself, so $b_E$ is a Blaschke product by Proposition~\ref{wuk}. The equality $U_{b_E} = E$ then follows from \eqref{boundary}. The order of $b_E$ is $n$ by Corollary~\ref{mapping}.
\end{proof}
If $E_1$ and $E_2$ are two unions of arcs related by \eqref{four}, it is clear from the definition that $h_{E_1}$ and $h_{E_2}$ have the same $n$th order Taylor polynomial at $0$. As $\phi$ is analytic at $0$, the same is true of $b_{E_1}$ and $b_{E_2}$.
\begin{corollary}\label{bigcorollary}
If $n \geq 0$ and $E_1$ and $E_2$ are each unions of at most $n$ arcs satisfying
\begin{equation}\label{hyp}
\hat{E_1}(k) = \hat{E_2}(k), \qquad 0 \leq k \leq n,
\end{equation}
then there are Blaschke products $b_1$ and $b_2$, each of order at most $n$, satisfying $E_j = U_{b_j}$ for $j = 1,2$ and
\begin{equation}\label{apply}
\hat{b_1}(k) = \hat{b_2}(k), \qquad 0 \leq k \leq n.
\end{equation}
\end{corollary}
\section{Blaschke products from Toeplitz matrices}\label{section.toep}
Fix a positive integer $n$ for the remainder of this section. Our goal is to show that Blaschke products $b_1$ and $b_2$ having order at most $n$ and satisfying \eqref{apply} must be equal. Let $L^2$ denote the space of square-integrable functions $\mathbb{T} \to \mathbb{C}$, with inner product
$$
\<f,g\> = \frac{1}{2\pi} \int_0^{2\pi} f(e^{it}) \overline{g(e^{it})} \, dt, \qquad f, g \in L^2.
$$
(We identify two functions if they agree almost everywhere.)
For $0 \leq k \leq n$ we let $\zeta^k$ denote the function $\mathbb{T} \to \mathbb{C}$ given by $z \mapsto z^k$. It is immediate that $\{\zeta^k: 0 \leq k \leq n\}$ is an orthonormal subset of $L^2$. We denote its span, the space of analytic polynomials of degree at most $n$, by $P$; we let $\pi: L^2 \to P$ denote the orthogonal projection.
\begin{definition}
If $f: \mathbb{T} \to \mathbb{C}$ is bounded, $T_f: P \to P$ denotes the linear map given by
$$
T_f \xi = \pi( f \xi), \qquad \xi \in P.
$$
Here $f \xi$ is the pointwise product of $f$ and $\xi$.
\end{definition}
If we let $\|T_f\|$ denote the norm of $T_f$ regarded as a linear operator on $P$ and write $\|f\|_{\infty} = \sup_{z \in \mathbb{T}} |f(z)|$, it is clear that
$$
\|T_f\| \leq \|f\|_{\infty}
$$
for any bounded $f$. It is also clear that for any such $f$
$$
\<T_f \zeta^k, \zeta^j\> = \hat{f}(j-k), \qquad 0 \leq j,k \leq n,
$$
so the matrix of $T_f$ with respect to the orthonormal basis $\{\zeta^k: 0 \leq k \leq n\}$ is constant along its diagonals (it is a \emph{Toeplitz matrix}).
If $f$ is a Blaschke product, then $f$ is analytic on $\overline{\mathbb{D}}$, so the matrix of $T_f$ is lower triangular with first column $(\hat{f}(k))_{k=0}^n$. Our hypothesis \eqref{apply} is thus that $T_{b_1} = T_{b_2}$, and to deduce that $b_1 = b_2$ it suffices to show how to recover a Blaschke product $b$ of order at most $n$ from the operator $T_b$ it induces on $P$.
\begin{lemma}\label{biglemma}
If $b$ is a Blaschke product of order at most $n$, then $\|T_b\| = 1$, and for any nonzero $r \in P$ satisfying $\|T_b r\| = \|r\|$ one has $T_b r = b r$.
\end{lemma}
This proof is a special case of the proof of \cite[Proposition 5.1]{sarasoninterp}.
\begin{proof}
There are nonzero polynomials $p$ and $q$, each of degree at most $n$, satisfying $b = p/q$. Clearly $T_b q = p$, and as $b$ maps $\mathbb{T}$ to itself, we have $|p(z)| = |q(z)|$ for all $z \in \mathbb{T}$, so $\|p\| = \|q\|$. We deduce that $\|T_b q\| = \|q\|$ and thus $\|T_b\| \geq 1$; since also $\|T_b\| \leq \|b\|_{\infty} = 1$, we conclude $\|T_b\| = 1$.
If $r \in P$ satisfies $\|T_b r\| = \|r\|$ we have
$$
\|r\|^2 = \|T_b r\|^2 = \|\pi(br)\|^2 \leq \|br\|^2 = \int_0^{2\pi} |b(e^{it})|^2 |r(e^{it})|^2 \, dt = \|r\|^2,
$$
from which $\|\pi(br)\| = \|br\|$ and thus $\pi(br) = br$ as desired.
\end{proof}
\begin{remark}\label{carath}
The argument of Lemma~\ref{biglemma} can be modified to show that if $f$ is bounded and analytic on $\overline{\mathbb{D}}$ and $\|f\|_{\infty} = 1$, then $\|T_f\| \leq 1$ with equality if and only if $f$ is a Blaschke product of order at most $n$. With more work, one can prove the rest of the classical Caratheodory-Fejer theorem: that every lower triangular $(n+1) \times(n+1)$ Toeplitz $M$ satisfying $\|M\| = 1$ is of the form $T_f$ for such an $f$.
\end{remark}
We can now prove Theorem~\ref{main}.
\begin{proof}[Proof of Theorem~\ref{main}]
By Corollary~\ref{bigcorollary} there are Blaschke products $b_1$ and $b_2$ of order at most $n$ satisfying $U_{b_j} = E_j$ for $j=1,2$ and $\hat{b_1}(k) = \hat{b_2}(k)$ for $0 \leq k \leq n$. This second fact implies that $T_{b_1} = T_{b_2}$. By Lemma~\ref{biglemma} there is nonzero $q \in P$ satisfying $\|T_{b_1} q\| = \|T_{b_2} q\| = \|q\|$ and
$$
b_1 = \frac{T_{b_1} q}{q} = \frac{T_{b_2} q}{q} = b_2,
$$
so $E_1 = U_{b_1} = U_{b_2} = E_2$.
\end{proof}
As the Fourier coefficients of a bounded function are coefficients with respect to an orthonormal basis of the Hilbert space $L^2$, one might wonder if Theorem~\ref{main} is a special case of a simpler result about arbitrary orthonormal bases of $L^2$. This is not the case. There are, for example, orthonormal bases $B$ for $L^2$ with the property that for every finite subset $F \subseteq B$, there is an arc $A$ with the property that every element of $F$ is constant on $A$. (The basis $(e^{2 \pi it} \mapsto f(t))_{f \in H}$, where $H$ is the \emph{Haar basis} of $L^2[0,1]$ constructed in \cite[\S III.1]{haar}, has this property.) In this situation, if $E \subseteq A$ and $E' \subseteq A$ are any two unions of arcs with the same total measure, one will have $\<\chi_E, f\> = \<\chi_{E'}, f\>$ for all $f \in F$: \emph{any} finite collection of coefficients with respect to $B$ must fail to distiguish infinitely many unions of $n$ arcs from one another.
\section{An algorithm}\label{algorithms}
Let $\mathcal{F}$ denote the map sending a union of at most $n$ arcs $E$ to the tuple $(\hat{E}(k))_{k=0}^n$ in $\mathbb{C}^{n+1}$. Suppose $c = (c_k)_{k=0}^n$ is given, and we desire to know whether or not $c$ in the range of $\mathcal{F}$. The arguments of the previous sections give us the following procedure. (We use the orthonormal basis of \S\ref{section.toep} to identify linear operators on $P$ with $(n+1)\times(n+1)$ matrices.)
\begin{enumerate}
\item Calculate the $n$th Taylor polynomial at $0$ for $\phi(\frac{c_0}{2} + \sum_{k=1}^n c_k z^k)$, and make its coefficients the first column of a lower-triangular Toeplitz matrix $M$.
\item Evaluate $\|M\|$.
If $\|M\| \neq 1$, then $c$ is not in the range of $\mathcal{F}$.
\item Otherwise $\|M\| = 1$ and by the Caratheodory-Fejer theorem (see Remark~\ref{carath}) there is a unique Blaschke product $f$ of order at most $n$ satisfying $M = T_f$. Find $F = U_f$ (e.g. by solving $f(z) = \pm 1$ to get the endpoints of the arcs) and calculate the coefficients of the $n$th order Taylor polynomial at $0$ for $b_F$.
If these coefficients are the first column of $M$ then $b_F = f$ and $c = \mathcal{F}(F)$; otherwise $c$ is not in the range of $\mathcal{F}$.
\end{enumerate}
\begin{remark} The third step of the algorithm is necessary as the map $E \mapsto b_E$ from unions of $n$ arcs to Blaschke products of order $n$ is not surjective. One can check, for example, that of the Blaschke products $b_t(z) = \frac{z^n - t}{1 - tz^n}$ for real $|t| < 1$, all of which satisfy $U_{b_t} = U_{b_0}$, only $b_0$ is in the range of $E \mapsto b_E$.
\end{remark}
If we know in advance that $c = \mathcal{F}(E)$ is in the range of $\mathcal{F}$, this algorithm can recover $E$ from $c$ in a somewhat explicit fashion. The matrix $M$ constructed from $c$ is $T_{b_E}$; Lemma~\ref{biglemma} implies that if we choose a nonzero $q \in P$ satisfying $\|M q\| = \|q\|$, we will have $b_E = \frac{M q}{q}$. If $q$ is chosen so as to have minimal degree, the polynomials $M q$ and $q$ will have no nontrivial common factors. In this case the degree of $q$ is the order of $b_E$, and the endpoints of the arcs of $E$--- the solutions to $b_E(z) = 1$ and $b_E(z) = -1$--- are the roots of the polynomials $Mq - q$ and $Mq + q$. A computer has no difficulty carrying out this procedure to find the arcs of $E$ to any given precision from the tuple $c = \mathcal{F}(E)$.
As this algorithm involves solving polynomial equations, we cannot expect symbolic formulas for these endpoints of the arcs of $E$ in terms of the Fourier coefficients $\hat{E}(k)$. Formulas for the polynomials $Mq \pm q$, however, can be obtained with some effort. The entries of $M$ are polynomials in $\exp(2 \pi i \hat{E}(0))$, $\hat{E}(1)$, \dots, $\hat{E}(n)$ with complex coefficients. As $M$ has norm $1$, a vector $q$ will satisfy $\|Mq\| = \|q\|$ if and only if $q$ is an eigenvector for the self-adjoint matrix $\adj{M}M$ corresponding to the eigenvalue $1$; we can find such a $q$ by using Gaussian elimination, for example. As the entries of $\adj{M} M$ are polynomials in the entries of $M$ and their complex conjugates, the coefficients of $q$ and $Mq \pm q$ will be rational functions in $\exp(2 \pi i \hat{E}(0))$, $\hat{E}(1)$, \dots, $\hat{E}(n)$ and their complex conjugates. Cases may arise in computing $Mq \pm q$ symbolically: in row reducing the symbolic matrix $\adj{M} M - I$, one needs to know whether or not certain functions of the matrix entries are zero--- but explicit formulas can be obtained in every case.
We give one example. Suppose that $E$ is a union of at most two arcs, with $\hat{E}(0)$, $\hat{E}(1)$, and $\hat{E}(2)$ given. Write $E_0 = \exp(2\pi i\hat{E}(0))$ and $E_k = -2\pi i k \hat{E}(k)$ for $k=1,2$. Carrying out the above procedure, one finds that if both $E_1$ and the denominator of
$$
a = \frac{E_2 \conj{E_1} + 2 E_1 - E_1^2 \conj{E_1} - 2 E_1 E_0}{E_1^2 E_0 + E_2 E_0 - E_2 + E_1^2},
$$
are nonzero, then the starting points of the arcs of $E$ are the solutions $z$ of the equation
$$
z^2 - az + \left(\frac{\conj{E_1} + (1 - E_0) a}{E_1 E_0}\right) = 0.
$$
The endpoints of the arcs of $E$ are given by a similar formula.
|
\section{Introduction}
\emph{Superseparability} may be defined as the polar opposite of
entanglement: two identical charged bosons, with state vectors that
have considerable spatio-temporal overlap, are unable to feel each
other's presence because the state vectors lie in disjoint Hilbert
spaces and cannot be superposed. Mathematically, this possibility
appears to be contained in von Neumann's Hilbert space formulation of
quantum mechanics, but whether or not it is realized in nature can
only be ascertained by experiment. This note describes the principle
of a possible experiment based on the magnetic Aharonov-Bohm{} effect.
The mathematical phenomena that suggest the experiment are subtleties
hidden in the notion of self-adjointness for unbounded operators.
Experience shows that these subtleties can generally be disregarded
in practical applications of quantum mechanics; among physicists,
only the mathematically minded are likely to be familiar with them. For
this reason, a brief review of the basic definitions and results, due
mostly to von Neumann, is provided in Section~\ref{SEC-SELF-ADJ}. It
follows an essential review of the historical background in
Section~\ref{SEC-HIST}. The paper begins with a description of the
experimental scheme in Section~\ref{SEC-SCHEME}; the theory of the
experiment is given, after the historical and mathematical
excursions, in Section~\ref{SEC-REEH}. If the experiment turns out to
be feasible, its results -- be they positive or negative -- will be
consequential, and implications of the possible results are discussed
briefly, in a non-speculative manner, in Section~\ref{SEC-DISCUSS}.
\section{Scheme of the experiment}\label{SEC-SCHEME}
The purpose of the experiment is to determine whether or not two
beams of identical bosons, prepared in inequivalent representations
of the CCR, can interfere with each other. A possible scheme for
such an experiment is shown in figure~1.
A coherent beam of charged bosons (e.g., $\alpha$-particles or
deuterons) from a source $S$ is split into two by a beam-splitter
$P$. One beam goes through the chamber $A$, the other through the
chamber $B$. Neither chamber contains a magnetic field, and the two
are electromagnetically isolated from each other. They contain the
magnetic flux lines $\Phi_{A,B}$ respectively (perpendicular to the
plane of the paper); at least one of these fluxes is continuously
variable over a certain range. The experiment consists of observing
changes in the interference pattern at the detector $D$ as
$\Delta\Phi=\Phi_B-\Phi_A$ is varied.
\begin{figure}[ht]
\vspace{5mm}
\beginpicture
\setcoordinatesystem units <1mm,1mm
\setplotarea x from -75 to 45, y from -20 to 15
\linethickness=2pt
\putrule from -30 -10 to 30 -10
\putrule from -30 10 to 30 10
\putrule from -30 0 to -10 0
\putrule from 10 0 to 30.4 0
\putrule from -10 0.3 to -10 -2.3
\putrule from -10.4 -2.3 to 0.4 -2.3
\putrule from 0 -2.3 to 0 2.3
\putrule from -0.4 2.3 to 10.4 2.3
\putrule from 10 2.3 to 10 -0.3
\putrule from -30 -10.3 to -30 -8
\putrule from -30 -6 to -30 6
\putrule from -30 8 to -30 10.3
\putrule from 30 -10.3 to 30 -2
\putrule from 30 2 to 30 10.3
\put {\circle*{4}} [B1] at -4.5 0
\put {\circle*{4}} [B1] at 5.5 0
\put {\circle*{6}} [B1] at -60 0
\linethickness=1pt
\putrule from -45 -10 to -45 10
\putrule from 43 -5 to 43 5
\linethickness=0.5pt
\setdashes <1mm>
\plot -45 8.5 43 0 -45 -8.5 /
\plot -45 8.5 -60 0 -45 -8.5 /
\footnotesize
\put {$S$} at -61.5 -4
\put {$P$} at -45 -13
\put {$D$} at 43 -8.3
\put {$\Phi_A$} at -6.5 2.5
\put {$\Phi_B$} at 4 -1.5
\put {$B$} at 0 -7
\put {$A$} at 0 7
\endpicture
\vspace{5mm}
\caption{Scheme for a noninterferometer}
\end{figure}
As will be shown in Section~\ref{SEC-REEH}, von Neumann's Hilbert
space formulation of quantum mechanics \cite{vN1932} suggests
that the two beams should interfere \emph{only} when
$\Delta\alpha=q\Delta\Phi/2\pi$ is an integer, where $q$ is the
charge of the boson, which will be $-e$ for deuterons and and $-2e$
for $\alpha$-particles. (We use units in which $\hbar=c=1.)$ For
non-integral values of $\Delta\alpha$, the two beams will belong to
disjoint Hilbert spaces and should not -- if the phrase
\emph{disjoint Hilbert spaces} has physical meaning -- be able to
interfere with each other. For this reason, the scheme of the figure
is called a \emph{noninterferometer}.\footnote{The terms
\emph{superseparability \emph{and} noninterferometer} were introduced
in \cite{SEN2010}.} It is also possible that the above interpretation of
quantum mechanics is invalid, and that both beams belong to the same
Hilbert space. In that case the interference pattern should merely
shift, as $\Delta\alpha$ is varied, returning to the original state
when $\Delta\alpha$ has changed exactly by unity; the fringe shift
should be periodic, with period $1$. If the experiment is realizable,
the case $\Phi_A=\Phi_B\neq 0$ will correspond to the standard
Aharonov-Bohm effect, so that this effect may be used to test the
electromagnetic isolation of the chambers $A$ and $B$. The reader is
referred to the monograph by Peshkin and Tonomura \cite{P-T1989} for a
historical account of the Aharonov-Bohm effect, and to the review
article \cite{T2005} and monograph \cite{T1999} for the
decisive experiments.
It should be recalled that if a flux $\Phi$ is quantized, then
$2e\Phi = 2\pi n$, where $n$ is an integer, so that $\Delta\alpha =
(n_B-n_A)/2$ for deuterons and $\Delta\alpha = n_B-n_A$ for
$\alpha$-particles. Therefore superseparability will never be
observed with $\alpha$-particles if both fluxes are quantized.
We shall end this section with a few reservations and a remark. The
configuration described above is an ideal which may be hard to
realize in the laboratory; that is why we have called it the `scheme'
of an experiment. It is unlikely that the chambers $A$ and $B$ can be
perfectly isolated from each other. It may not be possible to
prepare a state with a sharp value of $\alpha$ if the flux is not
quantized. Finally, the theory of Section~\ref{SEC-REEH} would be
applicable only if the interiors of chambers $A$ and $B$ are, for
purposes of the experiment, reasonable approximations to the punctured
plane. On the other hand, we know -- if only by hindsight -- that the
Aharonov-Bohm{} effect can be observed under conditions that are less than ideal.
Therefore the possibility that superseparability may also be testable
under less than ideal conditions should not be ruled out of hand.
\section{Historical background}\label{SEC-HIST}
We begin by recalling two basic facts. (i) The Born-Jordan
commutation relation $[p,q] = -\mathrm{i}\,\!I$ cannot be represented by
finite-dimensional matrices if $I$ is required to be the identity
matrix. (ii) If it is represented on an infinite-dimensional Hilbert
space $\mathfrak{H}$ with $I$ as the identity operator, then at least one of
$p$ and $q$ must be represented by an unbounded operator (see, for
example, \cite{SEN2010}.\footnote{All our Hilbert spaces will be over
the complex numbers, and will have countable orthonormal bases.}
Unbounded operators are not defined everywhere on a Hilbert space,
and are discontinuous wherever they are defined. They give rise to
mathematical phenomena that are not encountered in the theory of
finite dimensional matrices, and it requires considerable effort to
invest with meaning even the simplest of assertions, such as
$[A,B]=0$, if $A$ and $B$ are unbounded. The basic structures of
quantum mechanics, namely matrix mechanics, wave mechanics and
transformation theory were laid down in 1925--27,\footnote{Dirac's
\emph{Principles of Quantum Mechanics} \cite{D1930} and Heisenberg's
\emph{Physical Principles of Quantum Mechanics} \cite{H1930}
were both published in 1930.} but unbounded operators began to be
explored only in 1929--1930 \cite{vN1929-1930,S1932}. The
`first quantum revolution' (this term is due to Aspect \cite{A2004}) was
completed while unbounded operators were still terra incognita even
to mathematicians. In retrospect, one is struck by the fact that
transformation theory could be developed with scant understanding of
the operators that were to be transformed. By what magic was this
achieved?
The answer lies in an ansatz due to Hermann Weyl and a theo\-rem
proven by von Neumann.
In 1928, Weyl published his book \emph{Gruppentheorie und
Quantenmechanik} \cite{W1928}. In this book he replaced the canonical
commutation relations (CCR) for $N$ degrees of freedom by a
$2N$-parameter Lie group, which had the CCR as its Lie algebra. This
group has become known as the \emph{Weyl group}, and we shall denote
it by $\mathscr{W}_N$. We shall give the argument for $N=1$; the general case
merely requires a cumbersome modification of the notation (see \cite{W1928},
pp.\ 272--276).
Let $a, b \in \mathbb{R}$ and define, formally,
\begin{equation}\label{WEYL-GROUP-1}
u(a) = \exp\,(\mathrm{i} ap),\qquad v(b) = \exp\,(\mathrm{i} bq).
\end{equation}
From the properties of the exponential function, it follows
that
\begin{equation}\label{WEYL-GROUP-2}
u(a)u(a^{\prime}) = u(a+a^{\prime}), \qquad
v(b)v(b^{\prime}) = v(b+b^{\prime}).
\end{equation}
Write $u(-a)= u(a)^{-1}$, $v(-b)=v(b)^{-1}$ and
$u(0)=v(0)=\mathbf{1}$. Formal computation yields the result
\begin{equation}\label{WEYL-GROUP-3}
u(a)v(b)u(a)^{-1}v(b)^{-1} =
\mathrm{e}^{\mathrm{i} ab}\mathbf{1}.
\end{equation}
By definition, the Weyl group $\mathscr{W}_1$ consists of the set of elements
(\ref{WEYL-GROUP-1}), with multiplication defined by
(\ref{WEYL-GROUP-2}) and (\ref{WEYL-GROUP-3}). The element
$\mathbf{1}$ is the identity of the group. The group $\mathscr{W}_1$ is
nonabelian and noncompact, with $\mathbb{R}^2$ as the group manifold, and is
a Lie group. The same is true of the Weyl group $\mathscr{W}_N$ for $N$
degrees of freedom, except that its group manifold is $\mathbb{R}^{2N}$.
Being noncompact, the Weyl groups have no finite dimensional unitary
representations. In a unitary representation, the elements $u(a)$ and
$v(b)$ of $\mathscr{W}_1$ are represented by unitary operators $U(a)$ and
$V(b)$ on the Hilbert space $\mathfrak{H}$, and similar statements hold for
$\mathscr{W}_N$.\footnote{The definition of an infinite-dimensional unitary
representation includes a continuity condition that we have not
specified. The same condition is used in the definition of
one-parameter groups of unitaries.} A result known as Stone's theorem
asserts that a one-parameter group of unitaries $\{U(t)\}$ on a
Hilbert space has an infinitesimal generator $H$: $U(t) = \exp\,(\mathrm{i}
Ht)$, where $H$ is self-adjoint. It is bounded if $\{U(t)\}$ is
compact ($t\in S_1$, the circle) and unbounded if $\{U(t)\}$ is not
compact ($t \in \mathbb{R}$). A representation of $\mathscr{W}_N$ defines,
uniquely, a representation of its Lie algebra -- the CCR -- by
self-adjoint operators. In the representation so defined, at least
one member of any canonical pair $p, q$ is represented by an
unbounded operator.\footnote{Self-adjoint operators on
infinite-dimensional Hilbert spaces will be defined precisely in
{Section}~\ref{SEC-SELF-ADJ}. The exponential $\exp\,(\mathrm{i} At),\,
t\in\mathbb{R}$ of the unbounded self-adjoint operator $A$ needs definition,
but we shall content ourselves with the statement that it turns out
to have the expected properties.}
In 1930 von Neumann proved that, for finite $N$, the Weyl group
$\mathscr{W}_N$ has only one irreducible unitary representation
\cite{vN1930}. He gave the name \emph{Schr\"odinger operators} to the
representatives of the canonical variables $p_j, q_j$,
$j=1,\ldots,N$, and titled his paper `Die Eindeutigkeit
Schr\"odingersche Operatoren'. His result has become known as `von
Neumann's uniqueness theorem'. If the CCR were equivalent to the
Weyl group, it would explain why quantum mechanics could be developed
ahead of the theory of unbounded operators without falling into gross
error.
A Lie group defines a unique Lie algebra, but the converse is not
true. The simplest examples are the covering groups of compact
non-simply-connected Lie groups. Examples of this phenomenon that are
relevant to elementary particle physics were unearthed by Michel
as early as 1962 \cite{M1964}. The canonical commutation relations
are \emph{not} abstractly equivalent to the Weyl group; as we shall
see below, the $p_j,q_k$ will not even generate a Lie group unless
they are represented by self-adjoint operators. However, the
requirement of self-adjointness cannot be met in some simple and
realizable physical situations.
\section{Self-adjointness}\label{SEC-SELF-ADJ}
Let $\mathfrak{H}$ be a Hilbert space and $A$ an operator on it. If there
exists a positive number $K$ such that $||A\psi|| \leq K ||\psi||$
for all $\psi\in\mathfrak{H}$, then $A$ is said to be \emph{bounded}. If no
such $K$ exists, then $A$ is said to be \emph{unbounded}. An
unbounded operator $A$ is not defined everywhere on $\mathfrak{H}$; the
subset $\mathfrak{D}(A)\subsetneq\mathfrak{H}$ on which it is defined is called the
\emph{domain} of $A$. If $\mathfrak{D}(A)$ is not dense in $\mathfrak{H}$ then $A$ is
not (yet) mathematically manageable, and one generally assumes that
$A$ is \emph{densely defined}, i.e., $\mathfrak{D}(A)$ is dense in $\mathfrak{H}$.
In the rest of this section we shall deal only with unbounded
operators, and therefore the adjective `unbounded' will be omitted.
An operator $A$ is called \emph{closed} if the set of ordered pairs
$\{(\psi, A\psi)|\psi\in\mathfrak{D}(A)\}$ is a closed subset of
$\mathfrak{H}\times\mathfrak{H}$. An operator $A_1$ is an \emph{extension} of $A$ if
$\mathfrak{D}(A)\subset\mathfrak{D}(A_1)$ and $A_1\psi=A\psi$ for $\psi\in\mathfrak{D}(A)$;
one writes $A \subset A_1$. An operator is called \emph{closable} if
it has a closed extension. Every closable operator $A$ has a smallest
closed extension, which is denoted by $\bar{A}$.
In matrix theory, the adjoint is defined by $(T\boldsymbol{x}, \boldsymbol{y}) =
(\boldsymbol{x}, T^{\star}\boldsymbol{y})$. In operator theory, one has to take
domains into consideration. Let $\varphi,\xi\in\mathfrak{H}$ such that
$(A\psi,\varphi)=(\psi,\xi)$ for all $\psi\in\mathfrak{D}(A)$, and define
$A^{\star}$ by $A^{\star}\varphi=\xi$. Then $\mathfrak{D}(A^{\star})$ is
precisely the set of these $\varphi$. One can show that if $A$ is
densely defined, then $A^{\star}$ is closed. Furthermore, $A^{\star}$
is densely defined if and only if $A$ is closable, and if it is, then
$(\bar{A})^{\star} = A^{\star}$.
If $\mathfrak{D}(A)\subset\mathfrak{D}(A^*)$ and $A\varphi =
A^*\varphi$ for all $\varphi\in\mathfrak{D}(A)$, then $A$ is called
\emph{symmetric}.\footnote{Von Neumann used the term
\emph{Hermitian}, but current usage seems to limit this term to
operators on finite-dimensional vector spaces.} If $\mathfrak{D}(A) =
\mathfrak{D}(A^*)$ and $A\varphi=A^*\varphi$ for all $\varphi \in
\mathfrak{D}(A)$, then $A$ is called
\emph{self-adjoint}.\footnote{Von Neumann used the term
\emph{Hermitian hypermaximal}.} Self-adjoint operators form a
subclass of symmetric operators.
A symmetric operator may have no self-adjoint extension, it may have
many self-adjoint extensions, or it may have only one. In the last
case, it is called \emph{essentially self-adjoint}. One can show that
if $A$ is essentially self-adjoint, then its closure $\bar{A}$ is
self-adjoint, i.e., $\bar{A}$ is the unique self-adjoint extension of
$A$.
The fundamental differences between symmetric and self-adjoint
operators are:
\begin{enumerate}
\item The spectrum of a self-adjoint operator is a subset of the real
line, whereas the spectrum of a symmetric operator is a subset of the
complex plane; a symmetric operator is self-adjoint if and only if
its spectrum is a subset of the real line.
\item A self-adjoint operator can be exponentiated, i.e., if $A$ is
self-adjoint then $\exp\,(\mathrm{i} tA)$ is defined for all $t \in
\mathbb{R}$; a symmetric operator which is not self-adjoint
\emph{cannot} be exponentiated.
\end{enumerate}
The representation problem for the CCR (one degree of freedom) may
now be formulated as follows: \emph{Find all pairs of essentially
self-adjoint operators $P, Q$, densely defined on a common domain
$\mathfrak{D}$, such that $[P,Q]\varphi = -\mathrm{i}\varphi$ for every
$\varphi\in\mathfrak{D}$}. There are infinitely many such (inequivalent)
pairs; the interested reader is referred to Schmudgen \cite{SCH1983}
for details, and for references to earlier works.
If $A$ and $B$ are self-adjoint, are defined on a common dense domain
$\mathfrak{D}$ and commute on $\mathfrak{D}$, then $\exp\,
(\mathrm{i}aA)$ and $\exp\,(\mathrm{i}bB)$ are defined for all $a,b
\in \mathbb{R}$ and commute. However, if $A$ and $B$ are merely
essentially self-adjoint, are defined on $\mathfrak{D}$ and commute
on $\mathfrak{D}$, then $\exp\,(\mathrm{i}a\bar{A})$ and $\exp\,
(\mathrm{i}b\bar{B})$ \emph{do not necessarily commute}. This fact,
which at first seems highly counterintuitive, was unearthed by Nelson
in 1958; for details and references, see Reed and Simon
\cite{R-S1972,R-S1975}.
We shall conclude this section with an example. The group of
isometries of $\mathbb{R}^2$ consists of translations and rotations. The
group of isometries of the punctured plane $\mathbb{R}^2\setminus\{O\}$ is
the group of rotations about the origin $O$. What happens to the
translation operators on $\mathbb{R}^2$, namely $\exp\,(\mathrm{i} ap_x)$ and
$\exp\,(\mathrm{i} bp_y)$, $a,b \in \mathbb{R}$ (where
$p_x=-\mathrm{i}\partial/{\partial x},\;p_y=-\mathrm{i}\partial/{\partial y}$),
when the origin is excised?
The operators $\partial/{\partial x},\,\partial/{\partial y}$ are
defined on sets of differentiable functions. A function which is
differentiable on $\mathbb{R}^2$ is necessarily differentiable on
$\mathbb{R}^2\setminus\{O\}$, but the latter has a richer supply of
differentiable functions than $\mathbb{R}^2$, e.g., the function
$r^{-1}\exp\,(-r^2/2)$ (which is also square-integrable). Restricting
the domain enlarges the set of differentiable functions on which
$p_x$ and $p_y$ are defined. This enlargement changes the
\emph{spectra} of these operators, which in turn leads to the failure
of self-adjointness and exponentiability.
\section{Theory of the experiment}\label{SEC-REEH}
In 1988, Helmut Reeh showed that that the `Nelson phenomenon' could
be found in the Aharonov-Bohm effect \cite{R1988}. For brevity, let
us call a spinless particle of charge $q$ moving in a plane
perpendicular to a trapped magnetic flux -- the classical
Aharonov-Bohm example -- an \emph{AB-particle}. Owing to cylindrical
symmetry, the motion of an AB-particle is essentially
two-dimensional. Its canonical operators may be written, formally,
as
\begin{equation}\label{AB-OP-1}
\boldsymbol{p} = -\mathrm{i}\frac{\partial}{\partial\boldsymbol{x}} +
q\boldsymbol{A},\;\; \boldsymbol{q} = \text{multiplication by}\;\boldsymbol{x}.
\end{equation}
\noindent
Boldface symbols denote 2-vectors in the $XY$-plane. The vector
potential $\boldsymbol{A}$ (up to a gauge) can be written, in terms of the
magnetic flux $\Phi$, as
\begin{equation}\label{VP-FLUX}
\boldsymbol{A} = \frac{\Phi}{2\pi r}\boldsymbol{e},
\end{equation}
where $r = (x^2+y^2)^{1/2}$ and $\boldsymbol{e}$ is the unit vector at
$(x,y)$ tangent to the circle $r = \text{const}$:
\begin{equation*}
\boldsymbol{e} = \left(-\frac{y}{r}\raisebox{.5ex}{,}\relax{}\; \frac{x}{r}\right)\cdot
\end{equation*}
We shall set $ \alpha = q{\Phi}/{2\pi}$ and use (\ref{VP-FLUX}) to
rewrite the quantities $\boldsymbol{p}$ of (\ref{AB-OP-1}) as
\begin{equation}\label{P-A}
\boldsymbol{p}^{\alpha} = -\mathrm{i}\frac{\partial}{\partial\boldsymbol{x}} +
\alpha\boldsymbol{e},
\end{equation}
where the $\alpha$-dependence of $\boldsymbol{p}$ has been rendered explicit
on the left. The problem is to define the formal quantities
$p_x^{\alpha}$ and $p_y^{\alpha}$ in (\ref{P-A}) as operators on the
Hilbert space $L^2(\mathbb{R}^2 \setminus O) = L^2(\mathbb{R}^2)$;
excision of a single point, here the origin $O$, has no real effect
on an $L^2$-space, but -- as we have seen earlier -- changing the
domains of differentiation operators ever so slightly can have
drastic consequences. Reeh chose, for the common domain of
$p_x^{\alpha},\, p_y^{\alpha}$, the space $\mathscr{D}(\mathbb{R}^2
\setminus O)$ of smooth functions with compact support on
$\mathbb{R}^2\setminus O$. The space $\mathscr{D}(\mathbb{R}^2
\setminus O)$ is dense in $L^2(\mathbb{R}^2)$, and $p_x^{\alpha}$ and
$p_y^{\alpha}$ are distributions on it. If $\varphi \in
\mathscr{D}(\mathbb{R}^2 \setminus O)$, then it follows from
$\text{curl}\,\boldsymbol{A} = 0$ that $[p_x^{\alpha},\,
p_y^{\alpha}]\varphi = 0$.
Consider now the equation
\begin{equation}\label{EIGEN}
{p}^{\alpha}_x\varphi = \left(-\mathrm{i}\frac{\partial}{\partial x} -
\alpha\frac{y}{x^2+y^2}\right)\varphi
= \lambda\varphi.
\end{equation}
It is a linear homogeneous differential equation of the first order
which can be solved explicitly for any $\lambda\in\mathbb{C}$, and the same
holds for the equation ${p}_y^{\alpha}\psi = \lambda\psi$. The
solutions do not have compact support. By exploiting these solutions,
Reeh established the following results \cite{R1988}:
\begin{enumerate}
\item The operators $p_x^{\alpha}$ and $p_y^{\alpha}$
are not self-adjoint; they are essentially self-adjoint.
\item Let $\bar{p}_x^{\alpha}$ and $\bar{p}_y^{\alpha}$ be their
self-adjoint extensions, and define
\begin{equation*}
V_x^{\alpha}(a) = \exp\,(\mathrm{i} a\bar{p}_x^{\alpha}),
\qquad
V_y^{\alpha}(b) = \exp\,(\mathrm{i} b\bar{p}_y^{\alpha}).
\end{equation*}
Then
\begin{equation}\label{REEH-COMM}
V_x^{\alpha}(a)V_y^{\alpha}(b)V_x^{\alpha}(a)^{-1}
V_y^{\alpha}(b)^{-1}
= \mathrm{e}^{\mathrm{i}(\pi\alpha/2)\cdot[\epsilon(x) -
\epsilon(x+a)][\epsilon(y) - \epsilon(y-b)]}\,I,
\end{equation}
where $I$ is the identity operator, and
\begin{equation*}
\epsilon(t) = \left\{\begin{array}{rl}
1&\;\; t>1\\[2mm]
-1&\;\;t<1.
\end{array}
\right.
\end{equation*}
\end{enumerate}
Note that the product $[\ldots][\ldots]$ in the exponent on the
right-hand side of (\ref{REEH-COMM}) can only assume the values
$0,\pm4$, so that the entire right-hand side can only assume the
values $I, \exp\,(\pm 2\pi\mathrm{i}\alpha)I$. It follows that if $\alpha$
is an integer, then the right-hand side of (\ref{REEH-COMM}) equals
the identity operator $I$ \emph{for all admissible} $x,y,a,b$, but
\emph{not} if $\alpha$ is not an integer; in this case the group
generated by the operators $\{x, y, \bar{p}_x^{\alpha},
\bar{p}_y^{\alpha}\}$ is no longer isomorphic with the Weyl group
$W_2$. Clearly, the groups generated by these operators for $\alpha =
\alpha_1, \alpha_2$ are not isomorphic with each other if
$\alpha_1-\alpha_2$ is not an integer, and therefore the
representations of the CCR (for two degrees of freedom) they define
are not unitarily equivalent.
The experiment suggested in {Section}~\ref{SEC-SCHEME} is designed to
determine whether this mathematical inequivalence has observable
physical consequences.
\section{Interpretation of possible results}\label{SEC-DISCUSS}
\begin{enumerate}
\item If superseparability is observed with charged boson beams,
then -- irrespective of the psychological effect of the observation
-- it will confirm that the notion of inequivalent irreducible
representations of the CCR (for a finite number of degrees of
freedom) is physically meaningful; vectors from two inequivalent
representations cannot be superposed upon each other. It should then
prompt the investigation of other possible effects that arise from
the existence of inequivalent irreducible representations of the CCR.
Note that the discussion is at the level of the first quantization.
\item If, however, the phenomenon is \emph{not observed}, we shall
have to conclude that something basic is lacking in our understanding
of the linear space that underlies quantum mechanics: the question
that David Hilbert asked Rolf Nevanlinna in the late 1920's --
\emph{Tell me, Rolf, what is this Hilbert Space that the young people
are talking about?} -- may not yet have been answered to the physicist's
satisfaction.
\item The theoretical considerations of Section \ref{SEC-REEH} do
not apply to fermions. The creation-annihilation operators for a
fermion are bounded, and the canonical anticommutation
relations for a finite number of degrees of freedom have only one
irreducible unitary reprsentation. This was proved by Jordan and
Wigner in their very first paper on anticommutation relations in 1928.
Therefore one should not expect to find the phenomenon of
superseparability among fermions. If the experiment is performed with
electrons and superseparabi\-lity \emph{is} observed, it will pose
new and unsettling problems for theoretical physics.
\end{enumerate}
\section*{Acknowledgements}
I would like to thank T Eisenberg, N Panchapakesan, H Reeh, H Roos, G
L Sewell and A Shimony for their comments on earlier versions of this
paper.
|
\section{Introduction}
\label{sec:intro}
Fractal structures are ubiquitous in nature, coastlines~\cite{mandelbrot1967},
river networks~\cite{river,river2}, snowflakes~\cite{NittmanStanley1987},
growing colonies of
bacteria~\cite{Matsuyama1989,FujikawaMatsushita1989,FujikawaMatsushita1991},
mammalian lungs~\cite{SBHPS1994,BBSS1996,KS1997,AAAMBZSS1998,KT2007},
mammalian bloody vessels~\cite{KT2007}, just to mention few of
them~\footnote{The
interested reader could find many more example in the beautiful
books~\cite{vicsek1992,librofisica}.}. But
also mankind artifacts can
exhibit fractal features, for instance fractal antenna~\cite{HohlfeldCohen1999}
or fluctuations in markets prices~\cite{mandelbrot1963}.
A distinction can be made between {{\it mathematical} or deterministic
fractals~\cite{Mandelbrot1982} for which a complete geometric description can
be provided using simple tools such as homotheties, rotations and copying, and
{\it random} or {\it pseudo} fractals~\cite{vicsek1992} found in nature,
being the
latter characterized by exhibiting fractal properties, for instance
self--similarity, only when
statistical averages are computed, because unavoidable fluctuations and errors
can alter the regular--geometric patterns. Moreover such scale invariance
should be limited to a finite range of scale lengths because of physical
constraints.
It is worth remarking that
some of these physical fractals have functionalities,
e.g. transportation of gases in mammalian lungs, or charges in fractal
antenna, one can thus improve the
geometrical description by including flows and growths constraints. Networks are
therefore the most natural and useful tool to describe such growing complex
structures with flows constraints. We thus hereby propose the {\it Stochastic Weighted Fractal Networks}, SWFN for short, a new class of complex networks whose construction is directly
inspired by such physical fractal structures.
Starting from the pioneering works of Erd\H os and R\'enyi~\cite{ErdosReny1959}, network theory is nowadays a research
field in its own~\cite{AB2002,BLMCH2006} and the scientific activity is mainly
devoted to construct and characterize complex networks exhibiting some of the
remarkable properties of real networks, scale--free~\cite{BA1999},
small--world~\cite{WattsStrogatz1998}, communities~\cite{Fortunato2009},
weighted links~\cite{YJB2001,ZTZH2003,DM2004,BBPV2004,BBV2004,BBV2004b}, just to mention few of them.
In recent years we observed an increasing number of papers were authors proposed
models of deterministic (pseudo) fractal networks~\cite{BRV2001,JKK2002,DGM2002,RB2003,ZZCGFZ2007,Zhang2008,Guan2009,ZZZG2009,ZZZCG2007,carlettirighi} exhibiting scale-free and hierarchical structures. In a limited number of cases, models presented also a stochastic component~\cite{ZZSZG2008,WDS2007,CHLR2003,WDDS2006}.
The aim of the SWFN hereby introduced, is to provide a framework that could be used to (re)analyze flows on natural fractal structures using standard tools of transport theory on networks. Moreover SWFN share with physical fractals several interesting properties, for instance the self-similarity or the self-affinity, the presence of hierarchical structures and a stochastic growth process. Actually this allows us to generalize in a unifying scheme some of the above mentioned models existing in the literature.
The SWFN are constructed via a stochastic process and we are thus
able to analytically characterize their topology as a function of
the parameters involved in the construction, using {\it expectations} obtained
constructing several replicas.
Let us conclude this introduction with two remarks. First of all we named our
models \lq\lq fractal\rq\rq networks instead of \lq\lq pseudo fractal\rq\rq,
because some of the topological properties of SWFN depend on the fractal
dimension of some underlying fractal set, whose value ranges all the positive
real numbers, without any limitation. Second we rather
prefer to talk about \lq\lq stochastic\rq\rq networks to emphasize the
stochastic growth process instead of the randomness of some topological quantities; let us also stress that in the network theory \rq\rq randomness\rq\rq has a precise meaning that we cannot directly apply to this case.
The paper is organized as follows. In the next section we will introduce and
study a deterministic model, that generalize the one proposed
in~\cite{carlettirighi}, and that will serve as the basic building block to
construct the SWFN in
Section~\ref{sect:StochWFN}. Then we conclude with some possible applications
we sum up and draw our conclusions
\section{Deterministic Weighted Fractal Network}
\label{sect:nHWFN}
According to Mandelbrot~\cite{Mandelbrot1982} \lq\lq a fractal is by
definition a set for which the Hausdorff dimension strictly
exceeds the topological dimension\rq\rq. One of the most amazing and
interesting feature of fractals is their {\it self-similarity} or {\it
self-affinity}~\cite{Mandelbrot1985,Mandelbrot1986}, namely looking
at all scales we can find conformal or stretched copies of the whole set;
this is actually the idea used to build up fractals as fixed point of {\it
Iterated Function Systems}~\cite{barnsley1988,edgar1990}, IFS for
short. Such fractals have a Hausdorff dimension completely characterized by
the number of copies and the scaling factors of the IFS. Let us observe that
in this case this dimension coincides with the so called similarity
dimension~\cite{edgar1990}.
Recently, author proposed~\cite{carlettirighi} a new general framework aiming to
construct weighted networks with some a priori prescribed topology depending
on the two main parameters: the number of copies and the scaling
factors, hence on the fractal dimension of the \lq\lq underlying\rq\rq IFS
fractal. The aim of this Section is to generalize such construction to obtain
a larger class of networks; moreover exploiting
the iterative construction we will be able to completely and analytically
describe the network topology in terms of node strength distribution, average
(weighted) shortest path and (weighted) clustering coefficient.
Let us fix a positive integer $s>1$ and $s$ real numbers $f_1,\dots ,f_s\in
(0,1)$ and let us consider a (possibly) weighted network $G$ composed by $N$
nodes, one of which has been labeled {\it attaching node} and denoted by
$a$. We then introduce a map,
$\mathcal{T}_{s,\mathbf{f},a}$, depending on the parameters $s$,
$\mathbf{f}=(f_1,\dots,f_s)$ and on the labeled node $a$, whose action on
networks is described in Fig.~\ref{fig:construction}.
\begin{figure}
\centering
\includegraphics[width=7cm]{construction2b.eps}
\caption{{\it The map $\mathcal{T}_{s,\mathbf{f},a}$}. On the left a generic initial
graph $G$ with its attaching node $a$ (red on-line) and a generic weighted
edge $w\in G$ (blue on-line). On the right the new
graph $G^{\prime}$ obtained as follows: Let $G^{(1)},\dots,G^{(s)}$ be $s$
copies of $G$, whose weighted edges (blue on-line) have
been scaled respectively by a factor $f_1,\dots,f_s$, and let us denote by
$a^{(i)}$, for $i=1,\dots,s$, the node in $G^{(i)}$ image of the labeled
node $a\in G$, then link all those labeled nodes to $a\in G$ through edges of unitary weight. The connected
network obtained in this way will be by definition the image of $G$ through the map:
$G^{\prime}=\mathcal{T}_{s,\mathbf{f},a}(G)$.}
\label{fig:construction}
\end{figure}
So starting with a given initial network $G_0$ we can construct a family of
weighted networks
$(G_k)_{k\geq 0}$ iteratively applying the map $\mathcal{T}_{s,\mathbf{f},a}$:
$G_k=\mathcal{T}_{s,\mathbf{f},a}(G_{k-1})$.
This construction improves the one recently proposed~\cite{carlettirighi} by
avoiding the introduction of an extra node, moreover it offers a unifying
framework where several constructions presented in literature
can be included and generalized, e.g. the model presented in~\cite{JKK2002}
with $m=3$ can
be mapped into to the WFN with $s=3$, $\mathbf{f}=(1,1,1)$, i.e. no weights,
and $G_0=\bullet$. Finally this deterministic construction will be the basic
brick to develop the stochastic network introduced in the following Section~\ref{sect:StochWFN}.
Given $G_0$ and the map $\mathcal{T}_{s,\mathbf{f},a}$ we
are able to completely characterize the topology of each $G_k$ for $k\geq 1$ and also of the
limit network $G_{\infty}$, defined as the fixed point of the map,
$\mathcal{T}_{s,\mathbf{f},a}(G_{\infty})=G_{\infty}$.
\subsection{Results}
\label{sect:result}
The aim of this section is to describe the topology of the graphs $G_k$ for
all $k\geq 1$ and $G_{\infty}$, by analytically
studying their properties such as the average degree, the node
strength distribution, the average (weighted) shortest path and the (weighted)
clustering coefficient.
At each iteration step the graph $G_k$ grows as the number of its nodes
increases according to
\begin{equation}
\label{eq:numnodes}
N_k=(s+1)^kN_0\, ,
\end{equation}
being $N_0$ the
number of nodes in the initial graph, while the number of edges satisfies
\begin{equation}
\label{eq:numedges}
E_k=(s+1)^k(E_0+1)-1\, ,
\end{equation}
being $E_0$ the number of edges in $G_0$. Hence in the limit of
large $k$ the average degree is asymptotically given by
\begin{equation}
\label{eq:asymtavdeg}
\frac{E_k}{N_k}\underset{k\rightarrow
\infty}{\longrightarrow}\frac{E_0+1}{N_0}\, .
\end{equation}
Let us denote the weighted degree of
node $i\in G_k$, also called {\it node strength}~\cite{BBPV2004}, by
$\omega^{(k)}_i=\sum_{j}w_{ij}^{(k)}$, being $w_{ij}^{(k)}$ the weight
of the edge $(ij)\in G_k$; then using the recursive construction,
we can explicitly compute the total node
strength, $W_k=\sum_{i}\omega^{(k)}_{i}$, and easily show
that
\begin{equation}
\label{eq:wk}
W_k=\left[\frac{2s}{F}\left((F+1)^k-1\right)+(F+1)^kW_0\right]\, ,
\end{equation}
being $F=\sum_{j=1}^s f_j$. Let us observe that using the hypothesis $f_j<1$,
it trivially follows that $F<s$, hence we
can conclude that the average node strength goes to zero as $k$
increases: ${W_k}/{N_k}\underset{k\rightarrow \infty}{\longrightarrow} 0$.
\subsection{Node strength distribution.}
\label{ssec:degdist}
Let $g_k(x)$ denote the number of nodes in $G_k$ that have strength
$\omega^{(k)}_i=x$ and let us assume $g_0$ to have values in some finite
discrete subset of the positive reals, namely:
\begin{equation*}
g_0(x)>0 \; \text{if and
only if} \; x\in\{x_1,\dots,x_m\}\, ,
\end{equation*}
otherwise $g_0(x)=0$. Using the property of the map
$\mathcal{T}_{s,\mathbf{f},a}$ we get that after $k$ steps of the construction
the nodes strengths have been rescaled by a factor $f_1^{k_1}\dots f_s^{k_s}$,
where
the non-negative integers $k_i$ do satisfy $k_1+\dots +k_s\leq k$. Because this
can be done in $k!/(k_1!\dots k_s!)$ possible different ways, we get the
following
relation for the node strength distribution for the network $G_k$:
\begin{equation}
\label{eq:nhnodestrength}
g_k(f_1^{k_1}\dots f_s^{k_s}x)=\frac{k!}{k_1!\dots k_s!}g_0(x)\quad
\text{with $k_1+\dots+k_s\leq k$}\, .
\end{equation}
After sufficiently many steps and assuming that the main contribution arises
from the choice $k_1\sim \dots \sim k_s \sim k/s$, we can use Stirling formula
to get the approximate distribution (see Fig.~\ref{fig:construction1bis})
\begin{equation}
\label{eq:nhnodestrapprx}
\log g_k(x) \sim \frac{s\log s}{\log (f_1\dots f_s)} \log x\, ,
\end{equation}
so the nodes strength distribution follows a power law. Let us observe that in
the case of homogeneous scaling, i.e. all $f_j$ equal to some $f\in (0,1)$,
one can prove~\cite{carlettirighi} that Eq.~\eqref{eq:nhnodestrapprx} reduces
to $\log g_k(x)\sim -d_{fract}\log x$ where $d_{fract}=-\log s/\log f$ is the
fractal dimension of the underlying IFS fractal.
\begin{figure}
\centering
\includegraphics[width=9cm]{RenHistogk08092009_001.eps}
\caption{{\it Node Strengths Distribution}. Plot of the renormalized node
strengths
distribution $D^{-1}\log_{10}g_k(x)$, where $D=-s\log s/\log (f_1 \dots
f_s)$. Symbols refer to : $\Box$ the finite approximation $G_{11}$ with
$3145728$
nodes of the WFN with $s = 3$, $\mathbf{f}
=(1/\sqrt{2},1/\sqrt{3},1/\sqrt{5})$ and $G_0=\wfnone$;
$\bigcirc$ the finite approximation $G_9$ with $3359232$ nodes of the WFN
$s=5$, $\mathbf{f} =
(1/\sqrt{5},1/\sqrt{11},1/\sqrt{3},1/\sqrt{7},1/\sqrt{13})$ and
$G_0=\mathrel{\hspace{0.2em}\hbox{$\bullet$}\hspace{-0.2em}\hbox{$-$}\hspace{-0.2em}\hbox{$\bullet$}}$. The
reference line has slope $-1$, linear best
fits (data not shown) provides a slope $-1.037\pm 0.04$ and $R^2=0.798$ for
$\Box$ and $-1.00\pm 0.03$ and $R^2=0.8382$ for $\bigcirc$.}
\label{fig:construction1bis}
\end{figure}
\subsection{Average weighted shortest path.}
\label{ssec:meanpath}
By definition the average {\it weighted shortest path}~\cite{BLMCH2006} of the
graph $G_k$ is
\begin{equation}
\label{eq:wmpath}
\lambda_k=\frac{\Lambda_k}{N_k(N_k-1)}\, ,
\end{equation}
where
\begin{equation}
\label{eq:totwmpath}
\Lambda_k=\sum_{ij\in G_k} p_{ij}^{(k)}\, ,
\end{equation}
being $p_{ij}^{(k)}$ the weighted shortest path linking nodes $i$ and $j$ in
$G_k$. Taking advantage of the
recursive construction and adapting the ideas used in~\cite{carlettirighi}, we get the following recursive relation for $\Lambda_k$
\begin{equation}
\label{eq:Lkrec}
\Lambda_k=(F+1)\Lambda_{k-1}+2s(F+1)N_{k-1}\Lambda_{k-1}^{(a_{k-1})}+2s^2N^2_{k-1}\,
,
\end{equation}
where we introduced $\Lambda_k^{(a_k)}=\sum_{i\in G_k}p_{ia_k}^{(k)}$,
i.e. the sum of all weighted shortest
paths ending at the attaching node, $a_k\in G_k$. We can
prove that for large $k$ the asymptotic behavior of
$\Lambda_k^{(a_k)}$ is given by
\begin{equation}
\label{eq:Lkcasymt}
\Lambda_k^{(a_k)}\underset{k\rightarrow \infty}{\sim}
\frac{sN_0}{s-F}(s+1)^{k}\, ,
\end{equation}
and thus the recursive relation~\eqref{eq:Lkrec} can be explicitely solved
to provide the following asymptotic behavior
in the limit of large $k$ (see Fig.~\ref{fig:renoweigmeanpath})
\begin{equation}
\label{eq:ellkasym}
\lambda_k\underset{k\rightarrow
\infty}{\longrightarrow }
\frac{2s^2 (s+1)}{(s-F)[(1+s)^2-(1+F)]}\, .
\end{equation}
\begin{figure}
\centering
\includegraphics[width=8cm]{RenormalizedWMeanPath.eps}
\caption{{\it The average weighted shortest path}. Plot of the renormalized
average
weighted shortest path $\tilde{\lambda}_k$ versus the iteration number $k$,
where $\tilde{\lambda}_k=\lambda_k\frac{(s-F)[(1+s)^2-(1+F)]}{2s^2(s+1)}$
and $F=f_1+\dots+f_s$. Symbols refer to : $\Box$ the WFN $s=3$,
$\mathbf{f}=(1/\sqrt{2},1/\sqrt{3},1/\sqrt{5})$ and $G_0=\wfnone$;
$\bigcirc$ the WFN $s=$5,
$\mathbf{f}=(1/\sqrt{5},1/\sqrt{11},1/\sqrt{3},1/\sqrt{7},1/\sqrt{13})$ and
$G_0=\mathrel{\hspace{0.2em}\hbox{$\bullet$}\hspace{-0.2em}\hbox{$-$}\hspace{-0.2em}\hbox{$\bullet$}}$; $\Diamond$ the WFN $s=2$,
$\mathbf{f}=(1/\sqrt{3},1/\sqrt{5})$ and $G_0=\wfnthree$; $*$ the WFN
$s=2$, $\mathbf{f}=(1/\sqrt{3},1/\sqrt{5})$ and $G_0=\wfnfour$.}
\label{fig:renoweigmeanpath}
\end{figure}
One can explicitly compute the {\it average shortest path}, $\ell_k$, formally
obtained by setting $f_1=\dots=f_s=1$ in the previous formulas~\eqref{eq:wmpath}
and~\eqref{eq:totwmpath}. Hence slightly
modifying the results
presented above we can prove that asymptotically we
have (see Fig.~\ref{fig:ellek})
\begin{equation}
\label{eq:avmeanpath}
\ell_k \underset{k\rightarrow \infty}{\sim} \frac{2s}{(1+s)\log (s+1)}\log
\frac{N_k}{N_0}\, ,
\end{equation}
where growth law of $N_k$ given by~\eqref{eq:numnodes} has been used. Thus the network grows unbounded with
the logarithm of the network size, while the weighted shortest distances
stay bounded.
\begin{figure}
\centering
\includegraphics[width=8cm]{RenormalizedMeanPath.eps}
\caption{{\it The average shortest path $\ell_k$ as a function of the network
size} (semilog plot). Semilog plot of the renormalized average shortest
path $\tilde{\ell}_k$ versus the network size $N_k$, where
$\tilde{\ell}_k=\ell_k \frac{(s+1)\log (s+1)}{2s}$.
Symbols are the same of Fig.~\ref{fig:renoweigmeanpath}. The reference line
has slope $1$.
Linear best fits (data not shown) provides a slope $0.970\pm 0.017$ and
$R^2=0.9999$ for $\Box$, $0.9654 \pm 0.05$ and $R^2=0.9997$ for
$\bigcirc$, $0.97\pm 0.02$ and $R^2=0.9998$ for $\Diamond$ and $1.01\pm
0.03$ and $R^2=0.9993$ for $*$.}
\label{fig:ellek}
\end{figure}
\subsection{Clustering coefficient.}
\label{ssec:cluscoeff}
The clustering coefficient~\cite{WattsStrogatz1998,BLMCH2006} of the graph
$G_k$ is defined as the average over the whole set of
nodes of the local clustering coefficient $c_i^{(k)}$, namely $<c_k>=C_k/N_k$,
where $C_k=\sum_{i\in G_k}c^{(k)}_i$.
Because of the
construction algorithm new triangles are created in the network \lq\lq
boundary\rq\rq while their number doesn't change in the inner core, hence the local clustering coefficient, at
each step increases just by a factor $s+1$; thus after $k$--interactions we will have
$C_k=(1+s)^{k}C_0$, being $C_0=\sum_{i\in G_0}c^{(0)}_i$ the sum of
local clustering coefficients in the initial graph. We can thus conclude that
the clustering
coefficient of the graph is asymptotically
given by:
\begin{equation}
\label{eq:asymptclustcoeff}
<c_k>\underset{k\rightarrow \infty}{\longrightarrow }
\frac{C_0}{N_0}\, .
\end{equation}
On the other hand, one can introduce the links values to weigh the clustering
coefficient~\cite{SKOKK2007}, generalizing the previous relation, we can
easily prove that {\it weighted clustering coefficient} of the graph is
asymptotically given by:
\begin{equation}
\label{eq:asymptclustcoeff}
<\gamma_k>=\left(\frac{1+F}{1+s}\right)^k\frac{C_0}{N_0} \underset{k\rightarrow \infty}{\sim}
\frac{1}{N_k^{1-d}}\, ,
\end{equation}
where $d=\frac{\log (1+F)}{\log (1+s)}$, that results smaller than one because
of the assumption $f_j<1$.
\section{Stochastic Weighted Fractal Networks}
\label{sect:StochWFN}
The aim of this section is to present a class of complex weighted networks that grow according to a {\it stochastic} process and exhibit self-similar or self-affine structures, hereby named {\it Stochastic Weighted Fractal Networks}, for short {\it SWFN}, whose construction is directly inspired by the stochastic growth phenomena present in nature. The idea is
thus to mimic the growth of fractal structures in nature where \lq\lq
possible errors\rq\rq could modify regular patterns.
So let us hypothesize that the growth process is the result of a
stochastic process that selects the actual realization, i.e. the number of copies, between a number of
different possibilities. Thus at each iteration the number of
copies, $s$, is a stochastic variable distributed according to some probability distribution function $p(s)$. Once the numerical value for $s$ has been set, $s$ real numbers $f_1,\dots,f_{s}$ are drawn according to some probability
distribution function $q(f)$ with values in $(0,1)$. Finally a new network is constructed
by applying $\mathcal{T}_{s,(f_1,\dots ,f_{s}),a}$ to the actual network:
\begin{equation}
\label{eq:stoc1b}
G \underset{p(s)}{\longmapsto} G^{(s)}=\mathcal{T}_{s,(f_1,\dots
,f_{s}),a}G \, .
\end{equation}
\medskip
{\bf Remark.} \label{rem:fassump}{\it In the following we will assume the simplifying working
hypothesis that $f_1=\dots =f_{s}=\alpha/s$, i.e. $q(f)=\delta(f-\alpha/s)$, for some given and fixed $\alpha
\in (0,1)$, but of course the model applies to more
general cases.}
\medskip
One can repeat the construction $k$ times and thus obtain with
probability $p(s_k)\dots p(s_1)$, starting
from a network $G_0$, a new network, denoted by $G^{(s_k, \dots , s_1)}$:
\begin{equation}
\label{eq:kstepstoc1b}
G^{(s_k, \dots , s_1)}=
\mathcal{T}_{s_k,(f^{(k)}_1,\dots
,f^{(k)}_{s_k}),a}\circ \dots \circ \mathcal{T}_{s_1,(f^{(1)}_1,\dots
,f^{(1)}_{s_1}),a} G_0\, .
\end{equation}
The network growth results thus a stochastic process, hence we will
describe the main topological network measures in terms of {\it expectations}
obtained
repeating several times the construction. Of course we
could also consider and compute higher order momenta, but the computations
become
rapidly cumbersome, and thus we will non present these results except for
some simple cases, such as the number of nodes.
\subsection{Results: SWFN}
\label{ssec:resswfn-s}
At each step the number of nodes increases with respect the present ones, and
the exact amount depends on the number of branches drawn. Starting
from a network containing $N_0$ nodes we get a new network with
$N^{(s_1)}=(1+s_1)N_0$ nodes with probability $p(s_1)$. Iterating the
construction, after $k$ steps we can obtain with probability
$p(s_k)\dots p(s_1)$ a network with $N^{(s_k, \dots , s_1)}=(1+s_k)\dots
(1+s_1)N_0$ nodes. Hence the expected value for the number of nodes in a
network build after $k$ iterations, is given by:
\begin{eqnarray}
\label{eq:expnodes}
&\phantom{x}& <N_k>=\sum_{s_k,\dots,s_1}p(s_k)\dots p(s_1)N^{(s_k, \dots , s_1)}\\
&=&\sum_{s_k}p(s_k) (1+s_k)\sum_{s_{k-1},\dots,s_1} p(s_{k-1})\dots p(s_1)
N^{(s_{k-1}, \dots , s_1)}\notag \\
&=& (1+<s>)<N_{k-1}>\notag \, ,
\end{eqnarray}
where we denoted by $<s>=\sum_{s_k}p(s_k)s_k$ the average number of branches. We can thus conclude that the
expected number of nodes increases exponentially
\begin{equation}
\label{eq:nodes}
<N_k>=(1+<s>)^kN_0\, .
\end{equation}
Using similar ideas one can prove that the variance of the number of nodes
increases according to:
\begin{equation}
\label{eq:variancenodes}
\sigma_{N_k}^2=N_0^2\left[ (1+<s>)^2+\sigma^2_s\right]^k-(1+<s>)^{2k}N_0\, ,
\end{equation}
where $\sigma_s^2$ is the variance of the distribution of number of branches.
On the other hand the number of edges can increase, with probability $p(s_k)$, in one
iteration by $E^{(s_{k}, \dots , s_1)}=(1+s_k)E^{(s_{k-1}, \dots , s_1)}+s_k$
and thus
the expected number of edges do satisfy
\begin{equation}
\label{eq:edges}
<E_k>=(1+<s>)^k(E_0+1)-1\, .
\end{equation}
These findings are exact in the case of infinitely many replicas, nevertheless
numerical simulations presented in Fig.~\ref{fig:expNEavdegavstr} and in
Fig.~\ref{fig:expWavdegavstr} show the
good agreement also for finitely many repetitions.
\smallskip
{\bf Remark.} {\it The numerical simulations presented in the following will
be obtained assuming for the branch number a Poisson distribution
translated by one, more precisely to avoid a non zero probability of drawing zero branches, we
drawn with probability $p(k)=\lambda^k e^{-\lambda}/k!$ a non negative
integer $k$, and then we set the number of branches to $s=k+1$, in this way
we will get $<s>=\lambda+1$, $\sigma^2=\lambda$ and $s\geq 1$.
Of course our findings are more general and do not rely on the particular
choice for $p(s)$.}
\begin{figure*}
\centering
\includegraphics[width=8cm]{RenAvNodes.eps}\quad
\includegraphics[width=8cm]{RenAvEdges.eps}
\caption{{\it Expected values for number of Nodes and number of Edges}.
Renormalized quantities : $\Delta^{-1}\log (<N_k>/N_0)$ and
$\Delta^{-1}\log <E_k>$ where
$\Delta=\log(1+<s>)$. Symbols refer to : $\bigcirc$ the SWFN with parameters
$\lambda=4$, $\alpha=0.5$ and $G_0=\mathrel{\hspace{0.2em}\hbox{$\bullet$}\hspace{-0.2em}\hbox{$-$}\hspace{-0.2em}\hbox{$\bullet$}}$; $\Box$ the SWFN with parameters
$\lambda=2$, $\alpha=0.5$ and $G_0=\wfnone$; $\Diamond$ the SWFN with
parameters $\lambda=3$, $\alpha=0.8$ and $G_0=\wfnone$.
Expectations are obtained over $100$ replicas. Left panel, the reference
line has slope $1$, linear best fits (data not shown) give $0.9998\pm
0.03$ $R^2=0.9991$ for $\bigcirc$ and $1.017\pm 0.008$ $R^2=0.9999$ for
$\Box$, $1.008\pm 0.005$ $R^2=1.000$ for $\Diamond$. Right panel, the reference
line has slope $1$, linear best fits (data not shown) give $0.9569\pm 0.06$
$R^2= 0.9988$ for $\bigcirc$, $1.019\pm 0.03$ $R^2=0.9997$ for $\Box$ and
$1.06\pm 0.07$ $R^2=0.9955$ for $\Diamond$.}
\label{fig:expNEavdegavstr}
\end{figure*}
\smallskip
In a similar way we can compute the expected average degree after $k$ steps,
$<(E/N)_k>$, and
the expected average node strength after $k$ steps, $<(W/N)_k>$, where $W$ is
the total node strength for the given network realization, to get
(see Fig.~\ref{fig:expWavdegavstr}):
\begin{equation}
\label{eq:expavdeg}
\left< \left(\frac{E}{N}\right)_k\right>\underset{k \rightarrow
\infty}{\rightarrow} \frac{E_0+1}{N_0}
\quad \text{and} \quad \left< \left(\frac{W}{N}\right)_k\right>\underset{k
\rightarrow
\infty}{\rightarrow} 0\, .
\end{equation}
\begin{figure*}
\centering
\includegraphics[width=8cm]{RenAvDeg.eps}\quad
\includegraphics[width=8cm]{RenAvNodesStr.eps}
\caption{{\it Expected values for the average
degree and the average node strength}. Renormalized quantities :
$<(E/N)_k>/D$ where $D=(E_0+1)/N_0$. Symbols are the same of
Fig.~\ref{fig:expNEavdegavstr}. Expectations made over $100$ replicas.}
\label{fig:expWavdegavstr}
\end{figure*}
As we did in the previous section, we are able to analytically study other
relevant quantities such as the {\it expected} value for the {\it weighted shortest
path} $<\lambda_k>$, defined for each network realization
by~\eqref{eq:wmpath}. More precisely, starting from a network $G_0$ and
applying iteratively the above construction
we end up after $k$ iterations with
probability $p(s_k)\dots p(s_1)$ to a network $G^{(s_k,\dots ,s_1)}$, we can
thus define the weighted shortest path for the given network realization
by $\lambda^{(s_k,\dots
,s_1)}= \frac{\Lambda^{(s_k,\dots ,s_1)}}{\left(N^{(s_k,\dots
,s_1)}\right)^2}$. Then using the recursive construction we get:
\begin{widetext}
\begin{eqnarray}
\label{eq:lambdaGk}
\lambda^{(s_k,\dots ,s_1)}&=& \frac{(F_k+1) \Lambda^{s_{k-1},\dots
,s_1}+2s_k^2 \left(N^{(s_{k-1},\dots
,s_1)}\right)^2}{(1+s_k)^2\left(N^{(s_{k-1},\dots
,s_1)}\right)^2}+\frac{2s_k(F_k+1) \Lambda_{a_k}^{s_{k-1},\dots
,s_1}N^{(s_{k-1},\dots ,s_1)}}{(1+s_k)^2
\left(N^{(s_{k-1},\dots ,s_1)}\right)^2}\notag \notag \\
&=& \frac{F_k+1}{(1+s_k)^2}\lambda^{(s_{k-1},\dots ,s_1)}
+2\left(\frac{s_k}{s_k+1}\right)^2+\frac{2s_k(F_k+1)}{(1+s_k)^2}\hat{\lambda}^{(s_k,\dots
,s_1)}\, ,
\end{eqnarray}
\end{widetext}
where we used the growth rate of the number of nodes and~\eqref{eq:Lkrec} and
we introduced $\hat{\lambda}^{(s_k,\dots ,s_1)}~=~\Lambda_{a_k}^{(s_{k},\dots
,s_1)}/N^{(s_{k},\dots ,s_1)}$.
One can finally prove that the expected value for the
average weighted shortest path satisfies the recurrence equation:
\begin{widetext}
\begin{equation}
\label{eq:recexpwsp}
\left< \lambda_k\right>=\left< \lambda_{k-1}\right>\left<\frac{1+F}{(1+s)^2}\right>
+2\left< \left(\frac{s}{s+1}\right)^2\right> +2\left<
\frac{s(1+F)}{(1+s)^2}\right>\left< \hat{\lambda}_k\right> \, ,
\end{equation}
\end{widetext}
where we defined
\begin{eqnarray}
\label{eq:recexpwsphat}
\left<\frac{1+F}{(1+s)^2}\right>&=&\sum_k p(k) \frac{1+F_k}{(1+s_k)^2} \, ,\notag \\
\left< \left(\frac{s}{s+1}\right)^2\right>&=&\sum_k
p(k)\left(\frac{s_k}{s_k+1}\right)^2 \, \text{ and }\notag \\
\left<
\frac{s(1+F)}{(1+s)^2}\right>&=&\sum_k p(k)\frac{s_k(1+F_k)}{(1+s_k)^2}\, .
\end{eqnarray}
Under the simplifying assumption $f_1=\dots=f_{s_k}=\alpha/s_k$ we get
$F_k=\alpha$ and thus we
can simplify the previous equations and obtain (see Fig.~\ref{fig:avwpath}):
\begin{widetext}
\begin{equation}
\label{eq:expwspanal2}
<\lambda_k> \underset{k \rightarrow \infty}{\rightarrow}
\left<\left(\frac{s}{s+1}\right)^2\right>\frac{2}{1-(\alpha+1)\left<1/(1+s)^2\right>}+\left<\frac{s}{(1+s)^2}\right>\left<\frac{s}{1+s}\right>\frac{2(\alpha+1)}{1-(\alpha+1)\left<1/(1+s)^2\right>}
\frac{1}{1-(\alpha+1)\left<1/(1+s)\right>}\, .
\end{equation}
\end{widetext}
\begin{figure}
\centering
\includegraphics[width=8cm]{RenAvWPath.eps}
\caption{{\it Expected values for the average weighted shortest
path}. Renormalized
quantities: $<\tilde{\lambda}_k>= L^{-1}<\lambda_k>$, where $L$ is the
right hand side of Eq.~\eqref{eq:expwspanal2}.
Symbols are the same of Fig.~\ref{fig:expNEavdegavstr}. Expectations made over
$20$ replicas.}
\label{fig:avwpath}
\end{figure}
One can consider the {\it expected shortest path} by formally set all the scaling
factors equal to $1$ and similar technics allow to conclude that (see
Fig.~\ref{fig:avpath})
\begin{equation}
\label{eq:expshrtpath}
<\ell_k> \underset{k \rightarrow \infty}{\sim}\left<
\left(\frac{s}{s+1}\right)^2\right>\frac{2}{1-\left<1/(s+1)\right>}\frac{1}{\log
(1+<s>)}\log\frac{<N_k>}{N_0}\, .
\end{equation}
{\bf Remark.} {\it
Let us observe that in the case where only one value of $s$ is possible,
i.e. the probability distribution of the number of branches reduces to a
$\delta$--distribution, $p(s)=\delta_{s,s'}$,
then the above result coincide with the ones presented for the WFN in
Section~\ref{sect:nHWFN}.}
\begin{figure}
\centering
\includegraphics[width=8cm]{RenAvPath.eps}
\caption{{\it Expected values for the average shortest path as a function of the
network size} (semilog plot). Semilog plot of the renormalized expected
average shortest path $<\tilde{\ell}_k>=<\ell_k>/M$ versus the network size
$N_k$, where
$M$ is the right hand side of Eq.~\eqref{eq:expshrtpath}.
Symbols are the same of Fig.~\ref{fig:expNEavdegavstr}. Expectations made over $20$ replicas. The reference line
has slope $1$, linear best fits (data not shown) provides $0.95\pm 0.01$ with
$R^2=0.9999$ for $\Box$, $0.95\pm 0.07$ $R^2=0.9968$ for $\bigcirc$ and
$0.98\pm 0.01$ $R^2=0.9999$ for $\Diamond$.}
\label{fig:avpath}
\end{figure}
\section{Conclusions}
\label{sect:conclusion}
In this paper we proposed a unifying general framework for complex weighted networks sharing several properties with fractal sets, the {\it Stochastic Weighted Fractal Networks}. This theory, that generalizes to networks the construction of physical fractals, allows us to build complex networks with a prescribed topology, whose main quantities can be analytically predicted in terms of expectations and have been
shown to depend on the fractal dimension of some underlying fractal; for instance the networks are scale-free, the exponent being the related to the fractal dimension of the underlying IFS. Moreover the SWFN share with fractals, the self-similar or self-affine structure.
These networks exhibit the small-world property. In fact the average shortest path increases logarithmically with the system size; hence it is as small as the average shortest path of a random network with the same number of nodes and same average degree. On the other hand the clustering coefficient is asymptotically constant, thus larger than the clustering coefficient of a random network that shrinks to zero as the system size increases.
As already observed~\cite{carlettirighi} the self-similarity property of the SWFN make them suitable to model real problems involving some kind of diffusion over the network coupled with local looses of flow, here modeled via the parameters
$f<1$. Moreover the stochastic growth process allows us to introduce more realism in the construction.
|
\section{Introduction}
Type systems for lock-based race and deadlock static detection try to
contradict the idea put forward by some authors that ``the association
between locks and data is established mostly by
convention''~\cite{shavit:transactions-are-tomorrow-loads}.
Despite all the pathologies usually associated with locks (in the
aforementioned article and others), and specially at system's level,
locks are here to stay~\cite{cantrill.bonwick:real-world-concurrency}.
Deadlock detection should be addressed at the appropriate level
of abstraction, for, in general, compiled code that does not deadlock
allows us to conclude nothing of the source code.
Nevertheless, the problem remains valid at the assembly level and fits
quite nicely in the philosophy of typed assembly
languages~\cite{m.walker.c.g:sytemf-to-tal}.
By capturing a wider set of semantic properties, including the absence
of deadlocks, we improve compiler certification in systems where code
must be checked for safety before execution, in particular those with
untrusted or malicious components.
Our language targets a shared-memory machine featuring an array of
processors and a thread pool common to all
processors~\cite{cogumbreiro.martins.vasconcelos:compiling-pi-into-mtal,vasconcelos.martins:multithreaded-tal}. The
thread pool holds threads for which no processor is available, a
scheduler chooses a thread from this pool should a processor become
idle. Threads voluntary release processors---our model fits in the
cooperative multi-threading category. For increased flexibility (and
unlike many other models,
including~\cite{flanagan.abadi:types-safe-locking}) we allow forking
threads that hold locks, hence we allow the suspension of processes
while in critical regions.
A prototype implementation can be found at
\href{http://gloss.di.fc.ul.pt/mil}{http://gloss.di.fc.ul.pt/mil}.
\begin{figure}[t]
\begin{lstlisting}
main () {
f1,r3 := newLock; f3,r5 := newLock; f2,r4 := newLock -- 3 forks
r1 := r3; r2 := r4; fork liftLeftFork[f1,f2] -- 1st philosopher
r1 := r4; r2 := r5; fork liftLeftFork[f2,f3] -- 2nd philosopher
r1 := r5; r2 := r3; fork liftLeftFork[f3,f1] -- 3rd philosopher
done
}
liftLeftFork forall[l,m].(r1:<l>^l, r2:<m>^m) {
r3 := testSetLock r1
if r3 = 0b jump liftRightFork[l,m]
jump liftLeftFork[l,m]
}
liftRightFork forall[l,m].(r1:<l>^l, r2:<m>^m) requires {l} {
r3 := testSetLock r2
if r3 = 0b jump eat[l,m]
jump liftRightFork[l,m]
}
eat forall[l,m].(r1:<l>^l, r2:<m>^m) requires {l,m} {
-- eat
unlock r1 -- lay down the left fork
unlock r2 -- lay down the right fork
-- think
jump liftLeftFork[l,m]
}
\end{lstlisting}
\caption{The dining philosophers written in MIL}
\label{fig:philosophers}
\end{figure}
The code in Figure~\ref{fig:philosophers} presents a typical example
of a potential deadlock comprising a cycle of threads where each
thread requests a lock hold by the next thread. Imagine the code
running on a two-processors machine: after \lstinline|main| completes
its execution, each philosopher embarks on a busy-waiting loop, only
that two of them will be running in processors, while the third is
(and will indefinitely remain) in the run-pool.
Situations of deadlocks comprising suspended code are known to be
difficult to deal
with~\cite{kontothanassis.etal:scheduler-conscious-sync}.
Our notion of deadlocked state takes into account running and
suspended threads.
Another source of difficulties in characterizing deadlock states
derives from the low-level nature of our language that decouples the
action of lock acquisition from that of entering a critical section,
and that features non-blocking instructions only.
As such the meaning of ``entering a critical section'' cannot be of a
syntactic nature.
A characteristic of our machine is the syntactic dissociation of the
test-and-set-lock and the jump-to-critical operations, for which we
provide two distinct instructions, as found in conventional
instruction sets. Furthermore, there is no syntactic distinction
between a conventional conditional jump and a (conditional)
jump-to-critical instruction, and the test-set-lock and
jump-to-critical instructions can be separated by arbitrary assembly
code. As far as the type system goes, the thread holds the lock only
after the conditional jump, even though at runtime it may have been
obtained long before.
The main contribuitons of this paper are:
\begin{itemize}
\item A type system for deadlock elimination. We devise a type system
that establishes a strict partial order on lock acquisition, hence
enforcing that well typed MIL programs do not
deadlock---Theorem~\ref{thm:ts};
\item An algorithm for automatic program annotation. In order to check
the absence of deadlock, MIL programs must be annotated to reflect
the order by which locks must be acquired.
Annotating large assembly programs, either manually or as the result
of a compilation process, is not plausible.
We present an algorithm that takes a plain MIL program and produces
an annotated program together with a collection of constraints over
lock sets that are passed to a constraint solver. In case the
constraints are solvable the annotated program is
typeble---Theorem~\ref{thm:soundness}---hence free from deadlocks.
\end{itemize}
The outline of this paper is as follows. The next section introduces
the syntax of programs and machine states, together with the running
example. Then Section~\ref{sec:semantics} presents the operational
semantics and the notion of deadlocked
states. Section~\ref{sec:typing} describes the type system and the
first main result, typable states do not
deadlock. Section~\ref{sec:type-inference} introduces the annotation
algorithm and the second main result, the correctness of the algorithm
with respect to the type system. Finally, Section~\ref{sec:conclusion}
describes related work and concludes the paper.
\begin{myfigure}
\begin{alignat*}2
& \textit{registers} &
r \Space{::=}& \register 1 \;\mid\; \dots \;\mid\; \register\milkw R
\\
& \textit{lock values} &
b \Space{::=}& \lockbit {0} \;\mid\; \lockbit 1 \;\mid\; \lockbit {0}^\lambda
\\
& \textit{values} &
v \Space{::=}& r \;\mid\; n
\;\mid\; b
\;\mid\; l
\;\mid\; \instantiation\lambda
\;\mid\; \uninit
\\
& \textit{instructions} &
\iota \Space{::=}&
\\
& \quad\textit{control flow} &&
r \assign v \;\mid\;
\aopi \;\mid\;
\cjumpi \;\mid\;
\fork\ v
\\
& \quad\textit{memory} &&
\malloci \;\mid\;
\loadi \;\mid\;
\storei \;\mid\;
\\
& \quad\textit{locking} &&
\lambda\colon(\Lambda, \Lambda),r \assign \newlock \;\mid\;
\tsli \;\mid\;
\unlock\ v
\\
& \textit{inst.\ sequences}\quad &
I \Space{::=}&
\iota ; I \;\mid\;
\jumpi \;\mid\;
\done\\
& \textit{types} &
\tau \Space{::=}&
\milkw{int} \;\mid\;
\lambda \;\mid\;
\tupletype{\pol\tau}{\lambda} \;\mid\;
\codetype \Gamma \Lambda \;\mid\;
\universal\lambda\Lambda\Lambda\tau
\\
& \textit{register file types}\quad &
\Gamma \Space{::=}& \register 1\colon \tau_1,\dots,\register n\colon\tau_n
\\
& \textit{permissions} &
\Lambda \Space{::=}& \lambda_1,\dots,\lambda_n
\\
& \textit{heaps} &
H \Space{::=}& \{l_1 \colon h_1, \dots, l_n \colon h_n\}
\\
& \textit{heap values}\quad &
h \Space{::=}& \heaptuple{v_1\dots v_n}{\lambda} \;\mid\; \heapcode \tau I
\\
& \textit{thread pool} &
T \Space{::=}& \{\poolitem{l_1[\vec\lambda_1]}{R_1},\dots,\poolitem{l_n[\vec\lambda_n]}{R_n}\}
\\
& \textit{register files} &
R \Space{::=}& \{\register 1 \colon v_1, \dots, \register\milkw R \colon v_\milkw R\}
\\
& \textit{processors array}\qquad &
P \Space{::=}& \{1 \colon p_1, \dots, \milkw N \colon p_{\milkw N}\}
\\
& \textit{processor}\qquad &
p \Space{::=}& \processor R \Lambda I
\\
& \textit{states} &
S \Space{::=}& \program H T P \;\mid\; \milkw{halt}
\end{alignat*}
\caption{Syntax.}
\label{fig:syntax-instructions}
\end{myfigure}
\section{Syntax}
\label{sec:syntax}
The syntax of our language is generated by the grammar in
Figure~\ref{fig:syntax-instructions}. %
We rely on two mutually disjoint sets for \emph{heap labels}, ranged
over by~$l$, and for \emph{singleton lock types}, ranged over by
$\lambda$. Letter $n$ ranges over integer values.
Values $v$ comprise registers $r$, integer values $n$, lock values
$b$, labels $l$, type application $v[\lambda]$, and uninitialised
values~$\uninit$.
Lock value $\lockbit 0$ represents an open lock, whereas lock value
$\lockbit 1$ denotes a closed lock; the~$\lambda$ annotation in
$\lockbit 0^\lambda$ allows to determine the lock guarding the
critical section a processor is trying to enter and will be useful
when defining deadlocked states. Lock values are runtime entities,
they need to be distinct from conventional integer values for typing
purposes only.
Labels are used as heap addresses.
Uninitialised values represent meaningless data of a certain
type.
Most of the machine instructions $\iota$ presented in
Figure~\ref{fig:syntax-instructions} are standard in assembly
languages.
Distinct in MIL are the instructions for creating new
threads---$\milkw{fork}$ places in the run queue a new thread waiting for
execution---, for allocating memory---$\milkw{malloc} \mallocarg {\tau_1,
\dots, \tau_n} \lambda$ allocates a tuple in the heap protected by
lock $\lambda $ and comprising $n$ cells each of which containing an
uninitialised value of type $\tau_i$---, and for manipulating locks. In
this last group one finds $\milkw{newLock}$ to create a lock in the heap and
store its address in register $r$ ($\lambda$ describes the singleton
lock type associated to the new lock, further described below), $\milkw{testSetLock}$
to acquire a lock, and $\milkw{unlock}$ to release a lock.
Instructions are organised in sequences~$I$, ending in $\milkw{jump}$ or in
$\milkw{done}$.
Instruction \milkw{done} terminates a thread, voluntarily releasing the core,
giving rise to a cooperative multi-threading model of computation.
Types $\tau$ include the integer type $\milkw{int}$, the singleton lock
type~$\lambda$, the tuple type $\tupletype{\pol\tau}{\lambda}$ describing a tuple in the
heap protected by lock~$\lambda$, and the code type $\codetypea$
representing a code block abstracted on singleton lock types
$\pol\lambda$, expecting registers of the types in $\Gamma$ and
requiring locks as in $\Lambda$. Each universal variable is bound by
two sets of singleton lock types $\Lambda$, used for deadlock
prevention, as described below.
For simplicity we allow
polymorphism over singleton lock types only; for abstraction over
arbitrary types
see~\cite{cogumbreiro.martins.vasconcelos:compiling-pi-into-mtal}.
The \emph{abstract machine} is parametric on the number of
available processors $\milkw N$, and on the number of registers per
processor~$\milkw R$.
An abstract machine can be in two possible states $S$: halted or
running. A running machine comprises a heap $H$, a thread pool $T$,
and an array of processors $P$ of fixed length~$\milkw N$.
Heaps are maps from labels $l$ into \emph{heap values} $h$ that may be
either data tuples or code blocks.
\emph{Tuples} $\heaptuple{v_1, \dots, v_n}\lambda$ are vectors of
mutable values $v_i$ protected by some lock $\lambda$.
Code blocks $\heapcode \codetypea I$ comprise a signature (a code type) and
an instruction sequence~$I$, to be executed by a processor.
A thread pool~$T$ is a multiset of pairs
$\poolitem{l[\vec\lambda]}{R}$, each of which contains the address (a
label) of a code block in the heap, a sequence of singleton lock types
to act as arguments to the forall type of the code block, and a
register file.
A processor array $P$ contains~$\milkw N$ processors, each of which is
composed of a register file $R$ mapping the processor's registers to
values, a set of locks $\Lambda$ (the locks held by the thread running
at the processor, often call the thread's \emph{permission}), and a
sequence of instructions $I$ (the instructions that remain to
execute).
\paragraph{Lock order annotations}
Deadlocks are usually prevented by imposing a strict partial order on
locks, and by respecting this order when acquiring
locks~\cite{birrell:programming-with-threads,coffman.etal:system-deadlocks,flanagan.abadi:types-safe-locking}.
The syntax in Figure~\ref{fig:syntax-instructions} introduces
annotations that specify the locking order.
When creating a new lock, we declare the order between the newly
introduced singleton lock type and the locks known to the program. We
use the notation $\lambda\colon(\Lambda_1,\Lambda_2)$ to mean that
lock type $\lambda$ is greater than all lock types in set $\Lambda_1$
and smaller than each lock type in set $\Lambda_2$.
The annotated syntax differs from the original
syntax (\cite{cogumbreiro.martins.vasconcelos:compiling-pi-into-mtal,vasconcelos.martins:multithreaded-tal})
in two places:
at lock creation $\lambda\colon(\Lambda,\Lambda),r := \milkw{newLock}$;
and in universal types $\universal\lambda\Lambda\Lambda\tau$, where we explicitly specify the
lock order on newly introduced singleton lock types.
\paragraph{Example}
Figure~\ref{fig:philosophers} shows an example of a
\emph{non-annotated} program. Annotating such a program requires
describing the order for each lock introduced in code block
\lstinline|main|, say,
\begin{lstlisting}
f1::({},{}), r3 := newLock
f3::({f1},{}), r5 := newLock
f2::({f1},{f3}), r4 := newLock
\end{lstlisting}
and at the types for the three code blocks below.
\begin{lstlisting}
liftLeftFork forall[l::({},{})].forall [m::({l},{})].(r1:<l>^l, r2:<m>^m)
liftRightFork forall[l::({},{})].forall[m::({l},{})].(r1:<l>^l, r2:<m>^m) requires {l}
eat forall[l::({},{})].forall[m::({l},{})].(r1:<l>^l, r2:<m>^m) requires {l,m}
\end{lstlisting}
Notice that abstracting one lock at a time, as in the types just shown,
precludes declaring code blocks with non-strict partial orders on
locks, such as $\forall[l\colon(\emptyset,\{m\}),
m\colon(\{l\},\emptyset)].\tau$, which cannot be fulfilled by any
conceivable sequence of instructions.
\begin{myfigure}
\begin{gather*}
\tag\mkRrule{halt}
\frac{
\forall i. P(i) = \processor{\_\ }{\_\ }{\milkw{done}}
}{
\program \_ \emptyset P \rightarrow \milkw{halt}
}
\\[0.5em]
\tag\mkRrule{schedule}
\frac{
P(i) = \processor{\_\ }{\_\ }{\milkw{done}} \qquad
H(l) = \heapcode {\codetypeforall {\pol \lambda\colon(\_,\_)} {\_} \Lambda} I
}{
\program{H}{T\uplus\{\poolitem{l[\pol\lambda']}{R}\}}{P}
\rightarrow
\program H T {P\update{i}{\processor{R}{\Lambda}{I}\subs{\pol\lambda'}{\pol\lambda}}}
}
\\[0.5em]
\tag\mkRrule{fork}
\frac{
P(i) = \processor R {\Lambda\uplus\Lambda'} {(\fork\ v;I)}
\qquad
\operatorname{\hat R}(v) = l[\pol \lambda]
\qquad
H(l) = \heapcode {\codetypeforall {\_} {\_} {\Lambda'}} {\_}
}{
\program{H}{T}{P} \rightarrow \program{H}{T\cup\{\poolitem{l[\pol \lambda]}{R}\}}
{P\update i {\processor R {\Lambda} {I}}}
}
\\[0.5em]
\tag\mkRrule{newLock}
\frac{
P(i) = \processor R \Lambda {(\lambda\colon(\_, \_),r := \milkw{newLock};I)} \quad
l \not\in\operatorname{dom}(H) \quad
\lambda'\text{ fresh}
}{
\program H T P \rightarrow
\program {H\update{l}{\heaptuple{\lockbit 0}{\lambda'}}}
{T}
{P\update i {\processor{R\update rl}{\Lambda}{I\subs{\lambda'}\lambda}}}
}
\\[0.5em]
\tag\mkRrule{tsl}0
\frac{
P(i) = \processor R \Lambda {(\tsli;I)} \quad
\operatorname{\hat R}(v) = l \quad
H(l) = \heaptuple{\lockbit 0}{\lambda}
}{
\program H T P \rightarrow
\program {H\update{l}{\heaptuple{\lockbit 1}\lambda}}
{T}
{P\update i {\processor{R\update r{\lockbit 0 ^\lambda}}{\Lambda \uplus \{\lambda\}}{I}}}
}
\\[0.5em]
\tag\mkRrule{tsl}1
\frac{
P(i) = \processor R \Lambda {(\tsli;I)} \quad
H(\operatorname{\hat R}(v)) = \heaptuple{\lockbit 1}{\lambda} \quad
\lambda \not\in \Lambda
}{
\program H T P \rightarrow
\program H T {P\update i {\processor{R\update r{\lockbit 1}}{\Lambda}{I}}}
}
\\[0.5em]
\tag\mkRrule{unlock}
\frac{
P(i) = \processor R {\Lambda\uplus\{\lambda\}} {(\unlock\ v;I)} \quad
\operatorname{\hat R}(v) = l \quad
H(l) = \heaptuple{\_}{\lambda}
}{
\program H T P \rightarrow
\program{H\update l{\heaptuple{\lockbit 0}\lambda}}{T}{
P\update i {\processor R\Lambda I}}
}
\end{gather*}
\caption{Operational semantics (thread pool and locks).}
\label{fig:reduction-runqueue}
\end{myfigure}
\begin{myfigure}
\begin{gather*}
\tag\mkRrule{malloc}
\frac{
P(i) = \processor R \Lambda {(\malloci;I)}
\qquad
l \notin \operatorname{dom}(H)
}{
\program H T P \rightarrow
\program {H \update{l}{\heaptuple{?\pol \tau}\lambda}}
{T}
{P\update i {\processor{R\update r l}
{\Lambda}{I}}}
}
\\[0.5em]
\tag\mkRrule{load}
\frac{
P(i) = \processor R \Lambda {(\loadi;I)} \qquad
H(\operatorname{\hat R}(v)) = \heaptuple{v_1..v_n..v_{n+m}}\lambda \qquad
\lambda \in \Lambda
}{
\program H T P \rightarrow
\program H T {P\update i {\processor{R\update r{v_n}}{\Lambda}{I}}}
}
\\[0.5em]
\tag\mkRrule{store}
\frac{
P(i) = \processor R \Lambda {(\storei;I)} \qquad
R(r) = l \qquad
H(l) = \heaptuple{v_1..v_n..v_{n+m}}{\lambda} \qquad
\lambda \in \Lambda
}{
\program H T P \rightarrow
\program{H\update l{\heaptuple{v_1..\operatorname{\hat R}(v)..v_{n+m}}\lambda}}{T}{
P\update i {\processor R \Lambda I}}
}
\\[0.5em]
\tag\mkRrule{jump}
\frac{
P(i) = \processor R \Lambda \jumpi
\qquad
\operatorname{\hat R}(v) = l[\pol\lambda]
\qquad
H(l) = \heapcode{\forall[\pol\lambda'\colon(\_,\_)].\_}{I}
}{
\program H T P \rightarrow
\program H T {P\update i {\processor R \Lambda
{I\subs{\pol\lambda}{\pol\lambda'}}}}
}
\\[0.5em]
\tag\mkRrule{move}
\frac{
P(i) = \processor R \Lambda {(r \assign v;I)
}{
\program H T P \rightarrow
\program H T {P\update i {\processor {R\update{r}{\operatorname{\hat R}(v)}} \Lambda I}}
}
\\[0.5em]
\tag\mkRrule{arith}
\frac{
P(i) = \processor R \Lambda {(\aopi[r'];I)}
}{
\program H T P \rightarrow
\program H T {P\update i {\processor{R\update{r}{R(r') + \operatorname{\hat R}(v)}}{\Lambda}{I}}}
}
\\[0.5em]
\tag\mkRrule{branchT}
\frac{
P(i) = \processor R \Lambda {(\cjump{r}{v}{v'};\_)}
\qquad
R(r) = v
\qquad
\operatorname{\hat R}(v') = l[\pol\lambda]
\qquad
H(l) = \heapcode{\forall[\pol\lambda'\colon(\_,\_)].\_}{I}
}{
\program H T P \rightarrow
\program H T {P\update i {\processor R \Lambda {I\subs{\pol\lambda}{\pol\lambda'}}}}
}
\\[0.5em]
\tag\mkRrule{branchF}
\frac{
P(i) = \processor R \Lambda {(\cjump{r}{v}{\_};I)}
\qquad
R(r)\not= v
}{
\program H T P \rightarrow
\program H T {P\update i {\processor R \Lambda I}}
}
\end{gather*}
\caption{Operational semantics (memory and control flow).}
\label{fig:reduction-memory}
\end{myfigure}
\section{Operational Semantics and Deadlocked States}
\label{sec:semantics}
The operational semantics is defined in
Figures~\ref{fig:reduction-runqueue} and~\ref{fig:reduction-memory}.
The scheduling model of our machine is described by the first three
rules in Figure~\ref{fig:reduction-runqueue}.
The machine halts when all processors are idle and the thread pool is
empty (rule $\mkRrule{halt}$).
An idle processor (a processor that executes instruction $\milkw{done}$)
picks up an arbitrary thread from the thread pool and activates it
(rule $\mkRrule{schedule}$); the argument locks $\pol\lambda'$ replace the
parameters $\pol\lambda$ in the code for the processor.
For a $\milkw{fork}$ instruction, the machine creates a ``closure'' by
putting together the code label plus its arguments, $l[\pol \lambda]$,
and a copy of the registers, $R$, and by placing it in the thread
pool. The thread permission is partitioned in two: one part
($\Lambda$) stays with the thread, the other ($\Lambda'$) goes with
the newly created thread, as required by the type of its code.
Some rules rely on the evaluation function~$\operatorname{\hat R}$ that looks for
values in registers and in value application.
\begin{equation*}
\operatorname{\hat R}(v) =
\begin{cases}
R(v) & \text{if } v \text{ is a register}\\
\operatorname{\hat R}(v')[\lambda] & \text{if } v \text{ is } v'[\lambda]\\
v & \text{otherwise}
\end{cases}
\end{equation*}
In our model the heap tuple~$\tupletype{\lockbit 0}{\lambda}$
represents an open lock, whereas~$\tupletype{\lockbit 1}{\lambda}$
represents a closed lock. A lock is an uni-dimensional tuple holding
a \emph{lock value} because the machine provides for tuple allocation
only; lock $\lambda$ is used for type safety purposes, just like all
other singleton lock types.
Instruction~$\milkw{newLock}$ creates a new open lock in the heap and places
a reference $l$ to it in register~$r$.
Instruction $\milkw{testSetLock}$ loads the contents of the lock tuple into
register~$r$ and sets the heap value to~$\tupletype{\lockbit
1}{\lambda}$; it also makes sure that the lock is not in the thread's
permission (rules~$\mkRrule{tsl}0$ and~$\mkRrule{tsl}1$).
Further, applying the instruction to an unlocked lock adds
lock~$\lambda$ to the permission of the processor (rule~$\mkRrule{tsl}0$).
Locks are waved using instruction $\milkw{unlock}$, as long as the thread
holds the lock (rule~$\mkRrule{unlock}$).
Rules related to memory manipulation are described in
Figure~\ref{fig:reduction-memory}.
Rule~$\mkRrule{malloc}$ creates an heap-allocated $\lambda$-protected
uninitialised tuple and moves its address to register~$r$.
To store values in, and load from, a tuple we require that the lock
that guards the tuple is among the processor's permission.
In rules \mkRrule{branchT} and \mkRrule{branchF}, we ignore the lock annotation on
lock values, so that $\lockbit 0^\lambda$ is considered equal to
$\lockbit{0}$. The remaining rules are standard
(cf~\cite{m.walker.c.g:sytemf-to-tal}).
\paragraph{Deadlocked States}
The difficulty in characterising deadlock states stems from the fact
that processors never block and that threads may become (voluntary)
suspended while in critical a region. We aim at capturing conventional
techniques for acquiring locks, namely busy-waiting and
sleep-lock~\cite{vasconcelos.martins:multithreaded-tal}.
Towards this end, we need to restrict reduction of a given state $S$
to that of a single processor in order to control the progress of a
single core: let relation~$S \rightarrow_i S'$ denote a reduction step
on processor $i$ excluding rules \mkRrule{halt}, \mkRrule{schedule} and \mkRrule{unlock}.
\begin{defi}[Deadlocked states]
Let $S$ be the state $\program HTP$.
\begin{itemize}
\item A processor $\processor R\Lambda I$ holds lock $\lambda$ when
$\lambda\in\Lambda$; a suspend thread
$\poolitem{l[\vec\lambda']}{R}$ holds lock $\lambda$ when $H(l) =
\heapcode {\codetypeforall {\pol \lambda\colon(\_,\_)} {\_}
\Lambda} \_$ and $\lambda\in
\Lambda\subs{\pol\lambda'}{\pol\lambda}$;
\item A \emph{processor $p$ in $P$ immediately tries to enter a
critical section guarded by lock $\lambda$} if $p$ is of the
form $\processor{R}{\_}{(\cjumpi[\lockbit 0];\_)}$ and $R(r) = \lockbit
0^\lambda$;
\item For busy waiting, a \emph{thread in processor $p_i$ is
trying to enter a critical region guarded by $\lambda$} if $S
\rightarrow^*_i S'$ and processor $p_i$ in state~$S'$ immediately
tries to enter a critical section guarded by~$\lambda$;
\item For sleep-lock, a \emph{thread
$\poolitem{l[\pol\lambda']}{R}$ in thread pool $T$ is trying to enter a
critical region guarded by $\lambda$} if $ H(l) = \heapcode
{\codetypeforall {\pol\lambda\colon\_} {\_} \Lambda} I$, and
the thread in processor $p_1$ of state $S + P\{1\colon
\processor{R}{\Lambda}{I}\subs{\pol\lambda'}{\pol\lambda}\}$ is
trying to enter a critical region guarded by $\lambda$;
\item A state $S$ is \emph{deadlocked} if there exist locks
$\lambda_0,\dots,\lambda_n$, with $\lambda_0=\lambda_n$, and
indices~$d_0,\dots,d_{n-1}$ ($n>0$) such that for each $0\le i<n$,
either processor $p_{d_i}$ or suspended thread $t_{d_i}$ holds lock
$\lambda_i$ and is trying to enter a critical region guarded by
$\lambda_{i+1}$.
\end{itemize}
\end{defi}
Notice that $d_i\neq d_j$ does not imply $p_{d_i} \neq p_{d_j}$ and
similarly for threads in the thread pool, so that a state deadlocked
on locks $\lambda_0,\dots,\lambda_n$ may involve less than $n$
threads. We have excluded the \mkRrule{unlock} rule from the $\rightarrow_i$
reduction relation, yet releasing a lock is not necessarily an
indication that the thread is leaving a deadlocked state, for the
released lock may not be involved in the deadlock; a more general
definition of deadlocked state would take this fact into account.
\begin{myfigure}
\begin{gather*}
\frac{
\Psi \vdash \lambda\colon (\Lambda_1,\_)
\qquad
\lambda_1 \in \Lambda_1
}{
\Psi \vdash \lambda_1 \prec \lambda
}
\qquad
\frac{
\Psi \vdash \lambda\colon (\_,\Lambda_2)
\qquad
\lambda_2 \in \Lambda_2
}{
\Psi \vdash \lambda \prec \lambda_2
}
\qquad
\frac{
\Psi \vdash \lambda_1 \prec \lambda_2
\qquad
\Psi \vdash \lambda_2 \prec \lambda_3
}{
\Psi \vdash \lambda_1 \prec \lambda_3
}
\\
\frac{
\Psi\vdash\lambda_1\prec\lambda \quad\cdots\quad \Psi\vdash\lambda_n\prec\lambda
}{
\Psi\vdash\{\lambda_1,\dots,\lambda_n\} \prec \lambda
}
\qquad
\frac{
\Psi\vdash\lambda\prec\lambda_1 \quad\cdots\quad \Psi\vdash\lambda\prec\lambda_n
}{
\Psi\vdash \lambda \prec \{\lambda_1,\dots,\lambda_n\}
}
\end{gather*}
\caption{Less-than relation on locks and permissions}
\label{fig:less-than}
\end{myfigure}
\begin{myfigure}
\begin{gather*}
\tag{\mkTrule{type},\mkSrule{regFile}}
\frac{
\operatorname{ftv}(\tau) \subseteq \operatorname{dom}(\Psi)
}{
\Psi\vdash\tau
}
\qquad
\frac{
\Psi \vdash \tau_i
}{
\Psi
\vdash
\register 1\colon\tau_1,\dots,\register {n+m}\colon\tau_{n+m}
<:
\register 1\colon\tau_1,\dots,\register n\colon\tau_n
}
\\[0.5em]
\tag{\mkTrule{label},\mkTrule{reg},\mkTrule{int},\mkTrule{lock},\mkTrule{uninit}}
\frac{
\Psi \vdash \tau
}{
\Psi, l \colon \tau;\Gamma \vdash l\colon \tau
}
\qquad
\frac{
\Psi \vdash \tau
}{
\Psi; \Gamma_1,r_i\colon\tau,\Gamma_2 \vdash r_i\colon \tau
}
\qquad
\Psi;\Gamma \vdash n\colon \milkw{int}
\qquad
\Psi; \Gamma \vdash \lockbit 0,\lockbit 1,\lockbit 0^\lambda \colon \lambda
\qquad
\Psi; \Gamma \vdash \uninit \colon \tau
\\[0.5em]
\tag{\mkTrule{valApp}}
\frac{
\Psi\vdash\lambda'
\qquad
\Psi; \Gamma \vdash v \colon
\forall[\lambda\colon(\Lambda_1,\Lambda_2)]\tau
\qquad
\Psi \vdash \Lambda_1 \prec \lambda' \prec \Lambda_2
}{
\Psi;\Gamma \vdash \instantiation{\lambda'} \colon
\tau
\subs{\lambda'}{\lambda}
}
\end{gather*}
\caption{Rules for values
\myfbox{$\Psi;\Gamma \vdash v\colon\tau$}, for subtyping
\myfbox{$\Psi\vdash \Gamma <: \Gamma$}, and for types
\myfbox{$\Psi \vdash \tau$}.}
\label{fig:static-values}
\end{myfigure}
\begin{myfigure}
\begin{gather*}
\tag{\mkTrule{done}}
\Psi;\Gamma;\emptyset \vdash \done
\\[0.5em]
\tag{\mkTrule{fork}}
\frac{
\Psi;\Gamma \vdash v\colon \codetype{\Gamma'}{\Lambda}
\qquad
\Psi;\Gamma;\Lambda'\vdash I
\qquad
\Psi \vdash \Gamma <: \Gamma'
}{
\Psi;\Gamma;\Lambda\uplus\Lambda' \vdash \fork\ v;I
}
\\[0.5em]
\tag{\mkTrule{newLock}}
\frac{
\Psi,\lambda\colon(\Lambda_1, \Lambda_2);\Gamma\update{r}{\tupletype{\locktype}\lambda};
\Lambda \vdash I \qquad
\lambda \not \in \Psi, \Gamma, \Lambda
}{
\Psi;\Gamma;\Lambda \vdash \lambda\colon(\Lambda_1, \Lambda_2),r := \milkw{newLock}; I
}
\\[0.5em]
\tag{\mkTrule{tsl}}
\frac{
\Psi;\Gamma \vdash v\colon \tupletype\locktype\lambda
\qquad
\Psi;\Gamma\update{r}{\locktype};\Lambda \vdash I
\qquad
\lambda \not \in \Lambda
}{
\Psi;\Gamma;\Lambda \vdash \tsli; I
}
\\[0.5em]
\tag{\mkTrule{unlock}}
\frac{
\Psi;\Gamma \vdash v\colon\tupletype\locktype\lambda
\qquad
\Psi;\Gamma;\Lambda \vdash I
}{
\Psi;\Gamma;\Lambda\uplus\{\lambda\} \vdash \unlock\ v; I
}
\\[0.5em]
\tag{\mkTrule{critical}}
\frac{
\Psi;\Gamma \vdash r\colon \locktype
\qquad
\Psi;\Gamma \vdash v\colon\codetype{\Gamma'}{\Lambda\uplus\{\alpha\}}
\qquad
\Psi;\Gamma;\Lambda \vdash I
\qquad
\Psi \vdash \Gamma <: \Gamma'
\qquad
\Psi \vdash \Lambda \prec \lambda
}{
\Psi;\Gamma;\Lambda \vdash \cjumpi[\lockbit 0]; I
}
\end{gather*}
\caption{Typing rules for instructions (thread pool and locks)
\myfbox{$\Psi;\Gamma;\Lambda \vdash I$}.}
\label{fig:static-instructions-pool}
\end{myfigure}
\begin{myfigure}
\begin{gather*}
\tag{\mkTrule{malloc}}
\frac{
\Psi;
\Gamma\update r {\tupletype{\pol \tau}\lambda};\Lambda \vdash I
\qquad
\tau_i \neq \locktype[\lambda]
\qquad
\lambda \in \Lambda
}{
\Psi;\Gamma;\Lambda \vdash \malloci; I
}
\\[0.5em]
\tag{\mkTrule{load}}
\frac{
\Psi;\Gamma\vdash v\colon
\tupletype{\tau_1..\tau_{n+m}}{\lambda}
\qquad
\Psi;\Gamma\update r{\tau_n};\Lambda \vdash I
\qquad
\tau_n \neq \lambda'
\qquad
\lambda\in\Lambda
}{
\Psi;\Gamma;\Lambda \vdash \loadi;I
}
\\[0.5em]
\tag{\mkTrule{store}}
\frac{
\Psi;\Gamma \vdash v\colon\tau_n
\quad\;\;
\Psi;\Gamma \vdash r\colon
\tupletype{\tau_1..\tau_{n+m}}{\lambda}
\quad\;\;
\Psi;\Gamma\update{r}{\tupletype{\tau_1..\tau_{n+m}}{\lambda}};\Lambda \vdash I
\quad\;\;
\tau_n \neq \lambda'
\quad\;\;
\lambda\in\Lambda
}{
\Psi;\Gamma;\Lambda \vdash \storei; I
}
\\[-0.1em]
\tag{\mkTrule{move}}
\frac{
\Psi; \Gamma \vdash v\colon \tau
\qquad
\Psi;\Gamma\update r\tau;\Lambda \vdash I
}{
\Psi;\Gamma;\Lambda \vdash r \assign v;I
}
\\[0.5em]
\tag{\mkTrule{arith}}
\frac{
\Psi; \Gamma \vdash r'\colon \milkw{int}
\qquad
\Psi;\Gamma\vdash v\colon \milkw{int}
\qquad
\Psi;\Gamma\update r\milkw{int};\Lambda \vdash I
}{
\Psi;\Gamma;\Lambda \vdash \aopi[r'];I
}
\\[0.5em]
\tag{\mkTrule{branch}}
\frac{
\Psi; \Gamma \vdash r\colon \milkw{int}
\qquad
\Psi; \Gamma \vdash v\colon \milkw{int}
\qquad
\Psi; \Gamma \vdash v\colon\codetype{\Gamma}{\Lambda}
\qquad
\Psi;\Gamma;\Lambda \vdash I
}{
\Psi;\Gamma;\Lambda \vdash \cjumpi[v];I
}
\\[0.5em]
\tag{\mkTrule{jump}}
\frac{
\Psi;\Gamma \vdash v\colon\codetype{\Gamma'}{\Lambda}
\qquad
\Psi \vdash \Gamma <: \Gamma'
}{
\Psi;\Gamma;\Lambda \vdash \jumpi
}
\end{gather*}
\caption{Typing rules for instructions (memory and control flow)
\myfbox{$\Psi;\Gamma;\Lambda \vdash I$}.}
\label{fig:static-instructions-memory}
\end{myfigure}
\begin{myfigure}
\begin{gather*}
\tag{reg file, \fbox{$\Psi \vdash R \colon \Gamma$}}
\frac{
\forall i.
\Psi \vdash \Gamma(\register i)
\qquad
\Psi;\emptyset \vdash R(\register i)\colon \Gamma(\register i)
}{
\Psi\vdash R\colon\Gamma
}
\\[0.5em]
\tag{processors, \fbox{$\Psi \vdash P$}}
\frac{
\forall i.\Psi \vdash P(i)
}{
\Psi \vdash P
}
\qquad
\frac{
\Psi \vdash R\colon\Gamma
\qquad
\Psi;\Gamma';\Lambda \vdash I
\qquad
\Psi \vdash \Gamma <: \Gamma'
}{
\Psi \vdash \processor{R}{\Lambda}{I}
}
\\[0.5em]
\tag{thread pool, \fbox{$\Psi \vdash T$}}
\frac{
\forall i.
\Psi \vdash t_i
}{
\Psi \vdash \{t_1,\dots,t_n\}
}
\qquad
\frac{
\Psi;\emptyset\vdash v\colon
\codetype {\Gamma'} \_
\qquad
\Psi\vdash R\colon \Gamma
\qquad
\Psi \vdash \Gamma <: \Gamma'
}{
\Psi \vdash \poolitem{v}{R}
}
\\[0.5em]
\tag{heap value, \fbox{$\Psi \vdash h\colon\tau$}}
\frac{
\tau = \codetypeforall{\pol\lambda\colon(\pol \Lambda_1, \pol \Lambda_2)}{\Gamma}{\Lambda}
\qquad
\Psi, \pol\lambda\colon(\pol \Lambda_1, \pol \Lambda_2);\Gamma;\Lambda \vdash I
}{
\Psi \vdash \tau \{I\} \colon \tau
}
\qquad
\frac{
\forall i.\Psi;\emptyset \vdash v_i \colon \tau_i
}{
\Psi \vdash \heaptuple{\pol v}{\lambda} \colon
\tupletype{\pol \tau}\lambda
}
\\[0.5em]
\tag{heap, \fbox{$\Psi \vdash H$}}
\frac{
\forall l. \Psi \vdash H(l) \colon \Psi(l)
}{
\Psi \vdash H
}
\\[0.5em]
\tag{state, \fbox{$\Psi \vdash S$}}
\Psi \vdash \milkw{halt}
\qquad
\frac{
\Psi \vdash H \qquad \Psi \vdash T \qquad \Psi \vdash P
}{
\Psi \vdash \program HTP
}
\end{gather*}
\caption{Typing rules for machine states.}
\label{fig:static-machine}
\end{myfigure}
\section{A Type System for Deadlock Prevention}
\label{sec:typing}
\myparagraph{Type System}
Typing environments $\Psi$ map heap addresses $l$ to types $\tau$, and
singleton lock types $\lambda$ to lock kinds
$(\Lambda_1,\Lambda_2)$. An entry $\lambda\colon(\Lambda_1,\Lambda_2)$
in $\Psi$ means that $\lambda$ is larger than all lock types in
$\Lambda_1$ and smaller than any lock type in $\Lambda_2$, a notion
captured by relation $\prec$ described in Figure~\ref{fig:less-than}.
Instructions are also checked against a register file type $\Gamma$
holding the current types of the registers, and a set $\Lambda$ of
lock variables: the \emph{permission} of (the processor executing) the
code block.
The type system is presented in Figures~\ref{fig:static-values}
to~\ref{fig:static-machine}.
Typing rules for values are illustrated in
Figure~\ref{fig:static-values}.
Rule $\mkTrule{type}$ makes sure types are \emph{well-formed}, that all free
singleton lock types (or free type variables, $\operatorname{ftv}$) in a type are
bound in the typing environment.
A formula $\Gamma <: \Gamma'$ allows ``forgetting'' registers in
the register file type, and is particularly useful in jump
instructions where we want the type of the target code block to be
more general (ask for less registers) than those active in the current
code~\cite{m.walker.c.g:sytemf-to-tal}.
The rule for value application, \mkTrule{valApp}, checks that the argument
$\lambda'$ is within the interval $(\Lambda_1,\Lambda_2)$, as required
by the parameter $\lambda$.
The rules in Figure~\ref{fig:static-instructions-pool} capture the
policy for lock usage.
Rule $\mkTrule{done}$ requires the release of all locks before terminating the
thread.
Rule $\mkTrule{fork}$ splits permissions into sets $\Lambda$ and $\Lambda'$:
the former is transferred to the forked thread according to the
permissions required by the target code block, the latter remains with
the processor.
Rule $\mkTrule{newLock}$ assigns a lock type $\tupletype{\locktype}\lambda$ to
the register.
The new singleton lock type~$\lambda$ is recorded in
$\Psi$, so that it may be used in the rest of the instructions~$I$.
Rule $\mkTrule{tsl}$ requires that the value under test is a lock in the heap
(of type~$\tupletype{\locktype}\lambda$) and records the type of the
lock value~$\lambda$ in register~$r$.
This rule also disallows testing a lock already held by the processor.
Rule $\mkTrule{unlock}$ makes sure that only held locks are unlocked.
Rule $\mkTrule{critical}$ ensures that the processor holds the permission
required by the target code block, including the lock under test.
A processor is guaranteed to hold the tested lock only after
(conditionally) jumping to the critical region.
A previous test-and-set-lock instructions may have obtained the lock,
but the type system records that the processor holds the lock only
after the conditional jump.
The rule checks that the newly acquired lock is larger than all locks
in the possession of the thread.
The typing rules for memory and control flow are depicted in
Figure~\ref{fig:static-instructions-memory}.
Operations for loading from~($\mkTrule{load}$), and for storing
into~($\mkTrule{store}$), tuples require that the processor holds the right
permissions (the locks for the tuples it reads from, or writes to).
Both rules preclude the direct manipulation of lock values by
programs, via the $\tau_n\neq\lambda'$ assumptions.
The rules for typing machine states are illustrated in
Figure~\ref{fig:static-machine}.
The rule for a thread item in the thread pool checks that the type and
required registers~$R$ are as expected in the type of the code block
pointed by~$v$.
Similarly, the rule for type checking a processor also permits that
type~$\Gamma$ of the registers~$R$ be more specific than the register
file type~$\Gamma'$ required to type check the remaining
instructions~$I$.
The heap value rule for code blocks adds to $\Psi$ each singleton lock
type (together with its bounds), so that they may be used in the rest
of the instructions~$I$.
\myparagraph{Example}
As expected, the example is not typable with the annotations
introduced previously. The three \lstinline|newLock| instructions
place in $\Psi$ three entries $f_1\colon(\emptyset, \emptyset)$,
$f_2\colon(\{f_1\},\{f_3\})$,
$f_3\colon(\{f_1\},\emptyset)$. Then the value
\lstinline|(liftLeftFork[f2])[f1]|
(in the example: \lstinline|liftLeftFork[f1,f2]|)
in the first \lstinline|fork| instruction issues goals $\Psi\vdash
\emptyset \prec f_1 \prec \emptyset$ and $\Psi\vdash \{f_1\} \prec f_2
\prec \emptyset$, which are easy to guarantee given that $\Psi$
contains an entry $f_2\colon(\{f_1\},\{f_3\})$.
Likewise, the second \lstinline|fork| instruction, generates goals
$\Psi\vdash \emptyset \prec f_2 \prec \emptyset$ and $\Psi\vdash \{f_2\}
\prec f_3 \prec \emptyset$, which are again hold because of same entry.
However, the last \lstinline|fork| instruction requires $\Psi\vdash
\emptyset \prec f_3 \prec \emptyset$ and $\Psi\vdash \{f_3\} \prec f_2
\prec \emptyset$, the second of which does not hold.
Notice however that each of the three \lstinline|jump| instructions
are typable per se. For example, in code block
\lstinline|liftRightFork|, instruction
\lstinline|if r3 = 0b jump eat[l,m]| requires $\Psi \vdash \{l\}
\prec m$, which holds because the signature for the code block
includes the annotation $m\colon(\{l\},\emptyset)$.
\myparagraph{Typable States Do Not Deadlock}
The main result of the type system, namely that $\Psi \vdash S$ and
$S\rightarrow^* S'$ implies $S'$ not deadlocked, follows from Subject
Reduction and from Typable States Are Not Deadlocked, in a
conventional manner.
\begin{lem}[Substitution Lemma]
If $\Psi,\lambda\colon(\Lambda_1,\Lambda_2); \Gamma,
\Lambda \vdash I$ and $\Psi(\lambda)'= (\Lambda_1,\Lambda_2)$,
then $\Psi; \Gamma\sigma, \Lambda\sigma \vdash I\sigma$, where
$\sigma= \subs{\lambda'}{\lambda}$.
\end{lem}
\begin{thm}[Subject Reduction]
\label{thm:sr}
If $\Psi \vdash S$ and $S\rightarrow S'$, then $\Psi' \vdash S'$,
where $\Psi' = \Psi$ or $\Psi' =
\Psi,l\colon\heaptuple{\pol\tau}{\lambda}$ (with $l$ fresh) or
$\Psi' = \Psi,l\colon \heaptuple{\lambda}{\lambda},
\lambda\colon(\pol\Lambda_1,\pol\Lambda_2)$ (with $l,\lambda$
fresh).
\end{thm}
\begin{proof}
(Outline) By induction on the derivation of $S \rightarrow S'$ proceeding by case analysis
on the last rule of the derivation, using the substitution lemma for rules \mkRrule{schedule},
\mkRrule{fork}, \mkRrule{jump}, and \mkRrule{branchT}, as well as weakening in several rules.
\end{proof}
\begin{thm}[Typable States Are Not Deadlocked]
\label{thm:ts}
If $\Psi \vdash S$, then $S$ is not deadlocked.
\end{thm}
\begin{proof}
(Sketch) Consider the contra-positive and show that deadlocked states are not
typable. Without loss of generality suppose that $S$ is of the form
$\program{H}{\langle t_{d_0},\dots,t_{d_m}\rangle}{\{d_{m+1}\colon
p_{d_{m+1}}, \dots, d_n\colon p_{d_n}\}}$ with suspended threads
$t_{d_i}$ and processors $p_{d_j}$ not necessarily distinct.
Each of these threads and processors are trying to enter a critical
region.
%
For a processor $p_{d_i}$ we have that $S \rightarrow^*_{d_i} S'$
where $d_i$-th processor in $S'$ is of the form
$\processor{R}{\Lambda}{(\cjumpi[\lockbit 0];\_)}$ and $R(r) =
\heaptuple{\_}{\lambda^{d_{i+1}}}$. By Subject Reduction $\Psi'
\vdash S'$ where $\Psi'$ extends $\Psi$ as stated in
Theorem~\ref{thm:sr}. A simple derivation starting from rule
\mkTrule{critical} allows to conclude that $\Psi \vdash \Lambda_{d_i} \prec
\lambda_{d_{i+1}}$.
%
For a thread $t_{d_i} = \poolitem{l[\pol\lambda']}{R}$ in thread pool
of~$S$ we run the machine $S''$ obtained from $S$ by replacing
processor 1 with $\processor{R}{\Lambda}{I}
\subs{\pol\lambda'}{\pol\lambda}$, where $H(l) = \heapcode
{\codetypeforall {\pol\lambda\colon\_} {\_} {\Lambda}}
{I}$. Proceeding as for processes above, we conclude again that
$\Psi \vdash \Lambda_{d_i} \prec \lambda_{d_{i+1}}$.
We thus have $\Psi \vdash \Lambda_{d_0} \prec \lambda_{d_1}$, \dots
$\Psi \vdash \Lambda_{d_{n-1}} \prec \lambda_{d_n}$, $\Psi \vdash
\Lambda_{d_n} \prec \lambda_{d_0}$ which is not satisfiable.
\end{proof}
\section{Type Inference}
\label{sec:type-inference}
Annotating lock ordering on large assembly programs may not be an easy
task. In our setting, programmers (compilers, more often) produce
annotation free programs such as the one in
Figure~\ref{fig:philosophers}, and use an inference algorithm to
provide for the missing annotations.
\myparagraph{The Algorithm}
The \emph{annotation-free syntax} is obtained from that in
Figure~\ref{fig:syntax-instructions}, by removing the
$\colon(\Lambda,\Lambda)$ part both in the \milkw{newLock}{} instruction and
in the universal type.
Given an annotation-free program~$H$, algorithm $\mathcal W$ produces
a pair, comprising a typing environment $\Psi$ and an annotated
program~$H^\star$, such that $\Psi \vdash H^\star$, or else fails.
In the former case $H^\star$ is typable, hence does not deadlock
(Theorem~\ref{thm:sr}); in the latter case, there is no possible
labeling for~$H$.
We depend on a set of \emph{variables over permissions} (sets of
locks), ranged over by~$\nu$, disjoint from the set of heap labels and
from the set singleton lock types introduced in
Section~\ref{sec:syntax}.
Constraints are computed by an intermediate step in our algorithm.
\begin{defi}[Constraints and solutions]~
\begin{itemize}
\item We consider \emph{constraints} of three distinct forms:
$\Lambda \prec \lambda$, $\nu \prec \lambda$, and $\lambda \prec
\nu$, and denote by $C$ a set of constraints;
\item A \emph{substitution $\theta$} is a map from permission
variables $\nu$ to permissions $\Lambda$;
\item A substitution $\theta$ \emph{solves} $(\Psi, C)$ if
$\Psi\theta \vdash x\theta \prec y\theta$ for all $x\prec y \in
C$.
\end{itemize}
\end{defi}
Algorithm $\mathcal W$ runs in two phases: the first, $\mathcal A$,
produces a triple comprising a typing environment~$\Psi$, an annotated
program~$H^\star$, and a collection of constraints $C$, all containing
variables over permissions $\Lambda$. The set of constraints is then
passed to a constraint solver, that either produces a
substitution~$\theta$ or fails.
In the former case, the output of $\mathcal W$ is the pair
$(\Psi\theta, H^\star\theta)$; in the latter $\mathcal W$ fails.
In practice, we do not need to generate $H^\star$ or to perform the
substitutions; our compiler accepts $H$ if the produced collection of
constraints is solvable, and rejects it otherwise.
\begin{myfigure}
\begin{align*}
&\tagA[ \{l_i \colon \heapcode {\tau_i} {I_i}\}_{i\in I}] =
(\cup_{i\in I}\Psi_i, \{l_i \colon
\heapcode {{\tau_i^\star}} {I_i^\star}
\}_{i\in I}, \cup_{i\in I} C_i)
\\
&\quad\text{where }
\tau_i^\star = \codetypeforall{\pol\lambda_i {\colon
(\pol{\nu_i},\pol{\rho_i})}}{\Gamma_i}{\Lambda_i} = \tagT {\tau_i}
\\
&\quad\text{and }
(I_i^\star, \Psi_i , C_i) = \mathcal I(I_i ,
\{l_i\colon {\tau_i^\star}\}_{i\in I} \cup\{\pol\lambda_i\colon
(\pol{\nu_i},\pol{\rho_i})\},\Gamma_i,\Lambda_i)
\\
&\tagId {(\lambda,r := \milkw{newLock} ;I)} =
((\lambda{\colon(\nu,\rho)},r
:= \milkw{newLock}; I^\star), \Psi', C)
\\
&\quad \text{where}\;\;
(I^\star, \Psi', C) = \tagI I {\Psi\uplus\{\lambda{\colon(\nu,\rho)}\}}
{\Gamma\update{r}{\tupletype{\lambda}{\lambda}}} \Lambda
\\
&\tagId {(\cjumpi[\lockbit 0]; I)} =
((\cjumpi[\lockbit 0]; I^\star), \Psi', C_1 \cup C_2 \cup \{{\Lambda \prec \lambda}\})
\\
&\quad\text{where }
(\codetype{\Gamma'}{(\Lambda\uplus\{\lambda\})}, C_1) = \tagV{v}{\Psi}{\Gamma}
\\
&\quad\text{and } (I^\star, \Psi', C_2) = \tagId I
\\
&\quad\text{and }
\Psi \vdash \Gamma <: \Gamma'
\\
&\quad\text{and } \lambda = \Gamma(r)
\\
&\tagI {(\fork\ v; I)} \Psi \Gamma {\Lambda} =
((\fork\ v;I^\star), \Psi',C_1 \cup C_2)\\
&\quad\text{where }
(\codetype{\Gamma'}{\Lambda'}, C_1) = \tagV{v}{\Psi'}{\Gamma}
\\
&\quad\text{and }
(I^\star, \Psi', C_2) = \tagI I {\Psi} {\Gamma} {\Lambda\setminus\Lambda'}
\\
&\quad\text{and }
\Psi \vdash \Gamma <: \Gamma'
\\
&\tagV {\instantiation {\lambda} } \Psi \Gamma =
(
\tau
\subs{\lambda}{\lambda'}, C \cup \{{\nu
\prec \lambda \prec \rho}\})
\\
&\quad\text{where } (
\forall[\lambda'\colon(\nu,\rho)]\tau,
C) = \tagV v \Psi \Gamma
\\
& \tagT { \codetypeforall {\pol\lambda} {(\register
1\colon\tau_1,...,\register n\colon\tau_n)} {\Lambda} } =
\codetypeforall { \tagLock{\pol\lambda}{\pol\nu}{\pol\rho} }
{(\register 1\colon\tagT{\tau_1},...,\register
n\colon\tagT{\tau_n})} {\Lambda}
\end{align*}
\caption{The tagging algorithm (selected rules).}
\label{fig:algorithm}
\end{myfigure}
\myparagraph{Generating constraints}
Algorithm $\mathcal A$, described in Figure~\ref{fig:algorithm},
visits the program twice.
On a first step it builds an initial type environment $\Psi_0 =
\{l_i\colon {\tau_i^\star}\}_{i\in I}$ collecting the types for all
code blocks in the given program $\{l_i \colon \heapcode {\tau_i}
{I_i}\}_{i\in I}$, annotating with permission variables (denoted by
$\nu$ and $\rho$) the intervals for the locks bound in forall types;
on a second visit it generates the constraints and the annotated
syntax for the instructions in each code block.
The algorithm for instructions, $\mathcal I$, also shown in
Figure~\ref{fig:algorithm}, generates annotations for the singleton
lock type introduced in \lstinline|newLock| instructions, or further
constraints in the case of the jump-to-critical instruction.
In the case of a fork instruction, the algorithm calls function
$\mathcal V$ to obtain the required permission $\Lambda'$ and passes
the difference $\Lambda\setminus\Lambda'$ to the function that
annotates the continuation $I$.
The algorithm for values, $\mathcal V$, generates constraints in the
case of type application.
Finally, the algorithm for types annotates the singleton lock types in
forall types.
In the definition of all algorithms, permission-variables
$\nu,\rho,\pol\nu_i,\pol\rho_i$ are freshly introduced.
\myparagraph{Example}
For the running example, we first rename all bound variables so that
the type of code block \lstinline|liftLeftFork| mentions
\lstinline|l1| and \lstinline|m1|, that of \lstinline|liftLeftFork|
mentions \lstinline|l2| and \lstinline|m2|, and that of
\lstinline|eat| uses \lstinline|l3| and \lstinline|m3|.
For example:
\begin{lstlisting}
liftRightFork forall[l2].forall[m2].(r1:<l2>^L2, r2:<m2>^M2) requires {l2}
\end{lstlisting}
Then, algorithm $\mathcal A$ creates an initial environment $\Psi_0$
by generating twelve variables ($\rho_1$ to $\rho_{12}$) to annotate
the six locks ($\mathsf{l_i}$ and $\mathsf{m_i}$) in the three code
blocks that mention locks (\lstinline|liftLeftFork|,
\lstinline|liftRightFork|, and \lstinline|eat|). They are
$\mathsf{l_1}\colon(\rho_1,\rho_2)$, \dots,
$\mathsf{m_3}\colon(\rho_{11},\rho_{12})$.
Revisiting the signature of code block~\lstinline|liftRightFork|, we get:
\begin{lstlisting}
liftRightFork forall[m2::(rho7,rho8)].forall[l2::(rho5,rho6)].(r1:<l2>^L2, r2:<m2>^M2) requires {l2}
\end{lstlisting}
In the second pass, while in code block \lstinline|main|, algorithm
$\mathcal I$ generates six more permission variables ($\rho_{13}$ to
$\rho_{18}$) to annotate the new lock variables $\mathsf{f_1}$ to
$\mathsf{f_3}$ introduced with the \lstinline|newLock|
instructions. They are: $\mathsf{f_1}\colon(\rho_{13},\rho_{14})$
\dots $\mathsf{f_3}\colon(\rho_{17},\rho_{18})$.
The rest of the second pass generates new constraints in type
application and in jump-to-critical instructions. For example, in code
block \lstinline|liftRightFork|, and for value
\lstinline|eat[l2,m2]|, four constraints are generated: $\rho_9\prec
\mathsf{l_2}\prec \rho_{10}, \rho_{11} \prec \mathsf{m_2} \prec
\rho_{12}$. Then, in the jump-to-critical instruction,
\lstinline|if r3 = 0 jump eat[l2,m2]|, and since the thread holds lock
\lstinline|l2| (as witnessed by its signature
\lstinline|requires (l2)|), a new constraint $\{\mathsf{l_2}\} \prec
\mathsf{m_2}$ is generated.
The thus created set of constraints is then passed to a constraint
solver, which is bound to fail.
\myparagraph{Main result}
For soundness we start with a few lemmas.
\begin{lem}[Value soundness]
\label{lem:value-soundness}
If $\tagV v \Psi \Gamma = (\tau, C)$ and $\theta$ solves $(\Psi,C)$ then
$\Psi\theta; \Gamma\theta \vdash v\colon \tau\theta$.
\end{lem}
\begin{proof}
(Outline) The proof proceeds by induction on the inference tree for
$\Psi\theta; \Gamma\theta \vdash v\colon \tau\theta$ performing case
analysis on the last typing rule applied.
\end{proof}
\begin{lem}[Instruction soundness]
\label{lem:instr-soundness}
If $\tagId I = (I^\star, \Psi', C)$ and $\theta$ solves $(\Psi',C)$ then
$\Psi'\theta; \Gamma\theta; \Lambda\theta \vdash I^\star\theta$ and
$\Psi \subseteq \Psi'$.
\end{lem}
\begin{proof}
(Outline) The proof proceeds by induction on $I$. The cases for
conditional jump and fork use Lemma~\ref{lem:value-soundness}.
\end{proof}
\begin{thm}[Soundness]
\label{thm:soundness}
If $\mathcal W(H) = (\Psi, H^\star)$ then $\Psi \vdash H^\star$.
\end{thm}
\begin{proof}
(Outline) Follows directly from Lemma~\ref{lem:instr-soundness}
using typing rules for heap values and heaps. We use weakening on
typing environments before applying the heap rule.
\end{proof}
Conversely, we believe that if $\Psi \vdash H^\star$,
then $\mathcal W(\mathcal{E}(H^\star))$ does not fail, where $\mathcal{E}$ is
the obvious lock-order annotation erasure function. A stronger result
would include a notion of principal solutions.
\section{Related Work and Conclusion}
\label{sec:conclusion}
\myparagraph{Related work}
The literature on type systems for deadlock freedom in lock-based
languages is vast; space restrictions prohibit a general survey. We
however believe that the problem of type inference for deadlock
freedom in lock-based languages has been given not so much attention
in high-level languages, let alone low-level (assembly) languages.
Three characteristics separate our work from most proposals on the
topic: the non block structure of the locking primitives, the facts
that threads never block and that they may be suspended while holding
locks.
Following Coffman \textit{et al.}\@\xspace one can classify the problem of deadlock under
the categories of \emph{detection and recovery},
\emph{avoidance} and \emph{prevention }~\cite{DBLP:conf/ictac/Boudol09,coffman.etal:system-deadlocks}.
In the first category, detection and recovery, on finds for example
works that check deadlocks at runtime.
Cunningham \textit{et al.}\@\xspace infer locks for atomicity in an object-oriented
language, but use a runtime mechanism to detect when a thread's lock
acquisition would cause a
deadlock~\cite{cunningham.etal:lock-inference-proven}.
Java PathFinder~\cite{brat.etal:javapathfinder} and Driver
Verifier~\cite{ball.elal:thorough-static-analysis} identify violations
of the lock discipline during runtime tests. Agarwal
\textit{et al.}\@\xspace~\cite{agarwal.etal:run-time-detection-deadlocks,agarwal.etal:detecting-deadlocks}
present an algorithm that detects potential deadlocks involving any
number of threads.
Under the \emph{avoidance} category on finds, e.g., a recent work by
Boudol where a type and effect system allows for the design of an
operational semantics that refuses to lock a pointer whenever it
anticipates to take a pointer that is held by another
thread~\cite{DBLP:conf/ictac/Boudol09}.
Our work falls into the third category above, \emph{prevention}.
Flanagan and Abadi present a functional language with mutable
references where locking is block structured and threads physically
block~\cite{flanagan.abadi:types-safe-locking}. From this work we
borrowed the idea of singleton lock types to describe, at the type
level, a single lock.
Type based deadlock prevention has also been study in the realm of
object-oriented languages, where, e.g., Boyapati \textit{et al.}\@\xspace use a variant
of ownership types for preventing deadlocks in Java, performing
partial inference of annotations, but not of those related to lock
order~\cite{boyapati.lee.rinard:preventing-data-races}.
Suenaga proposes a concurrent functional language similar to Flanagan
and Abadi's mentioned above, except that it features non block
structured locking~\cite{DBLP:conf/aplas/Suenaga08}; his language
includes separate primitives for locking/unlocking, as in our case.
Albeit targeting at different level of abstraction, the results
(deadlock prevention) and the techniques (type inference) in both
works are similar, in particular the usage of a constraint-based
algorithm to infer types.
Differently from our case, Suenaga uses ownership types rather
than singleton lock types.
\myparagraph{Concluding remarks}
We have presented a type system that enforces a strict partial order
on lock acquisition, guaranteeing that well typed programs do not
deadlock. Towards this end we extended the syntax of our language to
incorporate annotations on the locking order. Acknowledging that the
annotation of large assembly programs (either manually or as the
result of a compilation process) is not plausible, we have introduced
an algorithm that infers the required annotations. The algorithm is
proved to be correct, hence that programs that pass our compiler are
exempt from deadlocks.
The current implementation of the algorithm generates, from a
non-annotated program, a set of constraints in the form of a Prolog
goal. The goal is then checked against a Prolog program that
implements the $\prec$ relation in Figure~\ref{fig:less-than}. We
consider the program typable if the goal succeeds. There is no point
in building the annotated syntax or performing the substitution, as
explained in Section~\ref{sec:type-inference}. Future work in this
area includes the automation of the whole process either by calling
the Prolog interpreter from within the compiler, or by implementing
relation $\prec$ directly in Java, the language of our type
checker/interpreter.
Future work also includes trying to assess the usage of our type
checker on larger programs, generated for example from an imperative
high-level language, and to further compare the singleton lock types
and the ownership types approaches for the description of non block
structured locking.
|
\chapter{Tangent Myller configurations $\mathfrak{M}_{t}$}
The theory of Myller configurations
$\mathfrak{M}(C,\overline{\xi},\pi)$ presented in the Chapter 2
has an important particular case when the tangent versor fields
$\overline{\alpha}(s)$, $\forall s\in (s_{1},s_{2})$ belong to the
corresponding planes $\pi(s)$. These Myller configurations will be
denoted by $\mathfrak{M}_{t} = \mathfrak{M}_{t}(C,
\overline{\xi},\pi)$ and named {\bf tangent} {\it Myller configuration}.
The geometry of $\mathfrak{M}_{t}$ is much more rich that the
geometrical theory of $\mathfrak{M}$ because in $\mathfrak{M}_{t}$
the tangent field has some special properties. So,
$(C,\overline{\alpha})$ in $\mathfrak{M}_{t}$ has only three
invariants $\kappa_{g}, \kappa_{n}$ and $\tau_{g}$ called {\it geodesic
curvature}, {\it normal curvature} and {\it geodesic torsion}, respectively, of
the curve $C$ in $\mathfrak{M}_{t}$.
The curves $C$ with $\kappa_{g} = 0$ are {\it geodesic lines} of
$\mathfrak{M}_{t}$; the curves $C$ with the property $\kappa_{n}=0$
are the {\it asymptotic lines} of $\mathfrak{M}_{t}$ and the curve $C$
for which $\tau_{g} = 0$ are the {\it curvature lines} for
$\mathfrak{M}_{t}$. The mentioned invariants have some geometric
interpretations as the geodesic curvature, normal curvature and
geodesic torsion of a curve $\cal{C}$ on a surfaces $S$ in the
Euclidean space $E_{3}$.
\section{The fundamental equations of $\mathfrak{M}_{t}$}
\setcounter{theorem}{0}\setcounter{equation}{0}
Consider a tangent Myller configuration $\mathfrak{M}_{t} =
(C,\overline{\xi},\pi)$. Thus we have
\begin{equation}
\langle \alpha(s), \nu(s)\rangle = 0,\; \forall s\in (s_{1},s_{2}).
\end{equation}
The Darboux frame $\cal{R}_{D}$ is $\cal{R}_{D}=(P(s);
\overline{\xi}(s), \overline{\mu}(s),\overline{\nu}(s))$ with
$\overline{\mu}(s) = \overline{\nu}(s)\times \overline{\xi}(s)$.
The fundamental equations of $\mathfrak{M}_{t}$ are obtained by
the fundamental equations (2.1.3), (2.1.4), Chapter 2 of a general
Myller configuration $\mathfrak{M}$ for which the invariant
$c_{3}(s)$ vanishes.
\begin{theorem}
The fundamental equations of the tangent Myller configuration
$\mathfrak{M}_{t}(C, \overline{\xi}, \pi)$ are given by the
following system of differential equations:
\begin{equation}
\displaystyle\frac{d\overline{r}}{ds} = c_{1}(s)\overline{\xi}(s) +
c_{2}(s)\overline{\mu}(s),\;\; (c_{1}^{2}+c_{2}^{2} = 1),
\end{equation}
\begin{eqnarray}
\displaystyle\frac{d\overline{\xi}}{ds} &=&
G(s)\overline{\mu}(s)+K(s)\overline{\nu}(s)\nonumber\\
\displaystyle\frac{d\overline{\mu}}{ds}& =& -G(s)\overline{\xi}(s) +
T(s)\overline{\nu}(s)\\\displaystyle\frac{d\overline{\nu}}{ds} &=&
-K(s)\overline{\xi}(s) -T(s)\overline{\mu}(s).\nonumber
\end{eqnarray}
\end{theorem}
The invariants $c_{1},c_{2},G, K,T$ have the same geometric
interpretations and the same denomination as in Chapter 2. So,
$G(s)$ is {\it the geodesic curvature} of the field
$(\overline{C},\overline{\xi})$ in $\mathfrak{M}_{t}$, $K(s)$ is
{\it the normal curvature} and $T(s)$ is {\it the geodesic
torsion} of $(C,\overline{\xi})$ in $\mathfrak{M}_{t}$.
The cases when some invariants $G, K, T$ vanish can be investigate
exactly as in the Chapter 2.
In this respect, denoting $\varphi = \sphericalangle
(\overline{\xi}_{2}(s),\overline{\nu}(s))$ and using the Frenet
formulae of the versor field $(C,\overline{\xi})$ we obtain the
formulae
\begin{equation}
G = K_{1}\sin \varphi,\;\; K = K_{1}\cos \varphi, \;\; T = K_{2}
+\displaystyle\frac{d\varphi}{ds}.
\end{equation}
In \S 5, Chapter 2 we get the relations between the invariants of
$(C,\overline{\xi})$ in $\mathfrak{M}_{t}$ and the invariants of
normal versor field $(C,\overline{\nu})$, (Theorem 2.5.1, Ch 2.) For
$\sigma = \sphericalangle (\overline{\xi}(s), \overline{\nu}_{3}(s))$
we have
\begin{equation}
K = \chi_{1}\sin \sigma,\;\; T = \chi_{1}\cos \sigma,\;\; G = \chi_{2}
+\displaystyle\frac{d\sigma}{ds}.
\end{equation}
Others results concerning $\mathfrak{M}_{t}$ can be deduced from
those of $\mathfrak{M}$.
For instance
\begin{theorem}
(Mark Krein) Assuming that we have:
$1.$ $\mathfrak{M}_{t}(C, \overline{\xi}, \pi)$ a tangent Myller
configuration of class $C^{3}$ in which $C$ is a closed curve,
having $s$ as natural parameter.
$2.$ The spherical image $C^{*}$ of $\mathfrak{M}_{t}$ determines
on the support sphere $\Sigma$ a simply connected domain of area
$\omega$.
$3.$ $\sigma = \sphericalangle (\overline{\xi}, \overline{\nu}_{3})$.
\end{theorem}
{\it In these conditions the following Mark-Krein$'$s formula}
holds:
\begin{equation}
\omega = 2\pi - \int_{C}G(s)ds +\int_{C}d\sigma.
\end{equation}
\section{The invariants of the curve $C$ in $\mathfrak{M}_{t}$}
\setcounter{theorem}{0}\setcounter{equation}{0}
A smooth curve $C$ having $s$ as arclength determines the tangent
versor field $(C,\overline{\alpha})$ with $\overline{\alpha}(s) =
\displaystyle\frac{d\overline{r}}{ds}$. Consequently we can consider a
particular tangent Myller configuration
$\mathfrak{M}_{t}(C,\overline{\alpha},\pi)$ defined only by curve $C$
and tangent planes $\pi(s)$.
In this case the geometry of Myller configurations
$\mathfrak{M}_{t}(C,\overline{\alpha},\pi)$ is called the geometry of
curve $C$ in $\mathfrak{M}_{t}$. The Darboux frame of curve $C$ in
$\mathfrak{M}_{t}$ is $\cal{R}_{D} = (P(s); \overline{\alpha}(s),
\overline{\mu}^{*}(s), \overline{\nu}(s))$, $\overline{\mu}^{*}(s)
= \overline{\nu}(s)\times \overline{\alpha}(s)$. It will be called
{\it the Darboux frame} of the curve $C$ in $\mathfrak{M}_{t}$.
\begin{theorem}
The fundamental equations of the curve $C$ in the Myller
configuration $\mathfrak{M}_{t}(C,\overline{\alpha},\pi)$ are given by
the following system of differential equations:
\begin{equation}
\displaystyle\frac{d\overline{r}}{ds} = \overline{\alpha}(s),
\end{equation}
\begin{eqnarray}
\displaystyle\frac{d\overline{\alpha}}{ds}&=&
\kappa_{g}(s)\overline{\mu}^{*}(s)+\kappa_{n}(s)\overline{\nu}(s),\nonumber\\
\displaystyle\frac{d\overline{\mu}^{*}}{ds}&=&
-\kappa_{g}(s)\overline{\alpha}(s)+\tau_{g}(s)\overline{\nu}(s),\\\displaystyle\frac{d\overline{\nu}}{ds}&=&
- \kappa_{n}(s)\overline{\alpha}(s) -
\tau_{g}(s)\overline{\mu}^{*}(s).\nonumber
\end{eqnarray}\end{theorem}
Of course $(3.2.1), (3.2.2)$ are the moving equations of the
Darboux frame $\cal{R}_{D}$ of the curve $C$.
The invariants $\kappa_{g}, \kappa_{n}$ and $\tau_{g}$ are called: the
{\it geodesic curvature}, {\it normal curvature} and {\it geodesic torsion} of the
curve $C$ in $\mathfrak{M}_{t}.$
Of course, we can prove a fundamental theorem for the geometry of
curves $C$ in $\mathfrak{M}_{t}:$
\begin{theorem}
A priori given $C^{\infty}$-functions $\kappa_{g}(s),$ $\kappa_{n}(s)$,
$\tau_{g}(s)$, $s\in [a,b]$, there exists a Myller configuration
$\mathfrak{M}_{t}(C,\overline{\alpha},\pi)$ for which $s$ is arclength
on the curve $C$ and given functions are its invariants. Two such
configurations differ by a proper Euclidean motion.\end{theorem}
The proof is the same as proof of Theorem 2.1.2, Chapter 2.
\begin{remark}\rm
Let $C$ be a smooth curve immersed in a $C^{\infty}$ surface $S$
in $E_{3}$. Then the tangent plans $\pi$ to $S$ along $C$ uniquely
determines a tangent Myller configuration
$\mathfrak{M}_{t}(C,\overline{\alpha},\pi)$. Its Darboux frame
$\cal{R}_{D}$ and the invariants $k_{g}, \kappa_{n},\tau_{g}$ of $C$
in $\mathfrak{M}_{t}$ are just the Darboux frame and geodesic
curvature, normal curvature and geodesic torsion of curve $C$ on
the surface $S$.
\end{remark}
Let $\cal{R}_{F} = (P(s); \overline{\alpha}_{1}(s),
\overline{\alpha}_{2}(s), \overline{\alpha}_{3}(s))$, be the Frenet frame of
curve $C$ with $\overline{\alpha}_{1}(s) = \overline{\alpha}(s)$.
The Frenet formulae hold:
\begin{equation}
\displaystyle\frac{d\overline{r}}{ds} = \overline{\alpha}_{1}(s)
\end{equation}
\begin{equation}
\left\{\begin{array}{l}
\displaystyle\frac{d\overline{\alpha}_{1}}{ds} = \kappa(s)\alpha_{2}(s);\vspace{1.5mm}\\
\displaystyle\frac{d\overline{\alpha}_{2}}{ds} = -\kappa(s)\overline{\alpha}_{1}(s) +
\tau(s)\overline{\alpha}_{3}(s)\nonumber\vspace{1.5mm}\\
\displaystyle\frac{d\overline{\alpha}_{3}}{ds} = -\tau(s)\overline{\alpha}_{2}(s)
\end{array}\right.
\end{equation}
where $\kappa(s)$ is the curvature of $C$ and $\tau(s)$ is the torsion of
$C$.
The relations between the invariants $\kappa_{g}, \kappa_{n},\tau_{g}$ and
$\kappa, \tau$ can be obtained like in Section 4, Chapter 2.
\begin{theorem}
The following formulae hold good
\begin{equation}
\kappa_{g}(s) = \kappa \sin \varphi^{*},\;\; \kappa_{n}(s) = \kappa \cos \varphi^{*},\;\;
\tau_{g}(s) = \tau +\displaystyle\frac{d\varphi^{*}}{ds},
\end{equation}
with $\varphi^{*} = \sphericalangle(\overline{\alpha}_{2}(s),
\overline{\nu}(s))$.
\end{theorem}
In the case when we consider the relations between the invariants
$\kappa_{g}, \kappa_{n}, \tau_{g}$ and the invariants $\chi_{1}, \chi_{2}$
of the normal versor field $(C,\overline{\nu})$ we have from the
formulae (3.1.5):
\begin{equation}
\kappa_{n} = \chi_{1}\sin \sigma, \; \tau_{g} = \chi_{1}\cos \sigma,\;\;
\kappa_{g} = \chi_{2} + \displaystyle\frac{d\sigma}{ds},
\end{equation}
where $\sigma = \sphericalangle (\overline{\alpha}(s),
\overline{\nu}_{3}(s))$.
It is clear that the second formula (3.2.5) gives us a theorem of
Meusnier type, and for $\varphi^{*} = 0$ or $\varphi^{*} = \pm \pi$ we have
$\kappa_{g} = 0, \kappa_{n} = \pm \kappa$, $\tau_{g} = \tau.$
For $\sigma = 0,$ or $\sigma = \pm \pi,$ from (3.2.6) we obtain $\kappa_{n} =
0,$ $\tau_{g} = \pm \chi_{1}$, $\kappa_{g} = \chi_{2}$.
\section{Geodesic, asymptotic and curvature\, \, \, lines in $\mathfrak{M}_{t}(C,\overline{\alpha},\pi)$}
\setcounter{theorem}{0}\setcounter{equation}{0}
The notion of parallelism of a versor field $(C,\overline{\alpha})$ in
$\mathfrak{M}_{t}$ along the curve $C$, investigated in the
Section 8, Chapter 2 for the general case, can be applied now for
the particular case of tangent versor field $(C,\overline{\alpha})$.
It is defined by the condition $\kappa_{g}(s) =0,$ $\forall s\in
(s_{1},s_{2})$. The curve $C$ with this property is called
{\it geodesic line} for $\mathfrak{M}_{t}$ (or {\it autoparallel curve}).
The following properties hold:
1. {\it The curve $C$ is a geodesic in the configuration
$\mathfrak{M}_{t}$ iff at every point $P(s)$ of $C$ the osculating
plane of $C$ is normal to $\mathfrak{M}_{t}$.}
2. {\it $C$ is geodesic in $\mathfrak{M}_{t}$ iff the equality
$|\kappa_{n}| = \kappa$ holds along $C$.}
3. {\it If $C$ is a straight line in $\mathfrak{M}_{t}$ then $C$ is a
geodesic of $\mathfrak{M}_{t}$.}
\bigskip
{\bf Asymptotics}
The curve $C$ is called {\it asymptotic} in $\mathfrak{M}_{t}$ if
$\kappa_{n} = 0,$ $\forall s\in (s_{1},s_{2})$. An asymptotic $C$ is
called an {\it asymptotic line}, too. The following properties can be
proved without difficulties:
1. {\it $C$ is asymptotic line in $\mathfrak{M}_{t}$ iff at every point
$P(s)\in C$ the osculating plane of $C$ coincides to the plane
$\pi(s)$.}
2. {\it If $C$ is a straight line then $C$ is asymptotic in
$\mathfrak{M}_{t}$.}
3. {\it $C$ is asymptotic in $\mathfrak{M}_{t}$ iff along $C$,
$|\kappa_{g}| = \kappa.$}
4. {\it If $C$ is asymptotic in $\mathfrak{M}_{t}$, then $\tau_{g} =
\tau$ along $C$.}
5. {\it If $\overline{\alpha}(s)$ is conjugate to $\overline{\alpha}(s)$ then
$C$ is asymptotic line in $\mathfrak{M}_{t}$.}
Therefore we may say that the asymptotic line in
$\mathfrak{M}_{t}$ are the autoconjugated lines.
\medskip
{\bf Curvature lines}
The curve $C$ is called the {\it curvature line} in the
configurations\linebreak $\mathfrak{M}_{t}(C, \overline{\xi},\pi)$
if the ruled surface $\cal{R}(C,\overline{\nu})$ is a developing
surface.
One knows that $\cal{R}(C,\overline{\nu})$ is a developing surface
iff the following equation holds:
$$
\langle\overline{\alpha}(s), \overline{\nu}(s),
\displaystyle\frac{d\overline{\nu}}{ds}(s)\rangle = 0, \;\; \forall s\in
(s_{1},s_{2}).
$$
Taking into account the fundamental equations of
$\mathfrak{M}_{t}$ one gets:
1. {\it $C$ is a curvature line in $\mathfrak{M}_{t}$ iff its geodesic
torsion $\tau(s) = 0$, $\forall s$.}
2. {\it $C$ is a curvature line in $\mathfrak{M}_{t}$ iff the versors
field $(C,\overline{\mu}^{*})$ are conjugated to the tangent
$\overline{\alpha}(s)$ in $\mathfrak{M}_{t}$.}
\begin{theorem}
If the curve $C$ of the tangent Myller con\-fi\-gu\-ration
$\mathfrak{M}_{t}(C,$ $\overline{\xi},\pi)$ satisfies two from the
following three conditions
a) $C$ is a plane curve.
b) $C$ is a curvature line in $\mathfrak{M}_{t}$.
c) The angle $\varphi^{*}(s) = \sphericalangle
(\overline{\alpha}_{2}(s),\overline{\nu}(s))$ is constant,
\noindent then the curve $C$ verifies the third condition,
too.\end{theorem}
For proof, one can apply the Theorem 3.2.3, Chapter 3.
More general, let us consider two tangent Myller configurations
$\mathfrak{M}_{t}(C,$ $\overline{\alpha},\pi)$ and
$\mathfrak{M}_{t}^{*}(C, \overline{\alpha},\pi^{*})$ and $\varphi^{**}(s) =
\sphericalangle (\overline{\nu}(s), \overline{\nu}^{*}(s))$.
We can prove without difficulties:
\begin{theorem}
Assuming satisfied two from the following three conditions
\begin{itemize}
\item[1.] $C$ is curvature line in $\mathfrak{M}_{t}$.
\item[2.] $C$ is curvature line in $\mathfrak{M}_{t}^{*}$.
\item[3.] The angle $\varphi^{**}(s)$ is constant for $s\in
(s_{1},s_{2})$.
\noindent Then, the third condition is also verified.
\end{itemize}
\end{theorem}
\section{Mark Krein$^{\prime}$s formula}
\setcounter{theorem}{0}\setcounter{equation}{0}
In the Section 10, Chapter 2, we have defined the spherical image
of a general Myller configuration $\mathfrak{M}$. Of course the
definition applies for tangent configurations $\mathfrak{M}_{t}$,
too.
But now will appear some new special properties.
If in the formulae from Section 10, Chapter 2 we take
$\overline{\xi}(s) = \overline{\alpha}(s)$, $\forall s$, the
invariants $G,K,T$ reduce to geodesic curvature, normal curvature
and geodesic torsion, respectively, of the curve $C$ in
$\mathfrak{M}_{t}$.
Such that we obtain the formulae:
\begin{equation}
\kappa_{n} = -\displaystyle\frac{ds^{*}}{ds}\cos \theta_{1},\;\; \theta_{1} =
\sphericalangle (\overline{\nu}_{2},\overline{\alpha})
\end{equation}
\begin{equation}
\tau_{g} = -\displaystyle\frac{ds^{*}}{ds}\cos \theta_{2},\;\; \theta_{2} =
\sphericalangle (\overline{\nu}_{2},\overline{\mu}^{*})
\end{equation}
which have as consequences:
1. {\it $C$ is asymptotic in $\mathfrak{M}_{t}$ iff $\theta_{1} = \pm
\displaystyle\frac{\pi}{2}$.}
2. {\it $C$ is curvature line in $\mathfrak{M}_{t}$ iff $\theta_{2} = \pm
\displaystyle\frac{\pi}{2}$.}
\smallskip
The following Mark Krein$'$s theorem holds:
\begin{theorem}
Assume that we have
\begin{itemize}
\item[$1.$] A Myller configuration $\mathfrak{M}_{t}(C,
\overline{\xi},\pi)$ of class $C^{k}$, $(k\geq 3)$ where $s$ is
the natural parameter on the curve $C$ and $C$ is a closed curve.
\item[$2.$] The spherical image $C^{*}$ of $\mathfrak{M}_{t}$ determine on
the support sphere $\Sigma$ a simply connected domain of area
$\omega$.
\item[$3.$] $\sigma = \sphericalangle (\overline{\nu}_{3}, \overline{\alpha})$.
In these conditions Mark Krein$'$s formula holds:
\begin{equation}
\omega =2\pi - \int_{C}\kappa_{g}(s)ds +\int_{C}d\sigma.
\end{equation}
\end{itemize}
\end{theorem}
Indeed the formula (2.10.9) from Theorem 2.10.1, Chapter 2 is
equivalent to (3.4.3).
The remarks from the end of Section 10, Chapter 2 are valid, too.
\section{Relations between the invariants $G,K,T$ of the versor field
$(C,\overline{\xi})$ in $\mathfrak{M}_{t}(C,\overline{\xi},\pi)$ and the
invariants of tangent versor field $(C,\overline{\alpha})$ in
$\mathfrak{M}_{t}$}
\setcounter{theorem}{0}\setcounter{equation}{0}
Let $\mathfrak{M}_{t}(C,\overline{\xi},\pi)$ be a tangent Myller
configuration, $\cal{R}_{D} = (P(s), \overline{\xi}(s),$$
\overline{\mu}(s),$ $\overline{\nu}(s))$ its Darboux frame and the
tangent Myller configuration
$\mathfrak{M}_{t}^{\prime}(C,\overline{a},\pi)$ determined by the
plane field $(C,\pi(s))$, with $\overline{\alpha}
=\displaystyle\frac{d\overline{r}}{ds}$-tangent versors field to the oriented
curve $C$. The Darboux frame $\cal{R}_{D}^{\prime} = (P(s);
\overline{\alpha}(s),$ $ \overline{\mu}^{*}(s),$ $\overline{\nu})$ and
oriented angle $\lambda = \sphericalangle
(\overline{\alpha},\overline{\xi})$ allow to determine $\cal{R}_{D}$
by means of formulas:
\begin{eqnarray}
\overline{\xi}&=& \overline{\alpha}\cos \lambda + \overline{\mu}^{*}\sin
\lambda\nonumber\\\overline{\mu}&=& -\overline{\alpha}\sin \lambda
+\overline{\mu}^{*}\cos \lambda\\\overline{\nu}&=&
\overline{\nu}.\nonumber
\end{eqnarray}
The moving equations (3.1.2), (3.1.3) and (3.2.1), (3.2.3) of
$\cal{R}_{D}$ and $\mathcal{R}^{\prime}_{D}$ lead to the following
relations between invariants $\kappa_{g}, \kappa_{n}$ and $\tau_{g}$ of
the curve $C$ in $\mathfrak{M}_{t}^{\prime}$ and the invariants
$G,K,T$ of versor field $(C,\overline{\xi})$ in
$\mathfrak{M}_{t}:$
\begin{eqnarray}
K&=& \kappa_{n}\cos \lambda + \tau_{g}\sin \lambda\nonumber\\T&=& -\kappa_{n}\sin \lambda
+ \tau_{g}\cos \lambda\\Gamma&=& \kappa_{g} + \displaystyle\frac{d\lambda}{ds}.\nonumber
\end{eqnarray}
The two first formulae (3.5.2) imply
\begin{equation}
K^{2} + T^{2} = \kappa_{n}^{2} + \tau_{g}^{2} =
\left(\displaystyle\frac{ds^{*}}{ds}\right)^{2}.
\end{equation}
The formula (3.5.2) has some important consequences:
\begin{theorem}
1. The curve $C$ has $\kappa_{g}(s)$ as geodesic curvature on the
developing surface $E$ generated by planes $\pi(s)$, $s\in
(s_{1},s_{2})$.
2. $G$ is an intrinsec invariant of the developing surface $E$.
\end{theorem}
\begin{proof} 1. Since the planes $\pi(s)$ pass through tangent line
$(P(s), \overline{\alpha}(s))$ to curve $C$, the developing surface
$E$, enveloping by planes $\pi(s)$ passes through curve $C$. So,
$\kappa_{g}$ is geodesic curvature of $C\subset E$ at point $P(s)$.
2. The invariants $\kappa_{g}(s)$ and $\lambda(s)$ are the intrinsic
invariants of surface $E$. It follows that the invariant $G$ has
the same property.\end{proof}
Now, we investigate an extension of Bonnet result.
\begin{theorem}
Supposing that the versor field $(C, \overline{\xi})$ in tangent
Myller configuration $\mathfrak{M}_{t}(C,\overline{\xi},\pi)$ has
two from the following three properties:
\begin{itemize}
\item[1.] $\overline{\xi}(s)$ is a parallel versor field in
$\mathfrak{M}_{t}$,
\item[2.] The angle $\lambda = \sphericalangle (\overline{\alpha},\overline{\xi})$ is
constant,
\item[3.] The curve $C$ is geodesic in $\mathfrak{M}_{t},$
\end{itemize}
then it has the third property, too.
\end{theorem}
The proof is based on the last formula (3.5.2). Also, it is not
difficult to prove:
\begin{theorem}
The normal curvature and geodesic torsion of the versor field
$(C,\overline{\mu}^{*})$ in tangent Myller configuration
$\mathfrak{M}_{t}(C,\overline{\xi},\pi)$, respectively are the
geodesic torsion and normal curvature with opposite sign of the
curve $C$ in $\mathfrak{M}_{t}$.
\end{theorem}
The same formulae (3.5.2) for $K(s)=0,$ $s\in (s_{1}, s_{2})$ and
$\tau_{g}(s)\neq 0$ imply:
\begin{equation}
{\rm{tg}} \lambda = -\displaystyle\frac{\kappa_{n}(s)}{\tau_{g}(s)}.
\end{equation}
In the theory of surfaces immersed in $E_{3}$ this formula was
independently established by E. Bortolotti [3] and Al. Myller
[31-34].
\begin{definition}
A curve $C$ is called a geodesic helix in tangent configuration
$\mathfrak{M}_{t}(C,\overline{\xi},\pi)$ if the angle $\lambda(s) =
\sphericalangle (\overline{\alpha},\overline{\mu})$ is constant $(\neq
0,\pi)$.
\end{definition}
The following two results can be proved without difficulties.
\begin{theorem}
If $C$ is a geodesic helix in
$\mathfrak{M}_{t}(C,\overline{\xi},\pi)$, then $$\kappa_{n}/\tau_{g} =
\pm \displaystyle\frac{\kappa}{\tau}.$$
\end{theorem}
Another consequence of the previous theory is given by:
\begin{theorem}
If $C$ is geodesic in $\mathfrak{M}_{t}(C,\overline{\xi},\pi)$, it
is a geodesic helix for $\mathfrak{M}_{t}$ iff $C$ is a cylinder
helix in the space $E_{3}$.
\end{theorem}
We can make analogous consideration, taking $\tau=0$ in the formula
(3.5.2).
We stop here the theory of Myller configurations
$\mathfrak{M}(C,\overline{\xi},\pi)$ in Euclidean space $E_{3}$.
Of course, it can be extended for Myller configurations
$\mathfrak{M}(C,\overline{\xi}, \pi)$ with $\pi$ a $k$-plane in
the space $E^{n}, n\geq 3$, $k<n$ or in more general spaces, as
Riemann spaces, [26], [27].
\newpage
\thispagestyle{empty}
\chapter[Applications of the theory of Myller configuration $\mathfrak{M}_{t}$]{Applications of theory of Myller configuration $\mathfrak{M}_{t}$ in the geometry of surfaces in
$E_{3}$}
A first application of the theory of Myller configurations
$\mathfrak{M}(C,\overline{\xi},\pi)$ in the Euclidean space $E_{3}$
can be realized to the geometry of surfaces $S$ embedded in
$E_{3}$. We obtain a more clear study of curves $C\subset S$, a
natural definition of Levi-Civita parallelism of vector field
$\overline{V}$, tangent to $S$ along $C$, as well as the notion of
concurrence in Myller sense of vector field, tangent to $S$ along
$C$. Some new concepts, as those of mean torsion of $S$, the total
torsion of $S$, the Bonnet indicatrix of geodesic torsion and its
relation with the Dupin indicatrix of normal curvatures are
introduced. A new property of Bacaloglu-Sophie-Germain curvature
is proved, too. Namely, it is expressed in terms of the total
torsion of the surface $S$.
\section{The fundamental forms of surfaces in $E_{3}$}
\setcounter{theorem}{0}\setcounter{equation}{0}
Let $S$ be a smooth surface embedded in the Euclidean space
$E_{3}$. Since we use the classical theory of surfaces in $E_{3}$
[see, for instance: Mike Spivak, Diff. Geom. vol. I, Bearkley,
1979], consider the analytical representation of $S$, of class $C^k$, $(k\geq 3,\ {\rm{or\ }}k=\infty)$:
\begin{equation}
\overline{r} = \overline{r}(u,v),\;\; (u,v)\in D.
\end{equation}
$D$ being a simply connected domain in plane of variables $(u,v)$,
and the following condition being verified:
\begin{equation}
\overline{r}_{u}\times \overline{r}_{v}\neq \overline{0}\;\;\RIfM@\expandafter\text@\else\expandafter\mbox\fi{
for }\forall (u,v)\in D.
\end{equation}
Of course we adopt the vectorial notations
\begin{eqnarray*}
\overline{r}_{u} = \displaystyle\frac{\partial \overline{r}}{\partial u},\;\;
\overline{r}_{v} = \displaystyle\frac{\partial\overline{r}}{\partial v}.
\end{eqnarray*}
Denote by $C$ a smooth curve on $S$, given by the parametric
equations
\begin{equation}
u = u(t),\;\; v = v(t),\;\; t\in (t_{1},t_{2}).
\end{equation}
In this chapter all geometric objects or mappings are considered
of $C^{\infty}$-class, up to contrary hypothesis.
In space $E_{3}$, the curve $C$ is represented by
\begin{equation}
\overline{r} = \overline{r}(u(t), v(t)),\;\; t\in (t_{1},t_{2}).
\end{equation}
Thus, the following vector field
\begin{equation}
d\overline{r} = \overline{r}_{u}du + \overline{r}_{v}dv
\end{equation}
is tangent to $C$ at the points $P(t) = P(\overline{r}(u(t)),
v(t))\in C.$ The vectors $\overline{r}_{u}$ and $\overline{r}_{v}$
are tangent to the parametric lines and $d\overline{r}$ is tangent
vector to $S$ at point $P(t).$
From (4.1.2) it follows that the scalar function
\begin{equation}
\Delta = \|\overline{r}_{u}\times \overline{r}_{v}\|^{2}
\end{equation}
is different from zero on $D$.
The unit vector $\overline{\nu}$
\begin{equation}
\overline{\nu} = \displaystyle\frac{\overline{r}_{u}\times
\overline{r}_{v}}{\sqrt{\Delta}}
\end{equation}
is normal to surface $S$ at every point $P(t).$
The tangent plane $\pi$ at $P$ to $S$, has the equation
\begin{equation}
\langle \overline{R}-\overline{r}(u,v), \overline{\nu}\rangle=0.
\end{equation}
Assuming that $S$ is orientable, it follows that $\overline{\nu}$
is uniquely determined and the tangent plane $\pi$ is oriented (by
means of versor $\overline{\nu}$), too.
The {\it first fundamental form} $\phi$ of surface $S$ is defined by
\begin{equation}
\phi = \langle d\overline{r}, d\overline{r}\rangle,\;\; \forall
(u,v)\in D
\end{equation}
or by $\phi(du,dv) = \langle d\overline{r}(u,v),
d\overline{r}(u,v)\rangle$, $\forall (u,v)\in D.$
Taking into account the equality (4.1.5) it follows that
$\phi(du,dv)$ is a quadratic form:
\begin{equation}
\phi(du,dv) = E du^{2} + 2F du dv + G dv^{2}.
\end{equation}
The coefficients of $\phi$ are the functions, defined on $D$:
\begin{equation}
\hspace*{1cm}E(u,v) = \langle \overline{r}_{u}, \overline{r}_{u}\rangle, F(u,v)
= \langle \overline{r}_{u}, \overline{r}_{v}\rangle,\; G(u,v) =
\langle \overline{r}_{v}, \overline{r}_{v} \rangle.
\end{equation}
But we have the discriminant $\Delta$ of $\phi:$
\begin{equation}
\Delta = EG - F^{2}.
\end{equation}
\begin{equation}
E>0,\; G>0,\; \Delta >0.
\end{equation}
Consequence: the first fundamental form $\phi$ of $S$ is
positively defined.
Since the vector $d\overline{r}$ does not depend on
parametrization of $S$, it follows that $\phi$ has a geometrical meaning
with respect to a change of coordinates in $E_{3}$ and with
respect to a change of parameters $(u,v)$ on $S$.
Thus $ds,$ given by:
\begin{equation}
ds^{2} = \phi(du,dv) = E du^{2} + 2F du dv + G dv^{2}
\end{equation}
is called the {\it element of arclength} of the surface $S$.
The arclength of a curve $C$ an $S$ is expressed by
\begin{equation}
s = \int_{t_{0}}^{t}\left\{E\(\displaystyle\frac{du}{d\sigma}\)^{2} + 2F
\displaystyle\frac{du}{d\sigma}\displaystyle\frac{dv}{d\sigma} + G\(\displaystyle\frac{dv}{d\sigma}\)^{2}\right\}^{1/2}d\sigma.
\end{equation}
The function $s = s(t),$ $t\in [a,b]\subset (t_{1},t_{2})$,
$t_{0}\in [a,b]$ is a diffeomorphism from the interval $[a,b]\to
[0,s(t)]$. The inverse function $t =t(s)$, determines a new
parametrization of curve $C$: $\overline{r} = \overline{r}(u(s),
v(s))$. The tangent vector $\displaystyle\frac{d\overline{r}}{ds}$ is a versor:
\begin{equation}
\overline{\alpha} = \displaystyle\frac{d\overline{r}}{ds} =
\overline{r}_{u}\displaystyle\frac{du}{ds} + \overline{r}_{v}\displaystyle\frac{dv}{ds}.
\end{equation}
For two tangent versors $\overline{\alpha} =
\displaystyle\frac{d\overline{r}}{ds}$, $\overline{\alpha}_{1} =
\displaystyle\frac{\delta\overline{r}}{\delta s}$ at point $P\in S$ the angle
$\sphericalangle(\overline{\alpha}, \overline{\alpha}_{1})$ is expressed
by
\begin{equation}
\hspace*{13mm}\cos\hspace{-1mm} \sphericalangle(\hspace{-0.5mm}\overline{\alpha}, \overline{\alpha}_{1}\hspace{-0.5mm})\hspace{-1mm} =\hspace{-1mm}
\displaystyle\frac{E du
\delta u\hspace{-0.8mm} +\hspace{-0.8mm} F(du \delta v\hspace{-0.8mm} +\hspace{-0.8mm} \delta u dv)\hspace{-0.8mm} +\hspace{-0.8mm} Gdv \delta v}{\sqrt{Edu^{2}\hspace{-0.8mm} +\hspace{-0.8mm} 2F du dv\hspace{-0.8mm} +\hspace{-0.8mm}
Gdv^{2}}\hspace{-1mm}\cdot\hspace{-1mm} \sqrt{E\delta u^{2}\hspace{-0.8mm} +\hspace{-0.8mm} 2F \delta u \delta v\hspace{-0.8mm} +\hspace{-0.8mm} G\delta v^{2}}}.
\end{equation}
{\it The second fundamental form} $\psi$ of $S$ at point $P\in S$ is
defined by
\begin{equation}
\psi = \langle \overline{\nu}, d^{2}\overline{r}\rangle = -
\langle d\overline{r}, d\overline{\nu}\rangle\;\; \forall (u,v)\in
D.
\end{equation}
Also, we adopt the notation $\psi(du,dv) = - \langle
d\overline{r}(u,v),d\nu(u,v)\rangle.$ $\psi$ is a quadratic form:
\begin{equation}
\psi(du,dv) =L du^{2} + 2M du dv + N dv^{2},
\end{equation}
having the coefficients functions of $(u,v)\in D:$
\begin{equation}
\begin{array}{l}
L(u,v) = \displaystyle\frac{1}{\sqrt{\Delta}} \langle \overline{r}_{u},
\overline{r}_{u}, \overline{r}_{uu}\rangle,\; M(u,v) =
\displaystyle\frac{1}{\sqrt{\Delta}}\langle \overline{r}_{u}, \overline{r}_{v},
\overline{r}_{uv} \rangle\\N(u,v) =
\displaystyle\frac{1}{\sqrt{\Delta}}<\overline{r}_{u}, \overline{r}_{v},
\overline{r}_{vv}>.
\end{array}
\end{equation}
Of course from $\psi =-\langle d\overline{r},
d\overline{\nu}\rangle$ it follows that $\psi$ has geometric
meaning.
The two fundamental forms $\phi$ and $\psi$ are enough to
determine a surface $S$ in Euclidean space $E_{3}$ under proper
Euclidean motions. This property results by the integration of the
Gauss-Weingarten formulae and by their differential consequences
given by Gauss-Codazzi equations.
\section{Gauss and Weingarten formulae}
\setcounter{theorem}{0}\setcounter{equation}{0}
Consider the moving frame$$ \cal{R} = (P(\overline{r});
\overline{r}_{u}, \overline{r}_{v}, \overline{\nu})
$$
on the smooth surface $S$ (i.e. $S$ is of class $C^\infty$).
The moving equations of $R$ are given by the Gauss and Weingarten
formulae.
The Gauss formulae are as follows:
\begin{equation}
\begin{array}{lll}
\overline{r}_{uu}&=& \left\{\begin{array}{l}1\\11\end{array}\right\} \overline{r}_{u} + \left\{\begin{array}{l}2\\ 11\end{array}\right\}
\overline{r}_{v} +
L\overline{\nu},\\\overline{r}_{uv}&=& \left\{\begin{array}{l}1\\
12\end{array}\right\}\overline{r}_{u} + \left\{\begin{array}{l}2\\ 12\end{array}\right\} \overline{r}_{v} +
M\overline{\nu}\\\overline{r}_{vv}&=&\left\{\begin{array}{l}1\\
22\end{array}\right\}\overline{r}_{u} +\left\{\begin{array}{l}2\\ 22\end{array}\right\} \overline{r}_{v} +
N\overline{\nu}.
\end{array}
\end{equation}
The coefficients $\left\{\begin{array}{cc} i\\jk
\end{array}\right\}$, $(i,j,k=1,2; u=u^{1}, v= u^{2})$ are called the
Christoffel symbols of the first fundamental form $\phi:$
\begin{equation}
\hspace*{5mm}{1\brace 11} =-\displaystyle\frac{1}{\sqrt{D}} \langle \overline{\nu},
\overline{r}_{v}, \overline{r}_{uu}\rangle, { 1\brace 21} ={
1\brace 21} = -\displaystyle\frac{1}{\sqrt{\Delta}}\langle \overline{\nu},
\overline{r}_{v}, \overline{r}_{uv}\rangle,
\end{equation}
$$
{ 1\brace 22} = -\displaystyle\frac{1}{\sqrt{\Delta}}\langle \overline{\nu},
\overline{r}_{v}, \overline{r}_{vv} \rangle,
$$
\begin{eqnarray}
\hspace*{14mm}{ 2\brace 11}
= - \displaystyle\frac{1}{\sqrt{\Delta}}\langle \overline{\nu}, \overline{r}_{u},
\overline{r}_{uu}\rangle, {2\brace 21}
= {2\brace 12}
= -\displaystyle\frac{1}{\sqrt{\Delta}}\langle \overline{\nu},\overline{r}_{u},
\overline{r}_{uv}\rangle,\nonumber\\ {2\brace 22} =
-\displaystyle\frac{1}{\sqrt{\Delta}}\langle
\overline{\nu},\overline{r}_{u},\overline{r}_{vv}\rangle.\nonumber
\end{eqnarray}
The Weingarten formulae are given by:
\begin{eqnarray}
\displaystyle\frac{\partial \overline{\nu}}{\partial u}&=& \displaystyle\frac{1}{\sqrt{\Delta}}\{(FM -
GL)\overline{r}_{u} + (FL - EM)\overline{r}_{v}\}\\\displaystyle\frac{\partial
\overline{\nu}}{\partial v}&=& \displaystyle\frac{1}{\sqrt{\Delta}}\{(FN -
GM)\overline{r}_{u} + (FM -EN)\overline{r}_{v}\}.\nonumber
\end{eqnarray}
Of course, the equations (4.2.1) and (4.2.3) express the variation
of the moving frame $R$ on surface $S$.
Using the relations $\overline{r}_{uuv} = \overline{r}_{uvu},$
$\overline{r}_{vuv} = \overline{r}_{vvu}$ and
$\displaystyle\frac{\partial^{2}\overline{\nu}}{\partial uv} =
\displaystyle\frac{\partial^{2}\overline{\nu}}{\partial v u}$ applied to (4.2.1), (4.2.3)
one deduces the so called fundamental equations of surface $S$,
known as the Gauss-Codazzi equations.
A fundamental theorem can be proved, when the first and second
fundamental from $\phi$ and $\psi$ are given and Gauss-Codazzi
equations are verified (Spivak [75]).
\newpage
\section{The tangent Myller configuration $\mathfrak{M}_{t}(C,$ $\overline{\xi},\pi)$ associated to a tangent versor field $(C,\overline{\xi})$ on a surface
$S$}
\setcounter{theorem}{0}\setcounter{equation}{0}
Assuming that $C$ is a curve on surface $S$ in $E_{3}$ and $(C,
\overline{\xi})$ is a versor field tangent to $S$ along $C$, there
is an uniquely determined tangent Myller configuration
$\mathfrak{M}_{t} = \mathfrak{M}_{t}(C, \overline{\xi},\pi)$ for
which $(C,\pi)$ is tangent field planes to $S$ along $C$.
The invariants $(c_{1}, c_{2}, G,K,T)$ of $(C, \overline{\xi})$ in
$\mathfrak{M}_{t}(C, \overline{\xi},\pi)$ will be called the
invariants of tangent versor field $(C, \overline{\xi})$ on
surface $S$.
Evidently, these invariants have the same values on every smooth
surface $S^{\prime}$ which contains the curve $C$ and is tangent
to $S$.
Using the theory of $\mathfrak{M}_{t}$ from Chapter 3, for $(C,
\overline{\xi})$ tangent to $S$ along $C$ we determine:
1. The Darboux frame
\begin{equation}
R_{D} = (P(s); \overline{\xi}(s), \overline{\mu}(s),
\overline{\nu}(s)),
\end{equation}
where $s$ is natural parameter on $C$ and
\begin{equation}
\overline{\nu} = \displaystyle\frac{1}{\sqrt{\Delta}} \overline{r}_{u}\times
\overline{r}_{v},\;\; \overline{\mu} = \overline{\nu}\times
\overline{\xi}.
\end{equation}
Of course $\cal{R}_{D}$ is orthonormal and positively oriented.
The invariants $(c_{1}, c_{2}, G,K,M)$ of $(C, \overline{\xi})$ on
$S$ are given by (3.1.2), (3.1.3) Chapter 3. So we have the
moving equations of $R.$
\begin{theorem}
The moving equations of the Darboux frame of tangent versor field
$(C,\overline{\xi})$ to $S$ are given by
\begin{equation}
\displaystyle\frac{d\overline{r}}{ds} = \overline{\alpha}(s) =
c_{1}(s)\overline{\xi} + c_{2}(s)\overline{\mu},\;\; (c_{1}^{2} +
c_{2}^{2} =1)
\end{equation}
and
\newpage
\begin{eqnarray}
\displaystyle\frac{d\overline{\xi}}{ds} &=& G(s)\overline{\mu}(s) +
K(s)\overline{\nu}(s),\nonumber\\ \displaystyle\frac{d\overline{\mu}}{ds} &=&
-G(s)\overline{\xi}(s) +
T(s)\overline{\nu}(s),\\\displaystyle\frac{d\overline{\nu}}{ds} &=&
-K(s)\overline{\xi} - T(s)\overline{\mu}(s),\nonumber
\end{eqnarray}
where $c_{1}(s), c_{2}(s), G(s), K(s), T(s)$ are invariants with
respect to the changes of coordinates on $E_{3}$, with respect of transformation of local coordinates on $S$, $(u,v)\to(\widetilde{u},\widetilde{v})$ and with respect
to transformations of natural parameter $s\to s_{0}+s$.
\end{theorem}
A fundamental theorem can be proved exactly as in Chapter 2.
\begin{theorem}
Let be $c_{1}(s),c_{2}(s)$, $[c_{1}^{2} + c_{2}^{2}=1]$, $G(s),
K(s), T(s)$, $s\in [a,b]$, a priori given smooth functions. Then
there exists a tangent Myller configuration
$\mathfrak{M}_{t}(C,\overline{\xi},\pi)$ for which $s$ is the
arclength of curve $C$ and the given functions are its invariants.
$\mathfrak{M}_{t}$ is determined up to a proper Euclidean motion.
The given functions are invariants of the versor field
$(C,\overline{\xi})$ for any smooth surface $S$, which contains
the curve $C$ and is tangent planes $\pi(s)$, $s\in [a,b]$.
\end{theorem}
The geometric interpretations of the invariants $G(s), K(s)$ and
$T(s)$ are those mentioned in the Section 2, Chapter 3.
So, we get
$$
G(s) = \lim_{\Delta s\to 0}\displaystyle\frac{\Delta\psi_{1}}{\Delta s}, \;\; \RIfM@\expandafter\text@\else\expandafter\mbox\fi{
with } \Delta \psi_{1} = \sphericalangle (\overline{\xi}(s),
{\rm{pr}}_{\pi(s)}\overline{\xi}(s+\Delta s)),
$$
$$
K(s) = \lim_{\Delta s\to 0}\displaystyle\frac{\Delta\psi_{2}}{\Delta s}, \;\; \RIfM@\expandafter\text@\else\expandafter\mbox\fi{
with } \Delta \psi_{2} = \sphericalangle (\overline{\mu},
{\rm{pr}}_{(P; \overline{\xi},\overline{\nu})}\overline{\xi}(s+\Delta
s)),
$$
$$
T(s) = \lim_{\Delta s\to 0}\displaystyle\frac{\Delta\psi_{3}}{\Delta s}, \;\; \RIfM@\expandafter\text@\else\expandafter\mbox\fi{
with } \Delta \psi_{3} = \sphericalangle (\overline{\mu},
{\rm{pr}}_{(P; \overline{\mu},\overline{\nu})}\overline{\xi}(s+\Delta
s)).
$$
For this reason $G(s)$ is called the {\it geodesic curvature} of $(C,
\overline{\xi})$ on $S$; $K(s)$ is called {\it the normal curvature} of
$(C, \overline{\xi})$ on $S$ and $T(s)$ is named the {\it geodesic
torsion} of the tangent versor field $(C, \overline{\xi})$ on $S$.
The calculus of invariants $G,K,T$ is exactly the same as it has been done in Chapter
2, (2.3.3), (2.3.4):
\begin{eqnarray}
&&G(s) = \left\langle \overline{\xi}, \displaystyle\frac{d\overline{\xi}}{ds},
\overline{\nu} \right\rangle,\;\; K(s) =
\left\langle\displaystyle\frac{d\overline{\xi}}{ds},\overline{\nu}\right\rangle = -\left\langle
\overline{\xi},\displaystyle\frac{d\overline{\nu}}{ds}
\right\rangle,\nonumber\\&&T(s) = \left\langle \overline{\xi},
\overline{\nu}, \displaystyle\frac{d\overline{\nu}}{ds}\right\rangle.
\end{eqnarray}
The analytical expressions of invariants $G,K,T$ on $S$ are given in next
section.
\section{The calculus of invariants $G,K,T$ on a surface}
\setcounter{theorem}{0}\setcounter{equation}{0}
The tangent versor $\overline{\alpha} = \displaystyle\frac{d\overline{r}}{ds}$ to
the curve $C$ in $S$ is
\begin{equation}
\overline{\alpha}(s) = \displaystyle\frac{d\overline{r}}{ds} = \overline{r}_{u}
\displaystyle\frac{du}{ds} + \overline{r}_{v}\displaystyle\frac{dv}{ds}.
\end{equation}
$\overline{\alpha}(s)$ is the versor of the tangent vector
$d\overline{r}$, from (4,4)
\begin{eqnarray}
d\overline{r} = \overline{r}_{u}du + \overline{r}_{v}dv.
\end{eqnarray}
The coordinate of a vector field $(C, \overline{V})$ tangent to
surface $S$ along the curve $C\subset S$, with respect to the
moving frame $\cal{R} = (P; \overline{r}_{u}, \overline{r}_{v},
\overline{\nu})$ are the $C^{\infty}$ functions $V^{1}(s),
V^{2}(s)$:
$$
\overline{V}(s)=V^{1}(s)\overline{r}_{u}+V^{2}(s)\overline{r}_{v}.
$$
The square of length of vector $\overline{V}(s)$ is:
$$
\langle \overline{V}(s),\overline{V}(s)\rangle = E(V^{1})^{2} +
2FV^{1}V^{2} + G(V^{2})^{2}
$$
and the scalar product of two tangent vectors $\overline{V}(s)$
and
$\overline{U}(s)=U^{1}(s)\overline{r}_{u}+U^{2}(s)\overline{r}_{v}$
is as follows
$$
\langle \overline{U}, \overline{V}\rangle = EU^{1}V^{1} +
F(U^{1}V^{2} + V^{1}U^{2}) +GU^{2}V^{2}.
$$
Let $C$ and $C^{\prime}$ two smooth curves on $S$ having $P(s)$ as
common point. Thus the tangent vectors $d\overline{r},
\delta\overline{r}$ at a point $P$ to $C$, respectively to $C^{\prime}$
are
$$
d\overline{r} = \overline{r}_{u}du + \overline{r}_{v}dv,
$$
$$
\delta \overline{r} = \overline{r}_{u} \delta u +\overline{r}_{v}\delta v.
$$
They correspond to tangent directions $(du,dv)$, $(\delta u, \delta v)$ on
surface $S$.
Evidently, $\overline{\alpha}(s) = \displaystyle\frac{d\overline{r}}{ds}$ and
$\overline{\xi}(s) = \displaystyle\frac{\delta \overline{r}}{\delta s}$ are the tangent
versors to curves $C$ and $C^{\prime}$, respectively at point
$P(s)$, and $(C, \overline{\alpha})$, $(C, \overline{\xi})$ are the
tangent versor fields along the curve $C$.
The Darboux frame $\cal{R}_{D} = (P(s); \overline{\xi}(s),
\overline{\mu}(s), \overline{\nu}(s))$ has the versors
$\overline{\xi}(s), \overline{\mu}(s),$ $\overline{\nu}(s)$ given
by
\begin{eqnarray}
\overline{\xi}(s) &=& \overline{r}_{u}\displaystyle\frac{\delta u}{\delta s} +
\overline{r}_{v}\displaystyle\frac{\delta v}{\delta s},\nonumber\\\overline{\mu}(s)&=&
\displaystyle\frac{1}{\sqrt{\Delta}}\left[(E \overline{r}_{v} - F
\overline{r}_{u})\displaystyle\frac{\delta u}{\delta s} +(F \overline{r}_{v} - G
\overline{r}_{u})\displaystyle\frac{\delta v}{\delta s}\right],\\\overline{\nu}(s) &=&
\displaystyle\frac{1}{\Delta}(\overline{r}_{u} \times \overline{r}_{v}).\nonumber
\end{eqnarray}
Taking into account (4.3.3) and (4.4.3) it follows:
\begin{eqnarray}
c_{1}&=& \langle \overline{\alpha}, \overline{\xi}\rangle = \displaystyle\frac{E du
\delta u +F (du \delta v + dv \delta u)+G dv \delta v}{\sqrt{\phi(du,
dv)}\sqrt{\phi(\delta u, \delta v)}}\\c_{2}&=& \langle
\overline{\alpha},\overline{\mu}\rangle=\displaystyle\frac{\sqrt{\Delta}(\delta u dv - \delta v
du)}{\sqrt{\phi(du,dv)}\sqrt{\phi(\delta u, \delta v)}}\nonumber
\end{eqnarray}
where
$$
\phi(du, dv) = \langle d\overline{r}, d\overline{r}\rangle,\;\;
\phi(\delta u, \delta v) = \langle \delta \overline{r}, \delta
\overline{r}\rangle.
$$
In order to calculate the invariants $G,K,T$ we need to determine
the vectors $\displaystyle\frac{d\overline{\xi}}{ds},
\displaystyle\frac{d\overline{\nu}}{ds}$.
By means of (4.4.3) we obtain
\begin{eqnarray}
\displaystyle\frac{d\overline{\xi}}{ds}&=& \displaystyle\frac{d}{ds}\left(\displaystyle\frac{\delta
\overline{r} }{\delta s}\right) =
\overline{r}_{uu}\displaystyle\frac{du}{ds}\displaystyle\frac{\delta u}{\delta s} +
\overline{r}_{uv}\left(\displaystyle\frac{du}{ds}\displaystyle\frac{\delta v}{\delta s} + \displaystyle\frac{\delta u
}{\delta s}\displaystyle\frac{dv}{ds}\right)
\nonumber\\&&+\overline{r}_{vv}\displaystyle\frac{dv}{ds} \displaystyle\frac{\delta v}{\delta s} +
\overline{r}_{u}\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\delta u }{\delta s}\right)
+\overline{r}_{v}\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\delta\nu}{\delta s}\right).
\end{eqnarray}
The scalar mixt $\langle\overline{r}_{u}, \overline{r}_{v},
\overline{\nu})\rangle$ is equal to $\sqrt{\Delta}$. It results:
\begin{eqnarray}
\left\langle \overline{\xi}, \displaystyle\frac{d\overline{\xi}}{ds},
\overline{\nu}\right\rangle& =&
\sqrt{\Delta}\left[\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\delta v }{\delta s}\right)
\displaystyle\frac{du}{ds} - \displaystyle\frac{d}{ds}\left(\displaystyle\frac{\delta u }{\delta s
}\right)\displaystyle\frac{dv}{ds}\right] \nonumber\\&&+ \langle
\overline{r}_{u},
\overline{r}_{uu},\overline{\nu}\rangle\displaystyle\frac{du}{ds}\left(\displaystyle\frac{\delta
u}{\delta s}\right)^{2}\nonumber+\\&&+ \langle \overline{r}_{u},
\overline{r}_{uv}, \overline{\nu}
\rangle\left(\displaystyle\frac{du}{ds}\displaystyle\frac{\delta v}{\delta s} +
\displaystyle\frac{dv}{ds}\displaystyle\frac{\delta u}{\delta s} \right)\displaystyle\frac{\delta u}{\delta s} + \\&&+
\langle \overline{r}_{u}, \overline{r}_{vv},
\overline{\nu}\rangle\displaystyle\frac{\delta u}{\delta s}\displaystyle\frac{dv}{ds}\displaystyle\frac{\delta v}{\delta
s }+\nonumber\\&&+\langle \overline{r}_{v}, \overline{r}_{uu},
\overline{\nu}\rangle \displaystyle\frac{du}{ds}\displaystyle\frac{\delta u}{\delta s}\displaystyle\frac{\delta v}{\delta
s } + \langle \overline{r}_{v}, \overline{r}_{uv},
\overline{\nu}\rangle\left(\displaystyle\frac{du}{ds}\displaystyle\frac{\delta v}{\delta s} +
\displaystyle\frac{dv}{ds} \displaystyle\frac{\delta u}{\delta s}\right)\displaystyle\frac{\delta v}{\delta
s}\nonumber\\&&+\langle \overline{r}_{v}, \overline{r}_{vv},
\overline{\nu}\rangle\displaystyle\frac{dv}{ds}\left(\displaystyle\frac{\delta v}{\delta s
}\right)^{2}.\nonumber
\end{eqnarray}
Taking into account (4.2.1), (4.2.2), (4.3.5) and (4.4.5) we have
\begin{proposition}
The geodesic curvature of the tangent versor field
$(C,\overline{\xi})$ on $S$, is expressed as follows:
\begin{eqnarray}
G(\delta,d)&=& \sqrt{\Delta}\{\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\delta v}{\delta
s}\right)\displaystyle\frac{\delta u}{\delta s} - \displaystyle\frac{d}{ds}\left(\displaystyle\frac{\delta u}{\delta s
}\right)\displaystyle\frac{\delta v}{\delta s} \nonumber\\&&+ \left( {2\brace 11}
\displaystyle\frac{du}{ds}+{2\brace12} \displaystyle\frac{dv}{ds}\right)\left(\displaystyle\frac{\delta u}{\delta
s}\right)^{2}\\&&\left(\left({ 1\brace 11}
-{2\brace 12}
\right)\displaystyle\frac{du}{ds} - \left( {2\brace 22}
- {1\brace 12}
\right)\displaystyle\frac{dv}{ds}\right)\displaystyle\frac{\delta u}{\delta s}\displaystyle\frac{\delta v}{\delta s}
\nonumber\\&& - \left({ 1\brace 21} \displaystyle\frac{du}{ds} + {1\brace 22}
\displaystyle\frac{dv}{ds}\right)\left(\displaystyle\frac{\delta v}{\delta s}\right)^{2}\}.\nonumber
\end{eqnarray}
\end{proposition}
Remark that the Christoffel symbols are expressed only by means of
the coefficients of the first fundamental form $\phi$ of surface
$S$ and their derivatives. It follows a very important result
obtained by Al Myller [34]:
\begin{theorem}
The geodesic curvature $G$ of a tangent versor field $(C,
\overline{\xi})$ on a surface $S$ is an intrisic invariant of $S$.
\end{theorem}
The invariant $G$ was named by Al. Myller the {\it deviation of
para\-llelism} of the tangent field $(C,\overline{\xi})$ on
surface $S$.
The expression (4.4.7) of $G$ can be simplified by introducing the
following notations
\begin{eqnarray}
&&\displaystyle\frac{D}{ds}\left(\displaystyle\frac{\delta u^{i}}{\delta s}\right) =
\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\delta u^{i}}{\delta s}\right) +\sum_{j,k=1}^{2}{
i\brace jk}
\displaystyle\frac{\delta u^{j}}{\delta s} \displaystyle\frac{d u^{k}}{ds}\nonumber\\&& (u= u^{1}; v =u^{2};
i=1,2).
\end{eqnarray}
The operator (4.4.8) is the classical operator of covariant
derivative with respect to Levi-Civita connection.
Denoting by
\begin{equation}
\hspace*{10mm}G^{ij}(\delta, d) = \sqrt{\Delta}\left\{\displaystyle\frac{D}{ds}\left(\displaystyle\frac{\delta
u^{i}}{\delta s }\right)\displaystyle\frac{\delta u^{j}}{\delta s} -
\displaystyle\frac{D}{ds}\left(\displaystyle\frac{\delta u^{j} }{\delta s}\right)\displaystyle\frac{\delta u^{i}}{\delta
s}\right\},(i,j=1,2)
\end{equation}
and remarking that $G^{ij}(\delta,d) = -G^{ji}(\delta,d)$ we have $G^{11}
= G^{22}=0$.
\begin{proposition}
The following formula holds:
\begin{equation}
G(\delta, d) = G^{12}(\delta,d).
\end{equation}
\end{proposition}
Indeed, the previous formula is a consequence of (4.4.7), (4.4.8)
and (4.4.9).
\begin{corollary}
The parallelism of tangent versor field $(C,\overline{\xi})$ to
$S$ along the curve $C$ is characterized by the differential
equation
\begin{equation}
G^{ij}(\delta,d) = 0\;\; (i,j =1,2).
\end{equation}
\end{corollary}
In the following we introduce the notations \begin{equation} G(s)
= G(\delta,d),\;\; K(s) = K(\delta,d),\; T(s) = T(\delta,d),
\end{equation}
since $\overline{\xi}(s) = \overline{r}_{u} \displaystyle\frac{\delta u}{\delta s} +
\overline{r}_{v}\displaystyle\frac{\delta v}{\delta s}$, $\overline{\alpha}(s) =
\overline{r}_{u}\displaystyle\frac{du}{ds}+\overline{r}_{v} \displaystyle\frac{dv}{ds}$.
\medskip
The second formula (4.3.5), by means of (4.4.5) gives us the
expression of normal curvature $K(\delta,d)$ in the form
\begin{equation}
\hspace*{12mm}K(\delta,d) = \displaystyle\frac{L du \delta u +M(du \delta v + dv \delta u) + N dv \delta
v}{\sqrt{Edu^{2} + 2F du dv + G dv^{2}}\sqrt{E\delta u^{2} + 2F\delta u \delta
v + G\delta v^{2}}}.
\end{equation}
It follows
\begin{equation}
K(\delta,d) = K(d,\delta)
\end{equation}
\begin{remark}\rm
The invariant $K(\delta,d)$ can be written as follows
$$
K(\delta,d) = \displaystyle\frac{\psi(\delta,d)}{\sqrt{\phi(d,d)}\sqrt{\phi(\delta,\delta)}}
$$
where $\psi(\delta,d)$ is the polar form of the second fundamental
form $\psi(d,d)$ of surface $S$.
\end{remark}
$K(\delta,d)=0$ gives us the property of conjugation of $(C,
\overline{\xi})$ with $(C,\overline{\alpha})$.
The calculus of invariant $T(\delta,d)$ can be made by means of
Weingarten formulae (4.2.3).
Since we have
\begin{eqnarray}
\displaystyle\frac{d\overline{\nu}}{ds} &=& \displaystyle\frac{1}{\Delta}\left\{[(FM -GL)
\overline{r}_{u} + (FL -
EM)\overline{r}_{v}]\displaystyle\frac{du}{ds}\nonumber\right.\\&&\left.+ \left[(FN
-GM)\overline{r}_{u} + (FM -
EN)\overline{r}_{v}\right]\displaystyle\frac{dv}{ds}\right\},
\end{eqnarray}
we deduce
\begin{eqnarray}
T(\delta,d)&=& \displaystyle\frac{1}{\sqrt{\Delta}}\left\{\left[(FM - GL) \displaystyle\frac{du}{ds} +
(FN-GM)\displaystyle\frac{dv}{ds}\right]\displaystyle\frac{\delta v}{\delta s}\right.\nonumber\\&&\left.-
\left[(FL-EM)\displaystyle\frac{du}{ds} + (FM- EN)\displaystyle\frac{dv}{ds}\right]\displaystyle\frac{\delta
u}{\delta s}\right\}.
\end{eqnarray}
For simplicity we write $T(\delta,d)$ from previous formula in the
following form
\begin{equation}
\hspace*{12mm}T(\delta,d) =
\displaystyle\frac{1}{\sqrt{\Delta}\sqrt{\phi(d,d)}\sqrt{\phi(\delta,\delta)}}\left\|
\begin{array}{ccc}
E \delta u\hspace{-0.5mm} +\hspace{-0.5mm} F\delta v& F\delta u\hspace{-0.5mm} +\hspace{-0.5mm}G \delta v\\Lambda du\hspace{-0.5mm} +\hspace{-0.5mm} Mdv & Mdu + N dv
\end{array} \right\|.
\end{equation}
Remember the expression of mean curvature $H$ and total curvature
$K_{t}$ of surface $S$:
\begin{equation}
H = \displaystyle\frac{EN - 2FM +GN}{2(EG -F^{2})},\;\; K_{t} = \displaystyle\frac{LN -
M^{2}}{EG-F^{2}}.
\end{equation}
It is not difficult to prove, by means of (4.4.16), the following
formula
\begin{equation}
T(\delta, d) - T(d,\delta) = 2\sqrt{\Delta} H\left(\displaystyle\frac{\delta u}{\delta
s}\displaystyle\frac{dv}{ds} - \displaystyle\frac{\delta v}{\delta s}\displaystyle\frac{du}{ds}\right).
\end{equation}
It has the following nice consequence:
\begin{theorem}
The geodesic torsion $T(\delta,d)$ is symmetric with respect of
directions $\delta$ and $d$, if and only if $S$ is a minimal surface.
\end{theorem}
Finally, $(C,\overline{\xi})$ is orthogonally conjugated with
$(C,\overline{\alpha})$ if and only if $T(\delta,d) = 0$.
\begin{remark}\rm
The invariant $K(\delta,d)$ was discovered by O. Mayer [20] and the
invariant $T(\delta,d)$ was found by E. Bertoltti [3].
\end{remark}
\section{The Levi-Civita parallelism of vectors tangent to $S$}
\setcounter{theorem}{0}\setcounter{equation}{0}
The notion of Myller parallelism of vector field
$(C,\overline{V})$ tangent to $S$ along curve $C$, in the
associated Myller configuration $\mathfrak{M}_{t}(C,
\overline{\xi},\pi)$ is exactly the Levi-Civita parallelism of tangent
vector field $(C,\overline{V})$ to $S$ along the curve $C$. Indeed,
taking into account that the tangent versor field
$(C,\overline{\xi})$ has
\begin{equation}
\overline{\xi}(s) = \xi^{1}(s)\overline{r}_{u} +
\xi^{2}(s)\overline{r}_{v}
\end{equation}
and expression of the operator $\displaystyle\frac{D}{ds}$ is
\begin{equation}
\displaystyle\frac{D\xi^{i}}{ds} = \displaystyle\frac{d\xi^{i}}{ds} +
\sum_{j,k=1}^{2}\xi^{j}\left\{\begin{array}{cc} i\\jk
\end{array}\right\} \displaystyle\frac{du^{k}}{ds},\; (i=1,2)
\end{equation}
we have
$$
G^{ij} = \displaystyle\frac{D\xi^{i}}{ds}\xi^{j} -
\displaystyle\frac{D\xi^{j}}{ds}\xi^{i},\; (i=1,2).
$$
Thus the parallelism of versor $(C,\overline{\xi})$ along $C$ on
$S$ is expressed by $G^{ij} = 0.$ But these equations are
equivalent to
\begin{equation}
\displaystyle\frac{D(\lambda(s)\xi^{i}(s))}{ds} = 0\; (i=1,2),\; (\lambda(s)\neq 0).
\end{equation}
This is the definition of Levi-Civita parallelism of vectors
$\overline{V}(s)=\lambda(s)\overline{\xi}(s)$, $(\lambda(s)=\|V\|)$ tangent to $S$ along
$C$. Writing $\overline{V} = V^{1}(s) \overline{r}_{u} +
V^{2}(s)\overline{r}_{v}$ and putting
\begin{equation}
\displaystyle\frac{DV^{i}}{ds} =
\displaystyle\frac{dV^{i}}{ds}+\sum_{j,k=1}^{2}V^{j}{i\brace jk}
\displaystyle\frac{du^{k}}{ds}
\end{equation}
the Levi-Civita parallelism is expressed by
\begin{equation}
\displaystyle\frac{DV^{i}}{ds} = 0.
\end{equation}
In the case $\overline{V}(s)=\overline{\xi}(s)$ is a versor field,
parallelism of $(C,\overline{\xi})$ in the associated Myller
configurations $\mathfrak{M}_{t}(C, \overline{\xi},\pi)$ is called
the parallelism of directions tangent to $S$ along $C$. We can
express the condition of parallelism by means of invariants of the
field $(C,\overline{\xi})$, as follows:
1. {\it $(C,\overline{\xi})$ is parallel in the Levi-Civita sense on $S$
along $C$ iff $G(\delta,d)=0.$}
2. {\it $(C, \overline{\xi})$ is parallel in the Levi-Civita sense on $S$
along $C$ iff the versor $\overline{\xi}(s^{\prime})$,
$\{|s^{\prime}-s|<\varepsilon, \varepsilon>0, s^{\prime}\in (s_{1},s_{2})\}$ is
parallel in the space $E_{3}$ with the normal plan $(P(s);
\overline{\xi}(s), \overline{\nu}(s))$.}
3. {\it A necessary and sufficient condition for the versor field $(C,
\overline{\xi})$ to be parallel on $S$ along $C$ in Levi-Civita sense
is that developing on a plane the ruled surface $E$ generated by
tangent planes field $(C,\pi)$ along $C$ to $E$- the directions
$(C,\overline{\xi})$ after developing to be parallel in Euclidean
sense.}
4. {\it The directions $(C,\overline{\xi})$ are parallel in the Levi-Civita
sense on $S$ along $C$ iff the versor field $(C,\overline{\xi}_{2})$ is normal to $S$.}
5. {\it $(C,\overline{\xi})$ is parallel on $S$ along $C$, iff $|K| =
K_{1}$.}
6. {\it If $(C,\xi)$ is parallel on $S$ along $C$, then $|K| = K_{1},
T=K_{2}$.}
7. {\it If the ruled surface $R(C,\overline{\xi})$ is not developable,
then $(C,\overline{\xi})$ is parallel in the Levi-Civita sense iff
$C$ is the striction line of surface $R(C,\overline{\xi})$.}
Taking into account the system of differential equations (4.5.5)
in the given initial conditions, we have assured the existence and
uniqueness of the Levi-Civita parallel versor fields on $S$ along
$C$.
Other results presented in Section 9, Chapter 2 can be
particularized here without difficulties.
A first application. The Tchebishev nets on $S$ are defined as a
net of $S$ for which the tangent lines to a family of curves of
net are parallel on $S$ along to every curve of another family of
curves of net and conversely.
\bigskip
\noindent{\bf A Bianchi result}
\begin{theorem}
In order that the net parameter $(u = u_{0}, v=v_{0})$ on $S$ to
be a Tchebishev is necessary and sufficient to have the following
conditions:
\begin{equation}
{1\brace 12}
= {2\brace 12}
= 0.
\end{equation}
\end{theorem}
Indeed, we have $G(\delta,d) = G(d,\delta) = 0$ if (4.5.6) holds.
But (4.5.6) is equivalent to the equations $\displaystyle\frac{\partial E}{\partial v} =
\displaystyle\frac{\partial G}{\partial u} = 0.$ So, with respect to a Tchebishev
parametrization of $S$ its arclength element $ds^{2} = \phi(d,d)$
is given by
$$
ds^{2} = E(u)du^{2} + 2F (u,v)du dv + G(v)dv^{2}.
$$
We finish this paragraph remarking that the parallelism of vectors
$(C, \overline{V})$ on $S$ or the concurrence of vectors
$(C,\overline{V})$ on $S$ can be studied using the corresponding
notions in configurations $\mathfrak{M}_{t}$ described in the
Chapter 3.
\section{The geometry of curves on a surface}
\setcounter{theorem}{0}\setcounter{equation}{0}
The geometric theory of curves $C$ embedded in a surface $S$ can
be derived from the geometry of tangent Myller configuration
$\mathfrak{M}_{t}(C,\overline{\alpha},\pi)$ in which
$\overline{\alpha}(s)$ is tangent versor to $C$ at point $P(s)\in C$
and $\pi(s)$ is tangent plane to $S$ at point $P$ for any $s\in
(s_{1},s_{2})$.
Since $\mathfrak{M}_{t}(c,\overline{\alpha},\pi)$ is geometrically
associated to the curve $C$ on $S$ we can define its Darboux frame
$\cal{R}_{D}$, determine the moving equations of $\cal{R}_{D}$ and
its invariants, as belonging to curve $C$ on surface $S$.
Applying the results established in Chapter 3, first of all we
have
\begin{theorem}
For a smooth curve $C$ embedded in a surface $S$, there exists a
Darboux frame $\cal{R}_{D} = (P(s); \overline{\alpha}(s),
\overline{\mu}^{*}(s), \overline{\nu}(s))$ and a system of
invariants $\kappa_{g}(s), \kappa_{n}(s)$ and $\tau_{g}(s)$, satisfying
the following moving equations
\begin{equation}
\displaystyle\frac{d\overline{r}}{ds} = \overline{\alpha}(s),\;\; \forall s\in
(s_{1},s_{2})
\end{equation}
and
\begin{eqnarray}
\displaystyle\frac{d\overline{\alpha}}{ds} &=& \kappa_{g}(s)\overline{\mu}^{*}
+\kappa_{n}(s)\overline{\nu},\nonumber\\\displaystyle\frac{d\overline{\mu}^{*}}{ds}
&=& -\kappa_{g}(s)\overline{\alpha} +
\tau_{g}(s)\overline{\nu},\\\displaystyle\frac{d\overline{\nu}}{ds}&=&
-\kappa_{n}(s)\overline{\alpha} - \tau_{g}(s)\overline{\mu}^{*},\;\;
\forall s\in (s_{1},s_{2}).\nonumber
\end{eqnarray}
\end{theorem}
The functions $\kappa_{g}(s), \kappa_{n}(s), \tau_{g}(s)$ are called the
{\it geodesic curvature}, the {\it normal curvature} and {\it
geodesic torsion} of curve $C$ at point $P(s)$ on surface $S,$
respectively.
Exactly as in Chapter 3, a fundamental theorem can be enounced and
proved.
The invariants $\kappa_{g}, \kappa_{n}$ and $\tau_{g}$ have the same
values along $C$ on any smooth surface $S^{\prime}$ which passes
throught curve $C$ and is tangent to surface $S$ along
$C$.
The geometric interpretations of these invariants and the cases of
curves $C$ for which $\kappa_{g}(s) = 0,$ or $\kappa_{n}(s) = 0$ or
$\tau_{g}(s)=0$ can be studied as in Chapter 3.
The curves $C$ on $S$ for which $\kappa_{g}(s) = 0$ are called (as in
Chapter 3) geodesics (or {\it autoparallel curves}) of $S$. If $C$
has the property $\kappa_{n}(s) = 0,$ $\forall s\in (s_{1},s_{2})$, it
is {\it asymptotic curve} of $S$ and the curve $C$ for which
$\tau_{g}(s) = 0,$ $\forall s\in(s_{1},s_{2})$ is the {\it
curvature line} of $S$.
The expressions of these invariants can be obtained from those of
the invariants $G(\delta,d)$, $K(\delta,d)$ and $T(\delta,d)$ for
$\overline{\xi}(s) = \overline{\alpha}(s)$ given in Chapter 4:
\begin{equation}
\kappa_{g} = G(d,d),\; \kappa_{n}(s) = K(d,d),\;\; \tau_{g}(s) = T(d,d).
\end{equation}
So, we have
\begin{equation}
\hspace*{10mm}\begin{array}{lll}
\kappa_{g}&=&
\sqrt{\Delta}\left\{\displaystyle\frac{D}{ds}\left(\displaystyle\frac{du^{i}}{ds}\right)\displaystyle\frac{du^{j}}{ds}-\displaystyle\frac{D}{ds}
\left(\displaystyle\frac{du^{j}}{ds}\right)\displaystyle\frac{du^{i}}{ds}\right\},\ (i,j=1,2)\vspace*{2mm}\\
\kappa_{n}&=& \displaystyle\frac{L du^{2} + 2M du dv +N dv^{2}}{E du^{2} + 2F du dv
+ G dv^{2}},\vspace*{2mm}\\ \tau_{g}&=& \displaystyle\frac{1}{\sqrt{\Delta}}
\displaystyle\frac{\left\|\begin{array}{cc}L du + Mdv& Mdu + N dv\vspace*{2mm}\\ Edu + Fdv &
Fdu + Gdv
\end{array}\right\|}{Edu^{2} + 2F du dv + Gdv^{2}}.
\end{array}
\end{equation}
Evidently, $\kappa_{g}$ is an intrinsic invariant of surface $S$ and
$\kappa_{n}$ is the ratio of fundamental forms $\psi$ and $\Phi$.
The Mark Krein formula (3.4.3), Chapter 3 gives now the
Gauss-Bonnet formula for a surface $S$.
\newpage
\section{The formulae of O. Mayer and E. Bortolotti}
\setcounter{theorem}{0}\setcounter{equation}{0}
The geometers O. Mayer and E. Bortolotti gave some new forms of
the invariants $K(\delta,d)$ and $T(\delta,d)$ which generalize the Euler
or Bonnet formulae from the geometry of surfaces in Euclidean
space $E_{3}$.
Let $S$ be a smooth surface in $E_{3}$ having the parametrization
given by curvature lines. Thus the coefficients $F$ and $M$ of the
first fundamental and the second fundamental form vanish.
We denote by $\theta = \sphericalangle \(\overline{\alpha},
\displaystyle\frac{\overline{r}_{u}}{\sqrt{E}}\)$ and obtain
\begin{equation}
\cos \theta = \displaystyle\frac{\sqrt{E}du}{\sqrt{Edu^{2} + G dv^{2}}},\;\; \sin
\theta = \displaystyle\frac{\sqrt{G}dv}{\sqrt{Edu^{2} + G dv^{2}}}.
\end{equation}
The principal curvature are expressed by
\begin{equation}
\displaystyle\frac{1}{R_{1}} = \displaystyle\frac{L}{E},\;\; \displaystyle\frac{1}{R_{2}} = \displaystyle\frac{N}{G}.
\end{equation}
The mean curvature and total curvature are
\begin{eqnarray}
H&=&\displaystyle\frac{1}{2}\left(\displaystyle\frac{1}{R_{1}} + \displaystyle\frac{1}{R_{2}}\right) =
\displaystyle\frac{1}{2}\left(\displaystyle\frac{L}{E} +
\displaystyle\frac{N}{G}\right)\nonumber\\K_{t}&=&\displaystyle\frac{1}{R_{1}}\displaystyle\frac{1}{R_{2}}
= \displaystyle\frac{LN}{EG}.
\end{eqnarray}
Consider a tangent versor field $(C,\overline{\xi})$,
$\overline{\xi} = \overline{r}_{u}\displaystyle\frac{\delta u}{\delta s} +
\overline{r}_{v}\displaystyle\frac{\delta v}{\delta s}$ and let be $\sigma =
\sphericalangle
\(\overline{\xi},\displaystyle\frac{\overline{r}_{u}}{\sqrt{E}}\)$. One gets
\begin{equation}
\cos \sigma = \displaystyle\frac{\sqrt{E}\delta u}{\sqrt{E\delta u^{2} + G\delta v^{2}}},\;
\sin \sigma = \displaystyle\frac{\sqrt{G}\delta v}{\sqrt{E \delta u^{2} + G\delta v^{2}}}.
\end{equation}
Consequently, the expressions (4.4.13) and (4.4.17) of
the normal \linebreak curvature and geodesic torsion of
$(C,\overline{\xi})$ on $S$ are as follows
\begin{equation}
K(\delta,d) = \displaystyle\frac{\cos \sigma \cos \theta}{R_{1}} + \displaystyle\frac{\sin \sigma \sin
\theta}{R_{2}}
\end{equation}
\begin{equation}
T(\delta,d) = \displaystyle\frac{\cos \sigma\sin \theta}{R_{2}} - \displaystyle\frac{\sin \sigma \cos
\theta}{R}.
\end{equation}
The first formula was established by O. Mayer [20] and second
formula was given by E. Bortolotti [3].
In the case $\overline{\xi}(s) = \overline{\alpha}(s)$ these formulae
reduce to the known Euler and Bonnet formulas, respectively:
\begin{eqnarray}
\kappa_{n}&=& \displaystyle\frac{\cos^{2}\theta}{R_{1}} +
\displaystyle\frac{\sin^{2}\theta}{R_{2}}\nonumber\\\tau_{g} &=&
\displaystyle\frac{1}{2}\left(\displaystyle\frac{1}{R_{2}} - \displaystyle\frac{1}{R_{1}}\right)\sin
2\theta.
\end{eqnarray}
For $\theta =0$ or $\theta = \displaystyle\frac{\pi}{2},\kappa_{n}$ is equal to
$\displaystyle\frac{1}{R_{1}}$ and $\displaystyle\frac{1}{R_{2}}$ respectively. For $\theta =
\pm \displaystyle\frac{\pi}{4}$, $\tau_{g}$ takes the extremal values.
\begin{equation}
\displaystyle\frac{1}{T_{1}} = \displaystyle\frac{1}{2}\left(\displaystyle\frac{1}{R_{2}} -
\displaystyle\frac{1}{R_{1}}\right),\; \displaystyle\frac{1}{T_{2}} =
-\displaystyle\frac{1}{2}\left(\displaystyle\frac{1}{R_{2}} - \displaystyle\frac{1}{R_{1}}\right).
\end{equation}
Thus
\begin{equation}
T_{m} = \displaystyle\frac{1}{2}\left(\displaystyle\frac{1}{T_{1}} +
\displaystyle\frac{1}{T_{2}}\right),\; T_{t} = \displaystyle\frac{1}{T_{1}}\displaystyle\frac{1}{T_{2}}
\end{equation}
are the {\it mean torsion} of $S$ at point $P\in S$ and the {\it total
torsion} of $S$ at point $P\in S.$
For surfaces $S$, $T_{m}$ and $T_{t}$ have the following
properties:
\begin{equation}
T_{m} = 0,\;\; T_{t} = -\displaystyle\frac{1}{4}\left(\displaystyle\frac{1}{R_{1}} -
\displaystyle\frac{1}{R_{2}}\right)^{2}.
\end{equation}
\begin{remark}\rm
1. As we will see in the next chapter the nonholomorphic manifolds in
$E_{3}$ have a nonvanishing mean torsion $T_{m}$.
2. $T_{t}$ from (4.7.10) gives us the Bacaloglu curvature of
surfaces [35].
\end{remark}
Consider in plane $\pi(s)$ the cartesian orthonormal frame $(P(s);
\overline{i}_{1}, \overline{i}_{2})$, $i_{1} =
\displaystyle\frac{\overline{r}_{u}}{\sqrt{E}}, i_{2} =
\displaystyle\frac{\overline{r}_{v}}{\sqrt{G}}$ and the point $Q\in \pi(s)$
with the coordinates $(x,y)$, given by
$$
\overrightarrow{PQ} = x\overline{i}_{1} + y \overline{i}_{2},\;
\overrightarrow{PQ} = |\kappa_{n}|^{-1}\overline{\alpha}.
$$
But $\overline{\alpha} =\cos \theta \overline{i}_{1} +\sin \theta i_{2}$. So
we have the coordinates $(x,y)$ of point $Q$:
\begin{equation}
x = |\kappa_{n}|^{-1}\cos \theta; \;\; y = |\kappa_{n}|^{-1}\sin \theta.
\end{equation}
The locus of the points $Q$, when $\theta$ is variable in interval
$(0,2\pi)$ is obtained eliminating the variable $\theta$ between the
formulae (4.7.7) and (4.7.11). One obtains a pair of conics:
\begin{equation}
\displaystyle\frac{x^{2}}{R_{1}} + \displaystyle\frac{y^{2}}{R_{2}} = \pm 1
\end{equation}
called the {\it Dupin indicatrix} of normal curvatures. This indicatrix
is important in the local study of surfaces $S$ in Euclidean
space.
Analogously we can introduce the Bonnet indicatrix. Consider in
the plane $\pi(s)$ tangent to $S$ at point $P(s)$ the frame
$(P(s), \overline{i}_{1}, \overline{i}_{2})$ and the point
$Q^{\prime}$ given by
$$
\overrightarrow{PQ}^{\prime} = |\tau_{g}|^{-1}\overline{\alpha} =
x\overline{i}_{1} +y \overline{i}_{2}.
$$
Then, the locus of the points $Q^{\prime}$, when $\theta$ verifies
(4.7.7) and $x =|\tau_{g}|^{-1}\cos \theta$, $y = |\tau_{g}|^{-1}\sin
\theta$, defines the Bonnet indicatrix of geodesic torsions:
\begin{equation}
\left(\displaystyle\frac{1}{R_{2}} - \displaystyle\frac{1}{R_{1}}\right)xy = \pm 1
\end{equation}
which, in general, is formed by a pair of conjugated equilateral
hyperbolas.
Of course, the relations between the indicatrix of Dupin and
Bonnet can be studied without difficulties.
Following the same way we can introduce the indicatrix of the
invariants $K(\delta,d)$ and $T(\delta,d)$.
So, consider the angles $\theta = \sphericalangle
(\overline{\alpha},i_{1})$, $\sigma =
\sphericalangle(\overline{\xi},\overline{i}_{1})$ and $U\in
\pi(s)$. The point $U$ has the coordinates $x,y$ with respect to
frame $(P(s); \overline{i}_{1}, i_{2})$ given by
\begin{equation}
x= |K(\delta, d)|^{-1}\cos \sigma,\;\; y = |K(\delta,d)|^{-1} \sin \sigma.
\end{equation}
The locus of points $U$ is obtained from (4.7.5) and (4.7.14):
\begin{equation}
\displaystyle\frac{x\cos \theta}{R_{1}} +\displaystyle\frac{y\sin \theta}{R_{2}} = \pm 1.
\end{equation}
Therefore (4.7.15) is the indicatrix of the normal curvature
$K(\delta,d)$ of versor field $(C,\overline{\xi}).$ It is a pair of
parallel straight lines.
Similarly, for
$$
x = |T(\delta,d)|^{-1}\cos \sigma,\;\; y = |T(\delta,d)|^{-1}\sin \sigma
$$
and the formula (4.7.6) we determine the indicatrix of geodesic
torsion $T_{g}(\delta,d)$ of the versor field $(C,\overline{\xi}):$
$$
\displaystyle\frac{x\sin \theta}{R_{2}} - \displaystyle\frac{y\cos \theta}{R_{1}} = \pm 1.
$$
It is a pair of parallel straight lines.
Finally, we can prove without difficulties the following formula:
\begin{equation}
\hspace*{14mm}\kappa_{n}(\theta)\kappa_{n}(\sigma)\hspace{-0.5mm}+\hspace{-0.5mm} \tau_{g}(\theta) \tau_{g}(\sigma)\hspace{-0.5mm}=\hspace{-0.5mm}2HK(\sigma,\theta)\cos
(\sigma-\theta)\hspace{-0.5mm}-\hspace{-0.5mm}K_{t}\cos 2(\sigma-\theta),
\end{equation}
where $\kappa_{n}(\theta)= \kappa_{n}(d,d),$ $\kappa_{n}(\sigma) = \kappa_{n}(\delta,\delta)$,
$K(\sigma,\theta) = K(d,\delta)$.
For $\sigma = \theta,$ (i.e. $\overline{\xi} = \overline{\alpha}$) the
previous formulas leads to a known Baltrami-Enneper formula:
\begin{equation}
\kappa_{n}^{2} + \tau_{g}^{2} - 2H \kappa_{n} +K_{t} = 0,
\end{equation}
for every point $P(s)\in S.$
Along the asymptotic curves we have $\kappa_{n} = 0,$ and the previous
equations give us the Enneper formula
$$
\tau_{g}^{2} +K_{t} = 0,\; (\kappa_{n}=0).
$$
Along to the curvature lines, $\tau_{g}=0$ and one obtains from
(4.7.17)
$$
\kappa_{n}^{2} - 2H \kappa_{n} +K_{t} = 0,\;\; (\tau_{g} = 0).
$$
The considerations made in Chapter 4 of the present book allow
to affirm that the applications of the theory of Myller
configurations the geometry of surfaces in Euclidean space $E_{3}$
are interesting. Of course, the notion of Myller configuration can
be extended to the geometry of nonholomonic manifolds in $E_{3}$
which will be studied in next chapter. It can be applied to the
theory of versor fields in $E_{3}$ which has numerous applications
to Mechanics, Hydrodynamics (see the papers by Gh. Gheorghiev and
collaborators).
Moreover, the Myller configurations can be defined and
investigated in Riemannian spaces and applied to the geometry of
submanifolds of these spaces, [24], [25]. They can be studied in the Finsler, Lagrange or Hamilton spaces [58], [68], [70].
\chapter*{Introduction}
In the differential geometry of curves in the Euclidean space $E_{3}$
one introduces, along a curve $C$, some versor fields, as tangent,
principal normal or binormal, as well as some plane fields as
osculating, normal or rectifying planes.
More generally, we can consider a versor field $(C,\overline{\xi})$
or a plane field $(C,\pi)$.
A pair $\{(C,\overline{\xi}), (C,\pi)\}$ for which
$\overline{\xi}\in \pi$, has been called in 1960 [23] by the present
author, a Myller configuration in the space $E_{3}$, denoted by
$\mathfrak{M}(C, \overline{\xi},\pi)$. When the planes $\pi$ are tangent to $C$ then we have a tangent
Myller configuration $\mathfrak{M}_{t}(C,\overline{\xi},\pi)$.
Academician Alexandru Myller studied in 1922 the notion of
parallelism of $(C,\overline{\xi})$ in the plane field $(C,\pi)$
obtaining an interesting generalization of the famous parallelism
of Levi-Civita on the curved surfaces. These investigations have
been continued by Octav Mayer which introduced new fundamental
invariants for $\mathfrak{M}(C, \overline{\xi}, \pi)$. The
importance of these studies was underlined by Levi Civita in
Addendum to his book {\it Lezioni di calcolo differentiale assoluto},
1925.
Now, we try to make a systematic presentation of the geometry of
Myller configurations $\mathfrak{M}(C,\overline{\xi},\pi)$ and
$\mathfrak{M}_{t}(C,\overline{\xi}, \pi)$ with applications to the
differential geometry of surfaces and to the geometry of
nonholonomic manifolds in the Euclidean space $E_{3}$.
Indeed, if $C$ is a curve on the surface $S\subset E_{3}$, $s$ is
the natural parameter of curve $C$ and $\overline{\xi}(s)$ is a
tangent versor field to $S$ along $C$ and $\pi(s)$ is tangent
planes field to $S$ along $C$, we have a tangent Myller
configuration $\mathfrak{M}_{t}(C,\overline{\xi},\pi)$ intrinsic
associated to the geometric objects $S,C,\overline{\xi}$.
Consequently, the geometry of the field $(C, \overline{\xi})$ on
surface $S$ is the geometry of the associated Myller
configurations $\mathfrak{M}_{t}(C,\overline{\xi},\pi)$. It is
remarkable that the geometric theory of $\mathfrak{M}_{t}$ is a
particular case of that of general Myller configuration
$\mathfrak{M}(C,\overline{\xi},\pi)$.
For a Myller configuration $\mathfrak{M}(C,\overline{\xi},\pi)$ we
determine a Darboux frame, the fundamental equations and a
complete system of invariants $G,K,T$ called, geodesic curvature,
normal curvature and geodesic torsion, respectively, of the versor
field $(C,\overline{\xi})$ in Myller configuration
$\mathfrak{M}(C,\overline{\xi}, \pi)$. A fundamental theorem, when
the functions $(G(s),K(s),$ $T(s))$ are given, can be
proven.
The invariant $G(s)$ was discovered by Al. Myller (and named by
him the deviation of parallelism). $G(s) = 0$ on curve $C$
characterizes the parallelism of versor field $(C,\overline{\xi})$
in $\mathfrak{M},$ [23], [24], [31], [32], [33]. The second invariant
$K(s)$ was introduced by O. Mayer (it was called the curvature of
parallelism). Third invariant $T(s)$ was found by E. Bortolotti
[3].
In the particular case, when $\mathfrak{M}$ is a tangent Myller
configuration $\mathfrak{M}_{t}(C,\overline{\xi},\pi)$ associated
to a tangent versor field $(C,\overline{\xi})$ on a surface $S$,
the $G(s)$ is an intrinsic invariant and $G(s)=0$ along $C$ leads
to the Levi Civita parallelism.
In configurations $\mathfrak{M}_{t}(C,\overline{\xi},\pi)$ there
exists a natural versor field $(C,\overline{\alpha})$, where
$\overline{\alpha}(s)$ are the tangent versors to curve $C$. The
versor field $(C,\overline{\alpha})$ has a Darboux frame $\cal{R} =
(P(s); \overline{\alpha}, \overline{\mu}^{*}, \overline{\nu})$ in
$\mathfrak{M}_{t}$ where $\overline{\nu}(s)$ is normal to plane
$\pi(s)$ and $\overline{\mu}^{*}(s) = \overline{\nu}(s)\times
\overline{\alpha}(s).$ The moving equations of $\cal{R}$ are:
\begin{eqnarray*} \displaystyle\frac{d\overline{r}}{ds} &=& \overline{\alpha}(s),\;\;
s\in (s_{1},s_{2})\\\displaystyle\frac{d\overline{\alpha}}{ds}&=&
\kappa_{g}(s)\overline{\mu}^{*} +\kappa_{n}(s)\overline{\nu},\\
\displaystyle\frac{d\overline{\mu}^{*}}{ds}&=& -\kappa_{g}(s)\overline{\alpha} +
\tau_{g}(s)\overline{\nu},\\\displaystyle\frac{d\overline{\nu}}{ds}&=&
-\kappa_{n}(s)\overline{\alpha} - \tau_{g}(s)\overline{\mu}^{*}.
\end{eqnarray*}
The functions $\kappa_{g}(s), \kappa_{n}(s)$ and $\tau_{g}(s)$ form a
complete system of invariants of the curve in $\mathfrak{M}_{t}$.
A theorem of existence and uniqueness for the versor fields
$(C,\overline{\alpha})$ in $\mathfrak{M}_{t}(C,$ $\overline{\alpha},\pi)$,
when the invariant $\kappa_{g}(s)$, $k_n$ and $\tau_{g}(s)$ are given, is proved. The function $\kappa_{g}(s)$ is called the geodesic
curvature of the curve $C$ in $\mathfrak{M}_{t}$; $\kappa_{n}(s)$ is
the normal curvature and $\tau_{g}(s)$ is the geodesic torsion of
$C$ in $\mathfrak{M}_{t}$.
The condition $\kappa_{g}(s) = 0$, $\forall s\in (s_{1},s_{2})$
characterizes the geodesic (autoparallel lines) in
$\mathfrak{M}_{t}$; $\kappa_{n}(s) = 0,$ $\forall s\in (s_{1},s_{2})$
give us the asymptotic lines and $\tau_{g}(s) = 0,$ $\forall s\in
(s_{1},s_{2})$ characterizes the curvature lines $C$ in
$\mathfrak{M}_{t}$
One can remark that in the case when $\mathfrak{M}_{t}$ is the
associated Myller confi\-gu\-ration to a curve $C$ on a surface $S$ we
obtain the classical theory of curves on surface $S$. It is
important to remark that the Mark Krien$'$s formula (2.10.3) leads
to the integral formula (2.10.9) of Gauss-Bonnet for surface $S$, studied by R. Miron in the book [62].
Also, if $C$ is a curve of a nonholonomic manifold $E_{3}^{2}$ in
$E_{3}$, we uniquely determine a Myller configuration
$\mathfrak{M}_{t}(C,\overline{\alpha},\pi)$ in which
$(C,\overline{\alpha})$ is the tangent versor field to $C$ and
$(C,\pi)$ is the tangent plane field to $E_{3}^{2}$ along $C$, [54], [64], [65].
In this case the geometry of $\mathfrak{M}_{t}$ is the geometry of
curves $C$ in the nonholonomic manifolds $E_{3}^{2}$. Some new
notions can be introduced as: concurrence in Myller sense of
versor fields $(C,\overline{\xi})$ on $E_{3}^{2}$, extremal
geodesic torsion, the mean torsion and total torsion of
$E_{3}^{2}$ at a point, a remarkable formula for geodesic torsion
and an indicatrix of Bonnet for geodesic torsion, which is not
reducible to a pair of equilateral hyperbolas, as in the case of
surfaces.
The nonholonomic planes, nonholonomic spheres of Gr. Moisil, can
be studied by means of techniques from the geometry of Myller
configurations $\mathfrak{M}_{t}$.
We finish the present introduction pointing out some important
developments of the geometry of Myller configurations:
The author extended the notion of Myller configuration in
Riemannian Geometry [24], [25], [26], [27]. Izu Vaisman [76] has studied
the Myller configurations in the symplectic geometry.
Mircea Craioveanu realized a nice theory of Myller configurations
in infinit dimensional Riemannian manifolds. Gheorghe Gheorghiev
developed the configuration $\mathfrak{M}$, [11], [55] in the
geometry of versor fields in the Euclidean space and applied it in
hydromechanics.
N.N. Mihalieanu studied the Myller configurations in Minkowski
spaces [61]. For Myller configurations in a Finsler, Lagrange or Hamilton spaces we refer to the paper [58] and to the books of R. Miron and M. Anastasiei [68], R. Miron, D. Hrimiuc, H. Shimada and S. Sab\u{a}u [70].
All these investigations underline the usefulness of geometry of
Myller configurations in differential geometry and its
applications.
\newpage
\cleardoublepage
\chapter{Versor fields. Plane fields in $E_{3}$}
First of all we investigate the geometry of a versor field
$(C,\overline{\xi})$ in the Euclidean space $E_{3}$, introducing an
invariant frame of Frenet type, the moving equations of this
frame, invariants and proving a fundamental theorem. The
invariants are called $K_{1}$-curvature and $K_{2}$-torsion of
$(C,\overline{\xi})$. Geometric interpretations for $K_{1}$ and
$K_{2}$ are pointed out. The parallelism of $(C,\overline{\xi})$,
concurrence of $(C,\overline{\xi})$ and the enveloping of versor
field $(C,\overline{\xi})$ are studied, too.
A similar study is made for the plane fields $(C,\pi)$ taking into
account the normal versor field $(C,\overline{\nu})$, with
$\overline{\nu}(s)$ a normal versor to the plane $\pi(s)$.
\section{Versor fields $(C,\overline{\xi})$}
\setcounter{theorem}{0}\setcounter{equation}{0}
In the Euclidian space a versor field $(C,\overline{\xi})$ can be
analytical represented in an orthonormal frame $\cal{R} =
(0;\overline{i}_{1}, \overline{i}_{2},\overline{i}_{3})$, by
\begin{equation}
\overline{r} = \overline{r}(s),\;\;\;\;\overline{\xi} = \overline{\xi}(s), \;\; s\in
(s_{1},s_{2})
\end{equation}
where $s$ is the arc length on curve $C$,
$$
\overline{r}(s) = \overline{OP}(s) = x(s)\overline{i}_{1} +
y(s)\overline{i}_{2} + z(s)\overline{i}_{3}\textrm{ and }
$$ $\overline{\xi}(s) =
\overline{\xi}(\overline{r}(s))\hspace{-0.5mm}=\hspace{-0.5mm} \overrightarrow{PQ} =
\xi^{1}(s)\overline{i}_{1}\hspace{-0.5mm} +\hspace{-0.5mm} \xi^{2}(s)\overline{i}_{2}\hspace{-0.5mm} +\hspace{-0.5mm}
\xi^{3}(s)\overline{i}_{3}$, $\|\overline{\xi}(s)\|^2\hspace{-0.5mm} =\hspace{-0.5mm}
\langle\overline{\xi}(s),\overline{\xi}(s)\rangle =1.$
\smallskip
All geometric objects considered in this book are assumed to be of
class $C^{k}$, $k\geq 3$, and sometimes of class $C^{\infty}$. The
pair $(C,\overline{\xi})$ has a geometrical meaning. It follows
that the pair $\(C,\displaystyle\frac{d\overline{\xi}}{ds}\)$ has a geometric
meaning, too. Therefore, the norm:
\begin{equation}
K_{1}(s) = \left\|\displaystyle\frac{d\overline{\xi}(s)}{ds}\right\|
\end{equation}
is an invariant of the field $(C,\overline{\xi})$.
We denote
\begin{equation}
\overline{\xi}_1(s) = \overline{\xi}(s)
\end{equation}
and let $\overline{\xi}_{2}(s)$ be the versor of vector
$\displaystyle\frac{d\overline{\xi}_{1}}{ds}$. Thus we can write
$$
\displaystyle\frac{d\overline{\xi}_{1}(s)}{ds} =K_{1}(s)\overline{\xi}_{2}(s).
$$
Evidently, $\overline{\xi}_{2}(s)$ is orthogonal to
$\overline{\xi}_{1}(s)$.
It follows that the frame
\begin{equation}
\cal{R}_{F} = (P(s); \overline{\xi}_{1}, \overline{\xi}_{2},
\overline{\xi}_{3}),\;\; \overline{\xi}_{3}(s) =
\overline{\xi}_{1}\times \overline{\xi}_{2}
\end{equation}
is orthonormal, positively oriented and has a geometrical meaning.
$\cal{R}_{F}$ is called the Frenet frame of the versor field
$(C,\overline{\xi})$.
We have:
\begin{theorem}
The moving equations of the Frenet frame $\cal{R}_{F}$ are:
\begin{equation}
\hspace*{5mm}\displaystyle\frac{d\overline{r}}{ds} =
a_{1}(s)\overline{\xi}_{1}+a_{2}(s)\overline{\xi}_{2} +
a_{3}(s)\overline{\xi}_{3}\;\;\;\; a_{1}^{2}(s)+a_{2}^{2}(s) +
a_{3}^{2}(s) =1
\end{equation}
and
\begin{eqnarray}
\displaystyle\frac{d\overline{\xi}_{1}}{ds} &=&
K_{1}(s)\overline{\xi}_{2},\nonumber\\\displaystyle\frac{d\overline{\xi}_{2}}{ds}
&=& -K_{1}(s)\overline{\xi}_{1}(s) +
K_{2}(s)\overline{\xi}_{3},\\\displaystyle\frac{d\overline{\xi}_{3}}{ds}& =&
-K_{2}(s)\overline{\xi}_{2}(s),\nonumber
\end{eqnarray}
where $K_{1}(s)>0.$ The functions $K_{1}(s), K_{2}(s), a_{1}(s),
a_{2}(s)$, $a_{3}(s),$ $s\in (s_{1},s_{2})$ are invariants of the
versor field $(C,\overline{\xi})$.\end{theorem}
The proof does not present difficulties.
The invariant $K_{1}(s)$ is called the curvature of
$(C,\overline{\xi})$ and has the same geometric interpretation as
the curvature of a curve in $E_{3}$. $K_{2}(s)$ is called the
torsion and has the same geometrical interpretation as the torsion
of a curve in $E_{3}$.
The equations (1.1.5), (1.1.6) will be called the {\it fundamental or
Frenet equations} of the versor field $(C,\overline{\xi})$.
In the case $a_{1}(s)= 1,$ $a_{2}(s) = 0,$ $a_{3}(s) =0$ the
tangent versor $\displaystyle\frac{d\overline{r}}{ds}$ is denoted by
$$
\overline{\alpha}(s) =
\displaystyle\frac{d\overline{r}}{ds}(s)\leqno{(1.1.5')}
$$
The equations (1.1.5), (1.1.6) are then the Frenet equations of a
curve in the Euclidian space $E_{3}$.
For the versor field $(C,\overline{\xi})$ we can formulate a
fundamental theorem:
\begin{theorem}
If the functions $K_{1}(s)>0,$ $K_{2}(s)$, $a_{1}(s), a_{2}(s),
a_{3}(s)$, $(a_{1}^{2} + a_{2}^{2} + a_{3}^{2} =1)$ of class
$C^{\infty}$ are apriori given, $s\in [a,b]$ there exists a curve
$C:[a,b]\to E_{3}$ parametrized by arclengths and a versor field
$\overline{\xi}(s), s\in [a,b]$, whose the curvature, torsion and
the functions $a_{i}(s)$ are $K_{1}(s), K_{2}(s)$ and
$a_{i}(s)$. Any two such versor fields $(C,\overline{\xi})$ differ
by a proper Euclidean motion.
\end{theorem}
For the proof one applies the same technique like in the proof of
Theorem 11, p. 45 from [75], [76].
\begin{remark}\rm
1. If $K_{1}(s) = 0,$ $s\in (s_{1}, s_{2})$ the versors
$\overline{\xi}_{1}(s)$ are parallel in $E_{3}$ along the curve
$C.$
2. The versor field $(C,\overline{\xi})$ determines a ruled
surface $S(C,\overline{\xi})$.
3. The surface $S(C,\overline{\xi})$ is a cylinder iff the
invariant $K_{1}(s)$ vanishes.
4. The surface $S(C,\overline{\xi})$ is with director plane iff
$K_{2}(s) = 0$.
5. The surface $S(C,\overline{\xi})$ is developing iff the
invariant $a_{3}(s)$ vanishes.
\end{remark}
If the surfaces $S(C,\overline{\xi})$ is a cone we say that the
versors field $(C,\overline{\xi})$ is {\it concurrent}.
\begin{theorem}
A necessary and sufficient condition for the versor field $(C,\overline{\xi})$, $(K_{1}(s)\neq 0)$,
to be concurrent is the following
$$
\displaystyle\frac{d}{ds}\left(\displaystyle\frac{a_{2}(s)}{K_{1}(s)}\right)-a_{1}(s) =
0,\;\; a_{3}(s) = 0,\;\; \forall s\in (s_{1},s_{2}).
$$
\end{theorem}
\medskip
\section{Spherical image of a versor field $(C,\overline{\xi})$}
\setcounter{theorem}{0}\setcounter{equation}{0}
Consider the sphere $\Sigma$ with center a fix point $O\in E_{3}$
and radius 1.
\begin{definition}
The spherical image of the versor field $(C,\overline{\xi})$ is the
curve $C^{*}$ on sphere $\Sigma$ given by $\overline{\xi}^{*}(s) =
\overrightarrow{OP^{*}}(s) = \overline{\xi}(s),$ $\forall s\in
(s_{1},s_{2})$.
\end{definition}
From this definition we have:
\begin{equation}
d\overline{\xi}^{*}(s) =\overline{\xi}_{2}(s) K_{1}(s)ds.
\end{equation}
Some immediate properties:
1. It follows
\begin{equation}
ds^{*} = K_{1}(s)ds.
\end{equation}
Therefore:
The arc length of the curve $C^{*}$ is
\begin{equation}
s^{*} = s_{0}^{*} +\int_{s_{0}}^{s}K_{1}(\sigma) d\sigma
\end{equation}
where $s_{0}^{*}$ is a constant and $[s_{0},s]\subset
(s_{1},s_{2})$.
2. The curvature $K_{1}(s)$ of $(C,\overline{\xi})$ at a
point $P^{*}(s)$ is expressed by
\begin{equation}
K_{1}(s) = \displaystyle\frac{ds^*}{ds}
\end{equation}
3. $C^{*}$ is reducible to a point iff $(C,\overline{\xi})$ is a
parallel versor field in $E_{3}$.
4. The tangent line at point $P^{*}\in C^{*}$ is parallel with
principal normal line of $(C,\overline{\xi})$.
5. Since $\overline{\xi}_{1}(s)\times \overline{\xi}_{2}(s) =
\overline{\xi}_{3}(s)$ it follows that the direction of binormal
versor field $\overline{\xi}_{3}$ is the direction tangent to
$\Sigma$ orthogonal to tangent line of $C^{*}$ at point $P^{*}$.
6. The geodesic curvature $\kappa_{g}$ of the curve $C^{*}$ at a point
$P^{*}$ verifies the equation
\begin{equation}
\kappa_{g} ds^{*} =K_{2}ds.
\end{equation}
7. The versor field $(C,\overline{\xi})$ is of null torsion (i.e.
$K_{2}(s)=0$) iff $\kappa_{g} = 0.$ In this case $C^{*}$ is an arc of
a great circle on $\Sigma$.
\section{Plane fields $(C,\pi)$}
\setcounter{theorem}{0}\setcounter{equation}{0}
A plane field $(C,\pi)$ is defined by the versor field
$(C,\overline{\nu}(s))$ where $\overline{\nu}(s)$ is normal to
$\pi(s)$ in every point $P(s)\in C.$ We assume the $\pi(s)$ is
oriented. Consequently, $\overline{\nu}(s)$ is well determined.
Let $\cal{R}_{\pi} = (P(s), \overline{\nu}_{1}(s),
\overline{\nu}_{2}(s), \overline{\nu}_{3}(s))$ be the Frenet frame,
$\overline{\nu}(s) = \overline{\nu}_{1}(s)$, of the versor field
$(C, \overline{\nu})$.
It follows that the fundamental equations of the plane field
$(C,\pi)$ are:
\begin{equation}
\displaystyle\frac{d\overline{r}}{ds}(s) = b_{1}(s){\overline{\nu}}_{1} +
b_{2}(s){\overline{\nu}}_{2} + b_{3}(s)\nu_{3},\;\;
b_{1}^{2}+b_{2}^{2}+b_{3}^{2}=1.
\end{equation}
\begin{eqnarray}
\displaystyle\frac{d\overline{\nu}_{1}}{ds} &=& \chi_{1}(s)\overline{\nu}_{2},\nonumber\\
\displaystyle\frac{d\overline{\nu}_{2}}{ds} &=& -\chi_{1}\overline{\nu}_{1} +
\chi_{2}(s)\overline{\nu}_{3}\\
\displaystyle\frac{d\overline{\nu}_{3}}{ds}&=&
-\chi_{2}(s)\overline{\nu}_{2}.\nonumber
\end{eqnarray}
The invariant $\chi_{1}(s) = \left\|\displaystyle\frac{d\overline{\nu}_{1}}{dr}\right\|$
is called the curvature and $\chi_{2}(s)$ is the torsion of the
plane field $(C,\pi(s))$.
The following properties hold:
1. The characteristic straight lines of the field of planes
$(C,\pi)$ cross through the corresponding point $P(s)\in C$ iff the
invariant $b_{1}(s) = 0$, $\forall s\in (s_{1},s_{2})$.
2. The planes $\pi(s)$ are parallel along the curve $C$ iff the
invariant $\chi_{1}(s)$ vanishes, $\forall s\in (s_{1},s_{2})$.
3. The characteristic lines of the plane field $(C,\pi)$ are
parallel iff the invariant $\chi_{2}(s) = 0$, $\forall s\in
(s_{1},s_{2})$.
4. The versor field $(C,\overline{\nu}_{3})$ determines the
directions of the characteristic line of $(C,\pi)$.
5. The versor field $(C,\overline{\nu}_{3})$, with
$\chi_{2}(s)\neq 0$, is concurrent iff:
$$
b_{1}(s) = 0,\;\; b_{3}(s) +
\displaystyle\frac{d}{dt}\left(\displaystyle\frac{b_{2}(s)}{\chi_{2}(s)}\right)=0.
$$
6. The curve $C$ is on orthogonal trajectory of the generatrices of
the ruled surface $\cal{R}(C,\overline{\nu}_{3})$ if $b_{3}(s) =
0,$ $\forall s\in (s_{1},s_{2})$.
7. By means of equations (1.3.1), (1.3.2) we can prove a
fundamental theorem for the plane field $(C,\pi).$
\chapter{Myller configurations $\mathfrak{M}(C,\overline{\xi},\pi)$}
The notions of versor field $(C,\overline{\xi})$ and the plane
field $(C,\pi)$ along to the same curve $C$ lead to a more general
concept named Myller configuration
$\mathfrak{M}(C,\overline{\xi},\pi)$, in which every versor
$\overline{\xi}(s)$ belongs to the corresponding plane $\pi(s)$ at
point $P(s)\in C.$ The geometry of $\mathfrak{M} =
\mathfrak{M}(C,\overline{\xi},\pi)$ is much more rich as the
geometries of $(C,\overline{\xi})$ and $(C,\pi)$ separately taken.
For $\mathfrak{M}$ one can define its geometric invariants, a Darboux frame
and introduce a new idea of parallelism or concurrence of versor
$(C,\overline{\xi})$ in $\mathfrak{M}$. The geometry of
$\mathfrak{M}$ is totally based on the fundamental equations of
$\mathfrak{M}$. The basic idea of this construction belongs to Al.
Myller [31], [32], [33], [34] and it was considerable developed by O. Mayer
[20], [21], R. Miron [23], [24], [62] (who proposed the name of Myller
Configuration and studied its complete system of invariants).
\section{Fundamental equations of Myller configuration}
\setcounter{theorem}{0}\setcounter{equation}{0}
\begin{definition}
A Myller configuration $\mathfrak{M} =
\mathfrak{M}(C,\overline{\xi},\pi)$ in the Euclidean space $E_{3}$
is a pair $(C,\overline{\xi})$, $(C,\pi)$ of versor field and
plane field, having the property: every $\overline{\xi}(s)$
belongs to the plane $\pi(s)$.
\end{definition}
Let $\overline{\nu}(s)$ be the normal versor to plane $\pi(s)$.
Evidently $\overline{\nu}(s)$ is uniquely determined if $\pi(s)$
is an oriented plane for $\forall s\in (s_{1},s_{2})$.
By means of versors $\overline{\xi}(s), \overline{\nu}(s)$ we can
determine the {\it Darboux frame} of $\mathfrak{M}:$
\begin{equation}
\cal{R}_{D} = (P(s); \overline{\xi}(s), \overline{\mu}(s),
\overline{\nu}(s)),
\end{equation}
where
\begin{equation}
\overline{\mu}(s) = \overline{\nu}(s)\times \overline{\xi}(s).
\end{equation}
$\cal{R}_{D}$ is geometrically associated to
$\mathfrak{M}$. It is orthonormal and positively oriented.
Since the versors $\overline{\xi}(s), \overline{\mu}(s),
\overline{\nu}(s)$ have a geometric meaning, the same pro\-per\-ties
have the vectors $\displaystyle\frac{d\overline{\xi}}{ds},$
$\displaystyle\frac{d\overline{\mu}}{ds}$ and $\displaystyle\frac{d\overline{\nu}}{ds}$.
Therefore, we can prove, without difficulties:
\begin{theorem}
The moving equations of the Darboux frame of $\mathfrak{M}$ are as
follows:
\begin{equation}
\displaystyle\frac{d\overline{r}}{ds} = \overline{\alpha}(s) =
c_{1}(s)\overline{\xi} + c_{2}(s)\overline{\mu} +
c_{3}(s)\overline{\nu};\;\; c_{1}^{2} + c_{2}^{2}+c_{3}^{2} =1
\end{equation}
and
\begin{eqnarray}
\displaystyle\frac{d\overline{\xi}}{ds} &=& G(s)\overline{\mu} +
K(s)\overline{\nu},\nonumber\\
\displaystyle\frac{d\overline{\mu}}{ds} &=& -G(s)\overline{\xi} +
T(s)\overline{\nu},\\\displaystyle\frac{d\overline{\nu}}{ds} &=&
-K(s)\overline{\xi} - T(s)\overline{\mu}\nonumber
\end{eqnarray}
and $c_{1}(s), c_{2}(s), c_{3}(s); G(s), K(s)$ and $T(s)$ are uniquely determined and are invariants.
\end{theorem}
The previous equations are called {\it the fundamental equations}
of the Myller configurations $\mathfrak{M}$.
Terms:
$G(s)$ -- is the geodesic curvature, $K(s)$ -- is the normal
curvature, $T(s)$ -- is the geodesic torsion of the versor field $(C,\overline{\xi})$ in Myller configuration $\mathfrak{M}$.
For $\mathfrak{M}$ a fundamental theorem can be stated:
\begin{theorem}
Let be a priori given $C^{\infty}$ functions $c_{1}(s), c_{2}(s),
c_{3}(s)$, $[c_{1}^{2} + c_{2}^{2} + c_{3}^{2}=1]$, $G(s),$
$K(s)$, $T(s)$, $s\in [a,b]$. Then there is a Myller configuration
$\mathfrak{M}(C,\overline{\xi},\pi)$ for which $s$ is the arclength of
curve $C$ and the given functions are its invariants. Two such
configuration differ by a proper Euclidean motion.
\end{theorem}
\noindent{\it Proof.}
By means of given functions $c_{1}(s), \ldots, G(s), \ldots$ we
can write the system of differential equations (2.1.3), (2.1.4).
Let the initial conditions $\overline{r}_{0} =
\overrightarrow{OP}_{0}$, $(\overline{\xi}_{0},
\overline{\mu}_{0}, \overline{\nu}_{0})$, an orthonormal,
positively oriented frame in $E_{3}$.
From (2.1.4) we find an unique solution $(\overline{\xi}(s),
\overline{\mu}(s),\overline{\nu}(s))$, $s\in [a,b]$ with the
property
$$
\overline{\xi}(s_{0}) = \overline{\xi}_{0},\;\;
\overline{\mu}(s_{0}) = \overline{\mu}_{0},\;\;
\overline{\nu}(s_{0}) = \overline{\nu}_{0},
$$
with $s_{0}\in [a,b]$ and $(\overline{\xi}(s), \overline{\mu}(s),
\overline{\nu} (s))$ being an orthonormal, positively oriented
frame.
Then consider the following solution of (2.1.3)
$$
\overline{r}(s) = \overline{r}_{0} +
\int_{s_{0}}^{s}[c_{1}(\sigma)\overline{\xi}(\sigma) +
c_{2}(\sigma)\overline{\mu}(\sigma) +c_{3}(\sigma)\overline{\nu}(\sigma)]d\sigma,
$$
which has the property $\overline{r}(s_{0}) = \overline{r}_{0}$,
and $\left\|\displaystyle\frac{d\overline{r}}{ds}\right\| =1.$
Thus $s$ is the arc length on the curve $\overline{r} = r(s)$.
Now, consider the configuration
$\mathfrak{M}(C,\overline{\xi},\pi(s))$, $\pi(s)$ being the plane
orthogonal to versor $\overline{\nu}(s)$ at point $P(s)$.
We can prove that $\mathfrak{M}(C,\overline{\xi},\pi)$ has as
invariants just $c_{1},c_{2},c_{3},G,K,T.$
The fact that two configurations $\mathfrak{M}$ and
$\mathfrak{M}^{\prime}$, obtained by changing the initial
conditions $(\overline{r}_{0}; \overline{\xi}_{0},
\overline{\mu}_{0}, \overline{\nu}_{0})$ to
$(\overline{r}_{0}^{\prime}, \overline{\xi}_{0}^{\prime},
\overline{\mu}_{0}^{\prime}, \overline{\nu}_{0}^{\prime})$ differ
by a proper Euclidian motions follows from the property that the
exists an unique Euclidean motion which apply $(\overline{r}_{0};
\overline{\xi}_{0}, \overline{\mu}_{0},\overline{\nu}_{0})$ to
$(\overline{r}^{\prime}_{0}, \overline{\xi}_{0}^{\prime},
\overline{\mu}_{0}^{\prime},
\overline{\nu}_{0}^{\prime})$.
\section{Geometric interpretations of invariants}
\setcounter{theorem}{0}
\setcounter{equation}{0}
The invariants $c_{1}(s), c_{2}(s), c_{3}(s)$ have simple
geometric interpretations:
$$
c_{1}(s) = \cos \sphericalangle (\overline{\alpha}, \overline{\xi}),\;
c_{2}(s) =\cos \sphericalangle (\overline{\alpha}, \overline{\mu}),\;
c_{3} = \cos \sphericalangle (\overline{\alpha}, \overline{\nu}).
$$
We can find some interpretation of the invariants $G(s), K(s)$ and
$T(s)$ considering a variation of Darboux frame
$$
R_{D}(P(s); \overline{\xi}(s),
\overline{\mu}(s),\overline{\nu}(s))$$$$\to
R^{\prime}_{D}(P^{\prime}(s+\Delta s), \overline{\xi}(s+\Delta s),
\overline{\mu}(s+\Delta s), \overline{\nu}(s+\Delta s)),
$$
obtained by the Taylor expansion
\begin{equation}
\begin{array}{lll}
\overline{r}(s+\Delta s)&=& \overline{r}(s)+\displaystyle\frac{\Delta s
}{1!}\displaystyle\frac{d\overline{r}}{ds} + \displaystyle\frac{(\Delta s)^{2}}{2!}
\displaystyle\frac{d^{2}r}{ds^{2}} + \ldots +\vspace*{2mm}\\ & & +\displaystyle\frac{(\Delta
s)^{n}}{n!}\left(\displaystyle\frac{d^{n}\overline{r}}{ds^{n}}
+ \overline{\omega}_{0}(s,\Delta
s)\right),\\\overline{\xi}(s+\Delta s)&=&
\overline{\xi}(s)+\displaystyle\frac{\Delta s }{1!}\displaystyle\frac{d\overline{\xi}}{ds} +
\displaystyle\frac{(\Delta s)^{2}}{2!} \displaystyle\frac{d^{2}\overline{\xi}}{ds^{2}} + \ldots
+\\ & & +\displaystyle\frac{(\Delta s)^{n}}{n!}\left(\displaystyle\frac{d^{n}\overline{\xi}}{ds^{n}}
+ \overline{\omega}_{1}(s,\Delta
s)\right),\vspace*{2mm}\\\overline{\mu}(s+\Delta s)&=&
\overline{\mu}(s)+\displaystyle\frac{\Delta s }{1!}\displaystyle\frac{d\overline{\mu}}{ds} +
\displaystyle\frac{(\Delta s)^{2}}{2!} \displaystyle\frac{d^{2}\overline{\mu}}{ds^{2}} + \ldots
+\\ & & +\displaystyle\frac{(\Delta
s)^{n}}{n!}\left(\displaystyle\frac{d^{n}\overline{\mu}}{ds^{n}}+ \overline{\omega}_{2}(s,\Delta
s)\right),\vspace*{2mm}\\\overline{\nu}(s+\Delta s)&=& \overline{\nu}(s)+\displaystyle\frac{\Delta s
}{1!}\displaystyle\frac{d\overline{\nu}}{ds} + \displaystyle\frac{(\Delta s)^{2}}{2!}
\displaystyle\frac{d^{2}\overline{\nu}}{ds^{2}} + \ldots +\vspace*{2mm}\\ & &\displaystyle\frac{(\Delta
s)^{n}}{n!}\left(\displaystyle\frac{d^{n}\overline{\nu}}{ds^{n}}+ \overline{\omega}_{3}(s,\Delta s)\right),
\end{array}
\end{equation}
where
\begin{eqnarray}
\lim\limits_{\Delta s\to 0}\overline{\omega}_{i}(s,\Delta s) = 0,\;\;
(i=0,1,2,3).
\end{eqnarray}
Using the fundamental formulas (2.1.3), (2.1.4) we can write, for
$n=1:$
\begin{eqnarray}
\overline{r}(s+\Delta s) &=& \overline{r}(s) + \Delta s
(c_{1}\overline{\xi} + c_{2} \overline{\mu} + c_{3}\overline{\nu}
+ \overline{\omega}_{0}(s,\Delta s)),\nonumber\\\overline{\xi}(s+\Delta s) &=&
\overline{\xi}(s) + \Delta s(G(s) \overline{\mu} + K(s)\overline{\nu}
+ \overline{\omega}_{1}(s,\Delta s) ),\\\overline{\mu}(s+\Delta s) &=&
\overline{\mu}(s) + \Delta s (-G(s)\overline{\xi} + T(s)\overline{\nu}
+ \overline{\omega}_{2}(s,\Delta s) )\nonumber\\\overline{\nu}(s)+\Delta s &=&
\overline{\nu}(s) + \Delta s(-K(s)\overline{\xi} - T(s)\overline{\mu}
+ \overline{\omega}_{3}(s,\Delta s)).\nonumber
\end{eqnarray} and (2.2.2) being verified.
Let $\overline{\xi}^{*}(s+\Delta s)$ be the orthogonal projection of
the versor $\overline{\xi}(s+\Delta s)$ on the plane $\pi(s)$ at point
$P(s)$ and let $\Delta \psi_{1}$ be the oriented angle of the versors
$\overline{\xi}(s), \overline{\xi}^{*}(s+\Delta s)$. Thus, we have
\begin{theorem}
The invariant $G(s)$ of versor field $(C,\xi)$ in Myller
configuration $\mathfrak{M}(C,\overline{\xi},\pi)$ is given by
$$
G(s) = \lim_{\Delta s\to 0}\displaystyle\frac{\Delta\psi_{1}}{\Delta s}.
$$
\end{theorem}
By means of second formula (2.2.3), this Theorem can be proved
without difficulties.
Therefore the name of {\it geodesic curvature} of $(C,\overline{\xi})$
in $\mathfrak{M}$ is justified.
Consider the plane $(P(s);
\overline{\xi}(s), \overline{\nu}(s))$-called the normal plan of
$\mathfrak{M}$ which contains the versor $\overline{\xi}(s)$.
Let be the vector $\overline{\xi}^{**}(s+\Delta s)$ the orthogonal
projection of versor $\overline{\xi}(s+\Delta s)$ on the normal plan
$(P(s); \overline{\xi}(s), \overline{\nu}(s))$. The angle $\Delta
\psi_{2} = \sphericalangle (\overline{\xi}(s),
\overline{\xi}^{**}(s+\Delta s))$ is given by the forma
$$
\sin \Delta \psi_{2} = \displaystyle\frac{\langle\overline{\xi}(s),
\overline{\xi}(s+\Delta s),
\overline{\mu}(s)\rangle}{\|\overline{\xi}^{**}(s+\Delta s)\|}.
$$
By (2.2.3) we obtain
$$
\sin \Delta \psi_{2} = \displaystyle\frac{K(s) +\langle \overline{\xi}(s),
\overline{\omega}_{1}(s, \Delta s), \overline{\mu}(s)
\rangle}{\|\overline{\xi}^{**}(s+\Delta s)\|}\Delta s.
$$
Consequently, we have:
\begin{theorem}
The invariant $K(s)$ has the following geometric interpretation
$$
K(s) = \lim_{\Delta s\to 0}\displaystyle\frac{\Delta \psi_{2}}{\Delta s}.
$$\end{theorem}
Based on the previous result we can call {\it $K(s)$ the normal
curvature} of $(C,\overline{\xi})$ in $\mathfrak{M}$.
A similar interpretation can be done for the invariant $T(s)$.
\begin{theorem}
The function $T(s)$ has the interpretation:
$$
T(s) = \lim\limits_{\Delta s\to 0}\displaystyle\frac{\Delta \psi_{3}}{\Delta s},
$$
where $\Delta \psi_{3}$ is the oriented angle between
$\overline{\mu}(s)$ and $\overline{\mu}^{*}(s+\Delta s)$-which is the
orthogonal projection of $\overline{\mu}(s+\Delta s)$ on the normal
plane $(P(s); \overline{\mu}(s),$
$\overline{\nu}(s))$.\end{theorem}
This geometric interpretation allows to give the name {\it
geodesic torsion} for the invariant $T(s)$.
\section{The calculus of invariants $G, K, T$}
\setcounter{theorem}{0}
\setcounter{equation}{0}
The fundamental formulae (2.1.3), (2.1.4) allow to calculate the
expressions of second derivatives of the versors of Darboux frame
$\cal{R}_{D}$. We have:
\begin{equation}\hspace*{8mm}
\begin{array}{lll}
\displaystyle\frac{d^{2}\overline{r}}{ds^{2}}&=& \left(\displaystyle\frac{dc_{1}}{ds} -
Gc_{2}-K c_{3} \right)\overline{\xi}+\left(\displaystyle\frac{dc_{2}}{ds} +
Gc_{1} - Tc_{3} \right)\overline{\mu} +\vspace*{2mm}\\&&+\left(\displaystyle\frac{dc_{3}}{ds}
+ K c_{1} + Tc_{2}\right)\overline{\nu}
\end{array}
\end{equation}
and
\begin{equation}
\hspace*{8mm}\begin{array}{lll}
\displaystyle\frac{d^{2}\overline{\xi}}{ds^{2}} &=& -(G^{2} +
T^{2})\overline{\xi} + \left(\displaystyle\frac{dG}{ds} -
KT\right)\overline{\mu} +\left(\displaystyle\frac{dK}{ds} +
GT\right)\overline{\nu},\vspace*{2mm}\\ \displaystyle\frac{d^{2}\overline{\mu}}{ds^{2}}
&=& -\left(\displaystyle\frac{dG}{ds} +KT \right)\overline{\xi} -
(G^{2}+T^{2})\overline{\mu} + \left(\displaystyle\frac{dT}{ds} -
GT\right)\overline{\nu},\vspace*{2mm}\\ \displaystyle\frac{d^{2}\nu}{ds^{2}} &=&
\left(-\displaystyle\frac{dK}{ds}+GT\right)\overline{\xi} - \left(\displaystyle\frac{dT}{ds}
+GT \right)\overline{\mu} - (K^{2}+T^{2})\overline{\nu}.
\end{array}
\end{equation}
These formulae will be useful in the next part of the book.
From the fundamental equations (2.1.3), (2.1.4) we get
\begin{theorem}
The following formulae for invariants $G(s), K(s)$ and $T(s)$
hold:
\begin{equation}
G(s) = \left\langle \overline{\xi},
\displaystyle\frac{d\xi}{ds},\overline{\nu}\right\rangle,
\end{equation}
\begin{equation}
K(s) = \left\langle \displaystyle\frac{d\overline{\xi}}{ds},
\overline{\nu}\right\rangle = -\left\langle \overline{\xi},
\displaystyle\frac{d\overline{\nu}}{ds}\right\rangle, T(s)=\left\langle\overline{\xi},\overline{\nu},\displaystyle\frac{d\overline{\nu}}{ds}\right\rangle.
\end{equation}
\end{theorem}
Evidently, these formulae hold in the case when $s$ is the arclength of
the curve $C$.
\section{Relations between the invariants of the field $(C,\overline{\xi})$ and the invariants of $(C,\overline{\xi})$ in
$\mathfrak{M}(C,\overline{\xi},\pi)$}
\setcounter{theorem}{0}
\setcounter{equation}{0}
The versors field $(C,\overline{\xi})$ in $E_{3}$ has a Frenet
frame $\cal{R}_{F} = (P(s), \overline{\xi}_{1},
\overline{\xi}_{2}, \overline{\xi}_{3})$ and a complete system of
invariants $(a_{1},a_{2},a_{3}; K_{1}, K_{2})$ verifying the
equations (1.1.5) and (1.1.6).
The same field $(C,\overline{\xi})$ in Myller configuration
$\mathfrak{M}(C,\overline{\xi},\pi)$ has a Darboux frame
$\cal{R}_{D} = (P(s), \overline{\xi}, \overline{\mu},
\overline{\nu})$ and a complete system of invariants $(c_{1},
c_{2}, c_{3}; G,K,T)$. If we relate $\cal{R}_{F}$ to $\cal{R}_{D}$
we obtain
\begin{eqnarray*}
\overline{\xi}_{1}(s) &=&
\overline{\xi}(s),\\\overline{\xi}_{2}(s)&=& \overline{\mu}(s)\sin
\varphi +\overline{\nu}(s)\cos \varphi,\\\overline{\xi}_{3}(s) &=&-\mu(s)cos
\varphi + \overline{\nu}(s)\sin \varphi,
\end{eqnarray*}
with $\varphi = \sphericalangle (\overline{\xi}_{2},\overline{\nu})$
and
$\langle\overline{\xi}_{1},\overline{\xi}_{2},\overline{\xi}_{3}\rangle
= \langle\overline{\xi},\overline{\mu},\overline{\nu}\rangle =1.$
Then, from (1.1.5) and (2.1.3) we can determine the relations between the two
systems of invariants.
From
$$
\displaystyle\frac{d\overline{r}}{ds} = \overline{\alpha}(s) =
a_{1}\overline{\xi}_{1} + a_{2}\overline{\xi}_{2} +
a_{3}\overline{\xi}_{3} = c_{1}\overline{\xi} +
c_{2}\overline{\mu}+c_{3}\overline{\nu}
$$
it follows
\begin{eqnarray}
c_{1}(s)&=&a_{1}(s)\nonumber\\c_{2}(s) &= &a_{1}(s)\sin \varphi
-a_{3}(s)\cos \varphi\\c_{3}(s)&=& a_{2}(s)\cos \varphi + a_{3}(s)\sin
\varphi.\nonumber
\end{eqnarray}
And, (1.1.6), (2.1.4) we obtain
\begin{eqnarray}
G(s)&=& K_{1}(s) \sin \varphi\nonumber\\K(s) &=& K_{1}(s)\cos \varphi\\T(s)
&=& K_{2}(s) + \displaystyle\frac{d\varphi}{ds}.\nonumber
\end{eqnarray}
These formulae, allow to investigate some important properties of
$(C,\overline{\xi})$ in $\mathfrak{M}$ when some invariants
$G,K,T$ vanish.
\section{Relations between invariants of normal field $(C,\overline{\nu})$ and invariants $G,K,T$}
\setcounter{theorem}{0}
\setcounter{equation}{0}
The plane field $(C,\pi)$ is characterized by the normal versor
field $(C,\overline{\nu})$, which has as Frenet frame $\cal{R}_{F}
= (P(s); \overline{\nu}_{1}, \overline{\nu}_{2},
\overline{\nu}_{3})$ with $\overline{\nu}_{1} = \overline{\nu}$
and has $(b_{1}, b_{2}, b_{3}, \chi_{1},\chi_{2})$ as a complete
system of invariants. They satisfy the formulae (1.3.1), (1.3.2).
But the frame $\cal{R}_{F}$ is related to Darboux frame
$\cal{R}_{D}$ of $(C,\overline{\xi})$ in $\mathfrak{M}$ by the
formulae
\begin{eqnarray}
\overline{\nu}_{1} &=&
\overline{\nu}(s)\nonumber\\-\overline{\nu}_{2}&=& \sin \sigma
\overline{\xi} +\cos \sigma \overline{\mu}\\\overline{\nu}_{3}&=&
-\cos \sigma \overline{\xi}+\sin \sigma \overline{\mu}\nonumber
\end{eqnarray}
where $\sigma = \sphericalangle (\overline{\xi}(s),
\overline{\nu}_{3}(s))$.
Proceeding as in the previous section we deduce
\begin{theorem}
The following relations hold:
\begin{eqnarray}
c_{1}&=& -b_{2}\sin \sigma + b_{3}\cos \sigma\nonumber\\-c_{2}&=&
b_{2}\cos \sigma + b_{3}\sin \sigma\\c_{3}&=& b_{2}\nonumber
\end{eqnarray}
and
\begin{eqnarray}
K&=& \chi_{1}\sin \sigma\nonumber\\T&=& \chi_{1}\cos \sigma\\Gamma&=&
\chi_{2}+\displaystyle\frac{d\sigma}{ds}.\nonumber
\end{eqnarray}
\end{theorem}
A first consequence of previous formulae is given by
\begin{theorem}
The invariant $K^{2}+T^{2}$ depends only on the plane field
$(C,\pi)$. We have
\begin{equation}
K^{2}+T^{2} = \chi_{1}.
\end{equation}
\end{theorem}
The proof is immediate, from (2.5.3).
\section{Meusnier$^{\prime}$s theorem. Versor fields $(C,\overline{\xi})$ conjugated with tangent versor $(C,\overline{\alpha})$}
\setcounter{theorem}{0}
\setcounter{equation}{0}
Consider the vector field $\overline{\xi}^{**}(s+\Delta s)$, $(|\Delta
s|<\varepsilon, \varepsilon>0)$, the orthogonal projection of versor
$\overline{\xi}(s+\Delta s)$ on the normal plane $(P(s);
\overline{\xi}(s), \overline{\nu}(s))$. Since, up to terms of
second order in $\Delta s$, we have
$$
\overline{\xi}^{**}(s+\Delta s) = \overline{\xi}(s) + \displaystyle\frac{\Delta
s}{1!}K(s)\overline{\nu}(s) + \overline{\theta}(s,\Delta s)\displaystyle\frac{(\Delta
s)^{2}}{2!},
$$
for $\Delta s\to 0$ one gets:
\begin{equation}
\displaystyle\frac{d\overline{\xi}^{**}}{ds} = K(s)\overline{\nu}(s).
\end{equation}
Assuming $K(s)\neq 0$ we consider the point $P_{c}^{**}$-called
the {\it center of curvature} of the vector field
$(C,\overline{\xi}^{**})$, given by
$$
\overrightarrow{PP}_{c}^{**} = \displaystyle\frac{1}{K(s)}\overline{\nu}(s).
$$
On the other hand the field of versors $(C,\overline{\xi})$ have a
center of curvature $P_{c}$ given by $\overrightarrow{PP}_{c} =
\displaystyle\frac{1}{K_{1}(s)}\overline{\xi}_{2}$.
The formula (2.4.2), i.e., $K(s) = K_{1}(s)\cos \varphi,$ shows that
the orthogonal projection of vector $\overrightarrow{PP}_{c}^{**}$
on the (osculating) plane $(P(s); \overline{\xi}_{1},
\overline{\xi}_{2})$ is the vector $\overrightarrow{PP}_{c}$.
Indeed, we have
\begin{equation}
\displaystyle\frac{\cos \varphi}{K} = \displaystyle\frac{1}{K_{1}}
\end{equation}
As a consequence we obtain a theorem of Meusnier type:
\begin{theorem}
The curvature center of the field $(C,\overline{\xi})$ in
$\mathfrak{M}$ is the orthogonal projection on the osculating
plane $(M; \overline{\xi}_{1}, \overline{\xi}_{2})$ of the
curvature center $P_{c}^{**}$.
\end{theorem}
\begin{definition}
The versor field $(C,\overline{\xi})$ is called conjugated with
tangent versor field $(C,\overline{\alpha})$ in the Myller
configuration $\mathfrak{M}(C,\overline{\xi},\pi)$ if the
invariant $K(s)$ vanishes.
\end{definition}
Some immediate consequences:
1. $(C,\overline{\xi})$ is conjugated with $(C,\overline{\alpha})$ in
$\mathfrak{M}$ iff the line $(P;\overline{\xi})$ is parallel in
$E_{3}$ to the characteristic line of the planes $\pi(s)$, $s\in
(s_{1},s_{2})$.
2. $(C,\overline{\xi})$ is conjugated with $(C,\overline{\alpha})$ in
$\mathfrak{M}$ iff $|T(s)| = \chi_{1}(s)$.
3. $(C,\overline{\xi})$ is conjugated with $(C,\overline{\alpha})$ in
$\mathfrak{M}$ iff the osculating planes
$(P;\overline{\xi}_{1},\overline{\xi}_{2})$ coincide to the planes
$\pi(s)$ of $\mathfrak{M}$.
4. $(C,\overline{\xi})$ is conjugated with $(C,\overline{\xi})$
iff the asimptotic planes of the ruled surface
$\cal{R}(C,\overline{\xi})$ coincide with the planes $\pi(s)$ of
$\mathfrak{M}$.
\section{Versor field $(C,\overline{\xi})$ with null geodesic torsion}
\setcounter{theorem}{0}
\setcounter{equation}{0}
A new relation of conjugation of versor field $(C,\overline{\xi})$
with the tangent versor field $(C,\overline{\alpha})$ is obtained in
the case $T(s) = 0.$
\begin{definition}
The versor field $(C,\overline{\xi})$ is called orthogonal
conjugated with the tangent versor field $(C,\overline{\alpha})$ in
$\mathfrak{M}$ if its geodesic torsion $T(s) = 0,$ $\forall s\in
(s_{1},s_{2})$.
Some properties
\begin{itemize}
\item[1.] $(C,\overline{\xi})$ is orthogonal conjugated with
$(C,\overline{\alpha})$ in $\mathfrak{M}$ iff $\overline{\mu}(s)$ are
parallel with the characteristic line of planes $\pi(s)$ along the
curve $C$.
\item[2.] $(C,\overline{\xi})$ is orthogonal conjugated with
$(C,\overline{\alpha})$ in $\mathfrak{M}$ if $|K_{1}(s)| =
\chi_{1}(s)$, along $C$.
\end{itemize}
\end{definition}
\begin{theorem}
Assuming that the versor field $(C,\overline{\xi})$ in the
configuration $\mathfrak{M}(C,\overline{\xi},\pi)$ has two of the
following three properties then it has the third one, too:
\begin{itemize}
\item[a.] The osculating planes $(P; \overline{\xi}_{1},
\overline{\xi}_{2})$ are parallel in $E_{3}$ along $C$.
\item[b.] The osculating planes
$(P;\overline{\xi}_{1},\overline{\xi}_{2})$ have constant angle
with the plans $\pi(s)$ on $C$.
\item[c.] The geodesic torsion $T(s)$ vanishes on $C$.
\end{itemize}
\end{theorem}
The proof is based on the formula $T(s) = K_{2}(s) +
\displaystyle\frac{d\varphi}{ds}$, $\varphi = \sphericalangle
(\overline{\xi}_{2},\overline{\nu})$.
Consider two Myller configurations
$\mathfrak{M}(C,\overline{\xi},\pi)$ and
$\mathfrak{M}^{\prime}(C,\overline{\xi},\pi^{\prime})$ which have
in common the versor field $(C,\overline{\xi})$. Denote by $\varphi =
\sphericalangle (\overline{\xi}_{2},\overline{\nu})$, $\varphi^{\prime}
= \sphericalangle(\overline{\xi}_{2},\overline{\nu}^{\prime})$.
Then the geodesic torsions of $(C,\overline{\xi})$ in
$\mathfrak{M}$ and $\mathfrak{M}^{\prime}$ are follows:
$$
T(s) = K_{2}(s) + \displaystyle\frac{d\varphi}{ds},\;\; T^{\prime}(s) = K_{2}(s) +
\displaystyle\frac{d\varphi^{\prime}}{ds}.
$$
Evidently, we have $\varphi-\varphi^{\prime} = \sphericalangle
(\overline{\nu},\overline{\nu}^{\prime})$.
By means of these relations we can prove, without difficulties:
\begin{theorem}
If the Myller configurations $\mathfrak{M}(C,\overline{\xi},\pi)$
and $\mathfrak{M}^{\prime}(C,\overline{\xi},\pi^{\prime})$ have two
of the following properties:
\begin{itemize}
\item[a)] $(C,\overline{\xi})$ has the null geodesic torsion,
$T(s)=0$ in $\mathfrak{M}$.
\item[b)] $(C,\overline{\xi})$ has the null geodesic torsion
$T^{\prime}(s)$in $\mathfrak{M}^{\prime}$.
\item[c)] The angle
$\sphericalangle(\overline{\nu},\overline{\nu}^{\prime})$is
constant along $C$, then $\mathfrak{M}$ and
$\mathfrak{M}^{\prime}$ have the third property.
\end{itemize}
\end{theorem}
\begin{remark}\rm
The versor field $(C,\overline{\nu}_{2})$ is orthogonally conjugated
with the tangent versors $(C,\overline{\alpha})$ in the configuration
$\mathfrak{M}(C,\overline{\xi},\pi)$.
\end{remark}
\section{The vector field parallel in Myller sense in configurations $\mathfrak{M}$}
\setcounter{theorem}{0}
\setcounter{equation}{0}
Consider $(C,\overline{V})$ a vector field, along the curve $C$.
We denote $\overline{V}(s) = \overline{V}(\overline{r}(s))$ and
say that $(C,\overline{V})$ is a vector field in the configuration
$\mathfrak{M} = \mathfrak{M}(C,\overline{\xi},\pi)$ if the vector
$\overline{V}(s)$ belongs to plane $\pi(s),$ $\forall s\in
(s_{1},s_{2})$.
\begin{definition}
The vectors field $(C,\overline{V})$ in
$\mathfrak{M}(C,\overline{\xi},\pi)$ is parallel in Myller sense
if the vector field $\displaystyle\frac{d\overline{V}}{ds}$ is normal to
$\mathfrak{M}$, i.e. $\displaystyle\frac{d\overline{V}}{ds} =
\lambda(s)\overline{\nu}(s),$ $\forall s\in
(s_{1},s_{2})$.\end{definition}
The parallelism in Myller sense is a direct generalization of
Levi-Civita parallelism of tangent vector fields along a curve
$C$ of a surface $S$.
It is not difficult to prove that the vector field
$\overline{V}(s)$ is parallel in Myller sense if the vector field
$$
\overline{V}^{\prime}(s+\Delta s) = pr_{\pi(s)}\overline{V}(s+\Delta s)
$$
is parallel in ordinary sens in $E_{3}$-up to terms of second
order in $\Delta s$.
In Darboux frame, $\overline{V}(s)$ can be represented by its
coordinate as follows:
\begin{equation}
\overline{V}(s) =V^{1}(s)\overline{\xi}(s) +
V^{2}(s)\overline{\mu}(s).
\end{equation}
By virtue of fundamental equations (2.1.4) we find:
\begin{equation}
\hspace*{8mm}\displaystyle\frac{d\overline{V}}{ds} = \left(\displaystyle\frac{dV^{1}}{ds} - G
V^{2}\right)\overline{\xi} +\left(\displaystyle\frac{dV^{2}}{ds} +
GV^{1}\right)\overline{\mu} + (KV^{1} +TV^{2})\overline{\nu}.
\end{equation}
Taking into account the Definition 2.8.1, one proves:
\begin{theorem}
The vector field $\overline{V}(s)$, $(2.8.1)$ is parallel in Myller
sense in configuration $\mathfrak{M}(C, \overline{\xi}, \pi)$ iff
coordinates $V^{1}(s), V^{2}(s)$ are solutions of the system of
differential equations:
\begin{equation}
\displaystyle\frac{dV^{1}}{ds}-GV^{2} = 0,\;\; \displaystyle\frac{dV^{2}}{ds}+GV^{1} = 0.
\end{equation}
\end{theorem}
In particular, for $\overline{V}(s) = \overline{\xi}(s)$, we
obtain
\begin{theorem}
The versor field $\overline{\xi}(s)$ is parallel in Myller sense
in $\mathfrak{M}(C,\overline{\xi},\pi)$ iff the geodesic curvature
$G(s)$ of $(C,\overline{\xi})$ in $\mathfrak{M}$ vanishes.
\end{theorem}
This is a reason that Al. Myller says that $G$ is {\it the
deviation of parallelism} [31]. Later we will see that $G(s)$ is
an intrinsec invariant in the geometry of surfaces in $E_{3}$.
By means of (2.8.3) we have
\begin{theorem}
There exists an unique vector field $\overline{V}(s)$, $s\in
(s_{1}^{\prime}, s_{2}^{\prime})\subset (s_{1},s_{2})$ parallel in
Myller sense in the configuration
$\mathfrak{M}(C,\overline{\xi},\pi)$ which satisfy the initial
condition $\overline{V}(s_{0}) = \overline{V}_{0}$, $s_{0}\in
(s_{1}^{\prime}, s_{2}^{\prime})$ and $\langle \overline{V}_{0},
\overline{\nu}(s_{0}) \rangle = 0.$
\end{theorem}
Evidently, theorem of existence and uniqueness of solutions of
system (2.8.3), is applied in this case.
In particular, if $G(s)$ = constant, then the general solutions of
(2.8.3) can be obtained by algebric operations.
An important property of parallelism in Myller sense is expressed
in the next theorem.
\begin{theorem}
The Myller parallelism of vectors in $\mathfrak{M}$ preserves the
lengths and angles of vectors.
\end{theorem}
\begin{proof}\rm If $\displaystyle\frac{d\overline{V}}{ds} =
\lambda(s)\overline{\nu}$, then $\displaystyle\frac{d}{ds}\langle \overline{V},
\overline{V} \rangle=0.$ Also,
$\displaystyle\frac{d\overline{V}^{\prime}}{ds}=\lambda(s)\overline{\nu}$,
$\displaystyle\frac{d\overline{U}}{ds} = \lambda^{\prime}(s)\overline{\nu}$, then
$\displaystyle\frac{d}{ds}\langle \overline{V}(s),
\overline{U}^{\prime}(s)\rangle = 0$
\end{proof}
\section{Adjoint point, adjoint curve and concurrence in Myller sense}
\setcounter{theorem}{0}
\setcounter{equation}{0}
The notions of adjoint point, adjoint curve and concurrence in
Myller sense in a configuration $\mathfrak{M}$ have been
introduced and studied by O. Mayer [20] and Gh. Gheorghiev
[11], [55]. They applied these notions, to the theory of surfaces,
nonholomorphic manifolds and in the geometry of versor fields in
Euclidean space $E_{3}$.
In the present book we introduce these notions in a different way.
Consider the vector field
$$
\overline{\xi}^{*}(s+\Delta s) = pr_{\pi(s)}\overline{\xi}(s+\Delta s).
$$
Taking into account the formula (2.2.1)$^{\prime}$ we can write up
to terms of second order in $\Delta s:$
\begin{equation}
\overline{\xi}^{*}(s+\Delta s) = \overline{\xi}(s) + \Delta s (G
\overline{\mu}(s) + \overline{\omega}^{*}(s,\Delta s))
\end{equation}
with
$$
\overline{\omega}^{*}(s, \Delta s) \to 0, (\Delta s\to 0).
$$
Let $C^{\prime}$ be the orthogonal projection of the curve $C$ on
the plane $\pi(s)$. A neighbor point $P^{\prime}(s+\Delta s)$ is
projected on plane $\pi(s)$ in the point $P^{*}(s+\Delta s)$ given by
\begin{equation}
\begin{array}{l}
\overline{r}^{*}(s+\Delta s) = \overline{r}(s) + \Delta s
(c_{1}\overline{\xi} + c_{2}\overline{\mu} +
\overline{\omega}_{0}^{*}(s, \Delta s)),\\ \overline{\omega}_{0}^{*}(s, \Delta s)
\to 0, (\Delta s\to 0).
\end{array}
\end{equation}
\begin{definition}
The adjoint point of the point $P(s)$ with respect to
$\overline{\xi}(s)$ in $\mathfrak{M}$ is the characteristic point
$P_{a}$ on the line $(P;\overline{\xi})$ of the plane ruled
surface $R(C^{*}, \overline{\xi}^{*})$.
\end{definition}
One proves that the position vector $\overline{R}(s)$ of adjoint
point $P_{a}$ for $G\neq 0$, is as follows:
\begin{equation}
\overline{R}(s) = \overline{r}(s) -
\displaystyle\frac{c_{2}}{G}\overline{\xi}(s).
\end{equation}
The vector field $(C^{*}, \overline{\xi}^{*})$ from (2.9.3) is
called {\it geodesic field}. A result established by O. Mayer [20]
holds:
\begin{theorem}
If the versor field $(C, \overline{\xi})$ is enveloping in space
$E_{3}$, then the adjoint point $P_{a}$ of the point $P(s)$ in
$\mathfrak{M}$ is the contact point of the line $(P,
\overline{\xi})$ with the cuspidale line.
\end{theorem}
\begin{definition}
The geometric locus of the adjoint points corresponding to the
versor field $(C, \overline{\xi})$ in $\mathfrak{M}$ is the
adjoint curve $C_{a}$ of the curve $C$ in $\mathfrak{M}$.
\end{definition}
The adjoint curve $C_{a}$ has the vector equations (2.9.3) for
$\forall s\in (s_{1},s_{2})$.
Now, we can introduce
\begin{definition}
The versor field $(C, \overline{\xi})$ is concurrent in Myller
sense in $\mathfrak{M}(C, \overline{\xi}, \pi)$ if, at every point
$P(s)\in C$ the geodesic vector field $(C^{*},
\overline{\xi}^{*})$ is concurrent.
\end{definition}
For $G(s)\neq 0$, we have
\begin{theorem}
The versor field $(C,\xi)$ is concurrent in Myller sense in
$\mathfrak{M}$ iff the following equation hold
\begin{equation}
\displaystyle\frac{d}{ds}\left(\displaystyle\frac{c_{2}}{G}\right) = c_{1}.
\end{equation}
\end{theorem}
For the proof see Section 1.1, Chapter 1, Theorem 1.1.3.
\section{Spherical image of a configuration $\mathfrak{M}$}
\setcounter{theorem}{0}
\setcounter{equation}{0}
In the Section 2, Chapter 1 we defined the spherical image of a
versor field $(C,\overline{\xi})$. Applying this idea to the
normal vectors field $(C,\overline{\nu})$ to a Myller
configuration $\mathfrak{M}(C, \overline{\xi}, \pi)$ we define the
notion of spherical image $C^{*}$ of $\mathfrak{M}$ as being
\begin{equation}
\overline{\nu}^{*}(s) = \overline{OP}^{*}(s) = \overline{\nu}(s)
\end{equation}
Thus, the relations between the curvature $\chi_{1}$ and torsion
$\chi_{2}$ of $(C,\overline{\nu})$ and geodesic curvature
$\kappa_{g}^{*}$ of $C^{*}$ at a point $P^{*}\in C^{*}$ and arclength
$s^{*}$ are as follows
$$
\chi_{1} = \displaystyle\frac{ds^{*}}{ds},\; \chi_{2}ds = \kappa_{g}^{*}ds^{*}.
$$
The properties enumerated in Section 2, ch 1, can be obtained for the
spheric image $C^{*}$ of configuration $\mathfrak{M}$.
Consider the versor field $\overrightarrow{P^{*}P^{*}_{1}} =
\xi^{*}(s) = \xi(s)$ and the angle $\theta = \sphericalangle
(\overline{\nu}_{2}, \overline{\xi})$. It follows
\begin{equation}
K = -\displaystyle\frac{ds^{*}}{ds}\cos \theta,\;\; (for \chi_{2}\neq 0).
\end{equation}
Thus, for $\theta = \pm \displaystyle\frac{\pi}{2}$ one gets $K=0,$ which leads to
a new interpretation of the fact that $\overline{\xi}(s)$ is
conjugated to $\overline{\alpha}(s)$ in Myller configurations.
Analogous one can obtain $T = -\displaystyle\frac{ds^{*}}{ds}\cos
\widetilde{\theta},$ $\widetilde{\theta} =
\sphericalangle(\overline{\nu}_{2}, \overline{\mu}^{*})$ (with
$\overline{\mu}^{*}(s) = \overline{\mu}(s)$, applied at the point
$P^{*}(s)\in C^{*}$). For $\widetilde{\theta} = \pm \displaystyle\frac{\pi}{2}$ it
follows that $\overline{\xi}(s)$ are orthogonally conjugated with
$\overline{\alpha}(s) .$
The problem is to see if the Gauss-Bomet formula can be extended
to Myller configurations $\mathfrak{M}(C, \overline{\xi}, \pi)$.
In the case $(\alpha(s)\in \pi(s)$, (i.e. $\overline{\alpha}(s)\bot
\overline{\nu}(s)$) such a problem was suggested by Thomson [18] and it
has been solved by Mark Krein in 1926, [18].
Here, we study this problem in the general case of Myller
configuration, when $\langle \overline{\alpha}(s),
\overline{\nu}(s)\rangle\neq 0.$
First of all we prove
\begin{lemma} Assume that we have:
1. $\mathfrak{M}(C, \overline{\xi},\pi)$ a Myller configuration of
class $C^{3}$, in which $C$ is a closed curve, having $s$ as
arclength.
2. The spherical image $C^{*}$ of $\mathfrak{M}$ determines on the
support sphere $\Sigma$ a simply connected domain of area $\omega$.
In this hypothesis we have the formula
\begin{equation}
\omega = 2\pi - \int_{C}\left(\overline{\nu}, \displaystyle\frac{d\overline{\nu}}{ds},
\displaystyle\frac{d^{2}\overline{\nu}}{ds^{2}}\right)/
\left\|\displaystyle\frac{d\overline{\nu}}{ds}\right\|^{2}ds.
\end{equation}
\end{lemma}
\begin{proof}\rm Let $\Sigma$ be the unitary sphere of
center $O\in E_{3}$ and a simply connected domain $D$, delimited
by $C^{*}$ on $\Sigma.$ Assume that $D$ remains to left with
respect to an observer looking in the sense of versor
$\overline{\nu}(s)$, when he is going along $C^{*}$ in the
positive sense.
Thus, we can take the following representation of $\Sigma$:
\begin{eqnarray}
x^{1}&=&\cos \varphi \sin \theta\nonumber\\
x^{2}&=&\sin \varphi \sin \theta\\x^{3}&=&\cos \varphi,\;\; \varphi\in [0, 2\pi), \theta
\in \(-\displaystyle\frac{\pi}{2}, \displaystyle\frac{\pi}{2}\).\nonumber
\end{eqnarray}
The curve $C^{*}$ can be given by
\begin{equation}
\varphi = \varphi(s),\; \theta = \theta(s),\;\; s\in [0,s_{1}]
\end{equation}
with $\varphi(s), \theta(s)$ of class $C^{3}$ and $C^{*}$ being closed:
$\varphi(0) = \varphi(s_{1})$, $\theta(0) = \theta(s_1)$.
The area $\omega$ of the domain $D$ is
\begin{equation}
\begin{array}{l}
\omega = \int_{0}^{s_{1}}\int_{0}^{\theta(s_{1})}\sin \theta d\theta =
\int_{0}^{s_{1}}(1-\cos \theta)\displaystyle\frac{d\varphi}{ds}ds =\\ = 2\pi
-\int_{0}^{s_{1}}\cos \theta \displaystyle\frac{d\varphi}{ds}ds.
\end{array}
\end{equation}
Noticing that the versor $\overline{\nu}^{*} = \overline{\nu}(s)$
has the coordinate (2.10.4) a straightforward calculus leads to
$$
\left\langle \overline{\nu}, \displaystyle\frac{d\overline{\nu}}{ds},
\displaystyle\frac{d^{2}\overline{\nu}}{ds^{2}} \right\rangle/
\left\|\displaystyle\frac{d\overline{\nu}}{ds}\right\|^{2} = \cos \theta
\displaystyle\frac{d\varphi}{ds} +\displaystyle\frac{d}{ds} \mbox{arctg}\, \displaystyle\frac{\sin
\theta\displaystyle\frac{d\varphi}{ds}}{\displaystyle\frac{d\theta}{ds}}.
$$
Denoting by $\psi$ the angle between the meridian $\varphi = \varphi_{0}$ and
curve $C^{*}$, oriented with respect to the versor
$\overline{\nu}$ we have
\begin{equation}
\mbox{tg} \psi =\(\sin \theta \displaystyle\frac{d\varphi}{ds}\)/ \displaystyle\frac{d\theta}{ds}.
\end{equation}
The previous formulae lead to
\begin{equation}
\left\langle \overline{\nu}, \displaystyle\frac{d\overline{\nu}}{ds},
\displaystyle\frac{d^{2}\overline{\nu}}{ds^{2}} \right\rangle/
\left\|\displaystyle\frac{d\overline{\nu}}{ds}\right\|^{2} = \cos \theta
\displaystyle\frac{d\varphi}{ds} + \displaystyle\frac{d\psi}{ds}.
\end{equation}
But in our conditions of regularity $\displaystyle\int_{C}\displaystyle\frac{d\psi}{ds}ds =
0.$ Thus (2.10.7), (2.10.8) implies the formula (2.10.3)
\end{proof}
It is not difficult to see that the formula (2.10.3) can be
generalized in the case when the curve $C$ of the configuration
$\mathfrak{M}$ has a finite number of angular points. The second
member of the formula (2.10.3) will be additive modified with the
total of variations of angle $\psi$ at the angular points
corresponding to the curve $C$.
Now, one can prove the generalization of Mark Krein formula.
\begin{theorem}
Assume that we have
1. $\mathfrak{M}(C, \overline{\xi}, \pi)$ a Myller configuration
of class $C^{3}$ $($i.e. $C$ is of class $C^{3}$ and $\xi(s)$,
$\overline{\nu}(s)$ are the class $C^{2}$$)$ in which $C$ is a
closed curve, having $s$ as natural parameter.
2. The spherical image $C^{*}$ of $\mathfrak{M}$ determine on the
support sphere $\Sigma$ a simply connected domain of area $\omega.$
3. $\sigma$ the oriented angle between the versors
$\overline{\nu}_{3}(s)$ and $\overline{\xi}(s)$. In these
conditions the following formula hold:
\begin{equation}
\omega = 2\pi -\int_{C}G(s)ds +\int_{C}d\sigma.
\end{equation}
\end{theorem}
\noindent {\bf Proof.} The first two conditions allow to apply the
Lemma 2.10.1. The fundamental equations (1.3.1), (1.3.2) of
$(C,\overline{\nu})$ give us for $\overline{\nu} =
\overline{\nu}_{1}$:
$$
\displaystyle\frac{d\overline{\nu}_{1}}{ds} = \chi_{1}\overline{\nu}_{2},\;
\displaystyle\frac{d_{2}\overline{\nu}_{1}}{ds^{2}} =
\displaystyle\frac{d\chi_{1}}{ds}\overline{\nu}_{2} +
\chi_{1}(-\chi_{1}\overline{\nu}_{1} +
\chi_{2}\overline{\nu}_{3}).
$$
So,
$$
\left\langle \overline{\nu}_{1}, \displaystyle\frac{d\overline{\nu}_{1}}{ds},
\displaystyle\frac{d^{2}\overline{\nu}_{1}}{ds^{2}} \right\rangle =
\chi_{1}^{2}\chi_{2}.
$$
Thus, the formula (2.10.3), leads to the following formula
$$
\omega = 2\pi -\int_{C}\chi_{2}(s)ds.
$$
But, we have $G(s) =\chi_{2}(s) +\displaystyle\frac{d\sigma}{ds}$, $G(s)$ being
the geodesic curvature of $(C,\overline{\xi})$ in $\mathfrak{M}$.
Then the last formula is exactly (2.10.9).
If $G=0$ for $\mathfrak{M}$, then we have $\omega = 2\pi. $
Indeed $G(s) = 0$ along the curve $C$ imply $\omega = 2\pi +2k \pi$,
$k\in \mathbb{N}$. But we have $0\leq \omega <4\pi,$ so $k=0.$
A particular case of Theorem 2.10.1 is the famous result of Jacobi:
{\it The area $\omega$ of the domain $D$ determined on the sphere $\Sigma$
by the closed curve $C^{*}$-spherical image of the principal normals
of a closed curve $C$ in $E_{3}$, assuming $D$ a simply connected
domain, is a half of area of sphere $\Sigma$.}
In this case we consider the Myller configuration $\mathfrak{M}(C,
\overline{\alpha}, \pi)$, $\pi(s)$ being the rectifying planes of $C$.
\newpage
\thispagestyle{empty}
\chapter{Applications of theory of Myller configurations in the geometry of nonholonomic manifolds from $E_{3}$}
The second efficient applications of the theory of Myller
configurations can be done in the geometry of nonholonomic
manifolds $E_{3}^{2}$ in $E_{3}$. Some important results, obtained
in the geometry of manifolds $E_{3}^{2}$ by Issaly l$'$Abee, D.
Sintzov [40], Gh. Vr\u{a}nceanu [44], [45], Gr. Moisil [72], M. Haimovici [14], [15] Gh. Gheorghiev [10], [11], I. Popa [12],
G. Th. Gheorghiu [13], R. Miron [30], [64], [65], I. Creang\u{a} [4], [5], [6], A. Dobrescu [7], [8] and I. Vaisman [41], [42], can be unitary presented by means of
associated Myller configuration to a curve embedded in
$E_{3}^{2}$. Now, a number of new concepts appears, the mean
torsion, the total torsion, concurrence of tangent vector field.
The new formulae, as Mayer, Bortolotti-formulas, Dupin and Bonnet
indicatrices etc. will be studied, too.
\section{Moving frame in Euclidean spaces $E_{3}$}
\setcounter{theorem}{0}\setcounter{equation}{0}
Let $R = (P; I_{1}, I_{2}, I_{3})$ be an orthonormed positively
oriented frame in $E_{3}$. The application $P\in E_{3}\to R = (P;
I_{1}, I_{2}, I_{3})$ of class $C^{k}$, $k\geq 3$ is a moving frame
of class $C^{k}$ in $E_{3}$. If $\overline{r} =
\overrightarrow{OP} = x \overline{e}_{1} + y \overline{e}_{2} +
z\overline{e}_{3}$ is the vector of position of point $P$, and $P$
is the application point of the versors $I_{1}, I_{2}, I_{3}$,
thus the moving equation of $R$ can be expressed in the form (see
Spivak, vol. I [76] and Biujguhens [2]):
\begin{equation}
d\overline{r} = \omega_{1}I_{1} + \omega_{2}I_{2} + \omega_{3}I_{3},
\end{equation}
where $\omega_{i}(i=1,2,3)$ are independent 1 forms of class $C^{k-1}$
on $E_{3}$, and
\begin{equation}
dI_{i} = \sum_{j=1}^{3}\omega_{ij}I_{j}, \;\; (i=1,2,3),
\end{equation}
with $\omega_{ij}$, $(i,j = 1,2,3)$ are the rotation Ricci
coefficients of the frame $R$. They are 1-forms of class
$C^{k-1}$, satisfying the skewsymmetric conditions
\begin{equation}
\omega_{ij} +\omega_{ji} = 0,\;\; (i,j = 1,2,3).
\end{equation}
In the following, it is convenient to write the equations (5.1.2)
in the form
$$
\begin{array}{c}
dI_{1} = rI_{2} -qI_{3}\\dI_{2} = p I_{3} - r I_{1}\\dI_{3} = q
I_{1} - p I_{2}.
\end{array}\leqno{(5.1.2)^{\prime}}
$$
Thus, {\it thus structure equations} of the moving frame $R$ can
be obtained by exterior differentiating the equations (5.1.1) and
(5.1.2)$^{\prime}$ modulo the same system of equations (5.1.1),
(5.1.2)$^{\prime}$.
One obtains, without difficulties
\begin{theorem}
The structure equations of the moving frame $R$ are
\begin{eqnarray}
&&d\omega_{1} = r \wedge \omega_{2},-q\wedge \omega_{3}, \;\; dp = r\wedge
q,\nonumber\\&&d\omega_{2} = p\wedge \omega_{3} - r \wedge \omega_{1},\;\; dq
= p\wedge r ,\\&&d\omega_{3} = q\wedge \omega_{1} - p\wedge \omega_{2},\;\; dr
= q\wedge p.\nonumber
\end{eqnarray}
\end{theorem}
In the vol. II of the book of Spivak [76], it is proved the theorem
of existence and uniqueness of the moving frames:
\begin{theorem}
Let $(p,q,r)$ be 1-forms of class $C^{k-1}$ ,$(k\geq 3)$ on $E_{3}$
which satisfy the second group of structure equation $(5.1.4)$,
thus:
$1^{\circ}.$ In a neighborhood of point $O\in E_{3}$ there is a
triple of vector fields $(I_{1}, I_{2}, I_{3})$, solutions of
equations (5.1.2)$^{\prime}$, which satisfy initial conditions
$(I_{1}(0), I_{2}(0), I_{3}(0))$-positively oriented, orthonormed
triple in $E_{3}$.
$2^{\circ}.$ In a neighborhood of point $O\in E_{3}$ there exists
a moving frame $R = (P; I_{1}, I_{2}, I_{3})$ orthonormed
positively oriented for which $\overline{r} = \overrightarrow{OP}$
is given by (5.1.1), where $\omega_{i} (i=1,2,3)$ satisfy the first
group of equations (5.1.4). $(I_{1}, I_{2}, I_{3})$ are given in
$1^{\circ}$ and the initial conditions are verified: $(P_{0} =
\overrightarrow{OO}; I_{1}(0), I_{2}(0), I_{3}(0))$-the
orthonormed frame at point $O\in E_{3}$.\end{theorem}
Remarking that the 1-forms $\omega_{1}, \omega_{2}, \omega_{3}$ are
independent we can express 1-forms $p,q,r$ with respect to
$\omega_{1}, \omega_{2}, \omega_{3}$ in the following form:
\begin{eqnarray}
&&p = p_{1}\omega_{1} + p_{2}\omega_{2} + p_{3}\omega_{3},\ q = q_{1}\omega_{1} +
q_{2}\omega_{2} + q_{3}\omega_{3},\\&& r = r_{1}\omega_{1} + r_{2}\omega_{2} +
r_{3}\omega_{3}.\nonumber
\end{eqnarray}
The coefficients $p_{i}, q_{i}, r_{i}$ are function of class
$C^{k-1}$ on $E_{3}$.
Of course we can write the structure equations (5.1.4) in terms of
these coefficients. Also, we can state the Theorem 5.1.2 by means of
coefficients of 1-forms $p,q,r$.
The 1-forms $p,q,r$ determine the rotation vector $\Omega$ of the
moving frame:
\begin{equation}
\Omega = p I_{1} + q I_{2} + r I_{3}.
\end{equation}
Its orthogonal projection on the plane $(P; I_{1}, I_{2})$ is
\begin{equation}
\overline{\theta} = p I_{1} + q I_{2}.
\end{equation}
Let $C$ be a curve of class $C^{k}$, $k\geq 3$ in $E_{3}$, given
by
\begin{equation}
\overline{r} = \overline{r}(s), \;\; s\in(s_{1},s_{2}),
\end{equation}
where $s$ is natural parameter. If the origin $P$ of moving frame
describes the curve $C$ we have $R = (P; I_{1}(s), I_{2}(s),
I_{3}(s))$, where $P(s) = P(\overline{r}(s))$, $I_{j}(s) =
I_{j}(\overline{r}(s))$, $(j=1,2,3)$.
So, the tangent versor $\displaystyle\frac{d\overline{r}}{ds}$ to curve $C$ is:
\begin{equation}
\displaystyle\frac{d\overline{r}}{ds} = \displaystyle\frac{\omega_{1}(s)}{ds}I_{1} +
\displaystyle\frac{\omega_{2}(s)}{ds}I_{2} + \displaystyle\frac{\omega_{3}(s)}{ds}I_{3}
\end{equation}
and $ds^{2}=\langle d\overline{r},d\overline{r}\rangle$ is:
\begin{equation}
ds^{2} = (\omega_{1}(s))^{2} + (\omega_{2}(s))^{2} + (\omega_{3}(s))^{2},
\end{equation}
where $\omega_{i}(s)$ are 1-forms $\omega_{i}$ restricted to $C$.
The restrictions $p(s), q(s), r(s)$ of the 1-forms $p,q,r$ to $C$
give us:
\begin{equation}
\begin{array}{lll}
\displaystyle\frac{p(s)}{ds} &=& p_{1}(s)\displaystyle\frac{\omega_{1}(s)}{ds} +
p_{2}(s)\displaystyle\frac{\omega_{2}(s)}{ds}
+p_{3}(s)\displaystyle\frac{\omega_{3}(s)}{ds},\\\vspace*{-.3cm}\\ \displaystyle\frac{q(s)}{ds} &=&
q_{1}(s)\displaystyle\frac{\omega_{1}(s)}{ds} + q_{2}(s)\displaystyle\frac{\omega_{2}(s)}{ds} +
q_{3}(s)\displaystyle\frac{\omega_{3}(s)}{ds},\\\vspace*{-.3cm}\\ \displaystyle\frac{r(s)}{ds}&=&r_{1}(s)\displaystyle\frac{\omega_{1}(s)}{ds}
+ r_{2}(s)\displaystyle\frac{\omega_{2}(s)}{ds} +
r_{3}(s)\displaystyle\frac{\omega_{3}(s)}{ds}.
\end{array}
\end{equation}
\section{Nonholonomic manifolds $E_{3}^{2}$}
\setcounter{theorem}{0}\setcounter{equation}{0}
\begin{definition}
A nonholonomic manifold $E_{3}^{2}$ on a domain $D\subset E_{3}$
is a nonintegrable distribution $\cal{D}$ of dimension 2, of class
$C^{k-1},$ $k\geq 3$.
\end{definition}
One can consider $\cal{D}$ as a plane field $\pi(P)$, $P\in D,$ the
application $P\to \pi(P)$ being of class $C^{k-1}$.
Also $\cal{D}$ can be presented as the plane field $\pi(P)$
orthogonal to a versor field $\overline{\nu}(P)$ normal to the
plane $\pi(P)$, $\forall P\in D.$
Assuming that $\overline{\nu}(P)$ is the versor of vector
$$\overline{V}(P) = X(x,y,z)\overline{e}_{1} +
Y(x,y,z)\overline{e}_{2} + Z(x,y,z)\overline{e}_{3}$$ and
$\overrightarrow{OP} = x \overline{e}_{1} + y \overline{e}_{2} +z
\overline{e}_{3}$ thus the nonholonomic manifold $E_{3}^{2}$ is
given by the Pfaff equation:
\begin{equation}
\omega = X(x,y,z)dx + Y(x,y,z)dy + Z(x,y,z)dz = 0
\end{equation}
which is nonintegrable, i.e.
\begin{equation}
\omega\wedge d\omega \neq0.
\end{equation}
Consider a moving frame $\cal{R} = (P; I_{1}, I_{2}, I_{3})$ in
the space $E_{3}$ and the nonholonomic manifold $E_{3}^{2}$, on a
domain $D$ orthogonal to the versors field $I_{3}$. It is given by
the Pfaff equations
\begin{equation}
\omega_{3} = \langle I_{3}, d\overline{r}\rangle = 0,\;\;\;
\mbox{on}\; D.
\end{equation}
By means of (5.1.4) we obtain
\begin{equation}
\omega_{3}\wedge d\omega_{3} = -(p_{1} +q_{2})\omega_{1}\wedge \omega_{2}\wedge
\omega_{3}.
\end{equation}
So, the Pfaff equation $\omega_{3} = 0$ is not integrable if and only
if we have
\begin{equation}
p_{1} + q_{2}\neq 0,\; \mbox{on}\; D.
\end{equation}
It is very known that:
In the case $p_{1} + q_{2} =0$ on the domain $D$, there are two
functions $h(x,y,z)\neq 0$ and $f(x,y,z)$ on $D$ with the property
\begin{equation}
h\omega_{3} = df.
\end{equation}
Thus the equation $h\omega_{3} = 0,$ have a general solution
\begin{equation}
f(x,y,z) = c,\;\;(c=const.),\;\; P(x,y,z)\in D.
\end{equation}
The manifold $E_{3}^{2}$, in this case is formed by a family of
surfaces, given by (5.2.7).
For simplicity we assume that the class of the manifold
$E_{3}^{2}$ is $C^{\infty}$ and the same property have the
geometric object fields or mappings which will be used in the
following parts of this chapter.
A smooth curve $C$ embedded in the nonholonomic manifolds
$E_{3}^{2}$, has the tangent vector $\overline{\alpha} =
\displaystyle\frac{d\overline{r}}{ds}$ given by (5.1.9) and $\omega_{3} = 0.$
This is
\begin{equation}
\overline{\alpha}(s) = \displaystyle\frac{d\overline{r}}{ds}
=\displaystyle\frac{\omega_{1}(s)}{ds}I_{1} + \displaystyle\frac{\omega_{2}(s)}{ds}I_{2},\; \forall
P(\overline{r}(s))\in D.
\end{equation}
The square of arclength element, $ds$ is
\begin{equation}
ds^{2} = (\omega_{1}(s))^{2} + (\omega_{2}(s))^{2}.
\end{equation}
And the arclength of curve $C$ on the interval $[a,b]\subset
(s_{1},s_{2})$ is \begin{equation} s =
\int_{a}^{b}\sqrt{(\omega_{1}(\sigma))^{2} + (\omega_{2}(\sigma))^{2}}d\sigma.
\end{equation}
A tangent versor field $\overline{\xi}(s)$ at the same point
$P(s)$ to another curve $C^{\prime},$ $P\in C^{\prime}$ and $P\in
C,$ can be given in the same form (5.2.8):
\begin{equation}
\overline{\xi}(s) = \displaystyle\frac{\delta\overline{r}}{\delta s} =
\displaystyle\frac{\widetilde{\omega}_{1}(s)}{\delta s}I_{1} +
\displaystyle\frac{\widetilde{\omega}_{2}(s)}{\delta s}I_{2},
\end{equation}
with $\delta s^{2} = (\widetilde{\omega}_{1}(s))^{2} +
(\widetilde{\omega}_{2}(s))^{2}$.
Thus, $\overline{\xi}(s)$ is a versor field tangent to the
nonholonomic manifold $E_{3}^{2}$ along to the curve $C$.
To any tangent vector field $(C,\overline{\xi})$ to $E_{3}^{2}$
along the curve $C$ belonging to $E_{3}^{2}$ we uniquely
associated the tangent Myller configuration $\mathfrak{M}_{t}(C,
\overline{\xi}, I_{3})$. We say that: {\it the geometry of the
associated tangent Myller configuration $\mathfrak{M}_{t}(C,
\overline{\xi},I_{3})$ is the geometry of versor field $(C,
\overline{\xi})$ on $E_{3}^{2}$}. In particular the geometry of
configuration $\mathfrak{M}_{t}(C,\overline{\alpha},I_{3})$ is {\it
the geometry of curve $C$} in the nonholonomic manifold
$E_{3}^{2}$.
The Darboux frame of $(C,\overline{\xi})$ on $E_{3}^{2}$ is
$\cal{R}_{D} = (P; \overline{\xi}, \overline{\mu}, I_{3})$, with
$\overline{\mu} = I_{3}\times \overline{\xi}$, i.e.:
\begin{equation}
\overline{\mu} = \displaystyle\frac{\widetilde{\omega}_{1}(s)}{\delta s}I_{2} -
\displaystyle\frac{\widetilde{\omega}_{2}(s)}{\delta s}I_{1}.
\end{equation}
In Darboux frame $\cal{R}_{D}$, the tangent versor $\overline{\alpha}
= \displaystyle\frac{d\overline{r}}{ds}$ to $C$ can be expressed in the
following form:
\begin{equation}
\overline{\alpha}(s) = \cos \lambda(s) \overline{\xi}(s) + \sin
\lambda(s)\overline{\mu}(s).
\end{equation}
Taking into account the relations (5.2.8), (5.2.11) and (5.2.13),
one gets:
\begin{eqnarray}
\displaystyle\frac{\omega_{1}}{ds}& =& \cos \lambda \displaystyle\frac{\widetilde{\omega}_{1}}{\delta s} -
\sin
\lambda \displaystyle\frac{\widetilde{\omega}_{2}}{\delta s},\nonumber\\
\displaystyle\frac{\omega_{2}}{ds}&=& \sin \lambda \displaystyle\frac{\widetilde{\omega}_{1}}{\delta s} +
\cos \lambda \displaystyle\frac{\widetilde{\omega}_{2}}{\delta s}.
\end{eqnarray}
For $\lambda(s) = 0$, we have $\overline{\alpha}(s) = \overline{\xi}(s)$.
\section{The invariants $G,R,T$ of a tangent versor field $(C,\overline{\xi})$ in $E_{3}^{2}$}
\setcounter{theorem}{0}\setcounter{equation}{0}
The invariants $G,K,T$ of $(C,\overline{\xi})$ satisfy the
fundamental equations
\begin{equation}
\displaystyle\frac{d\overline{\xi}}{ds} = G \overline{\mu}+K I_{3},\
\displaystyle\frac{d\overline{\mu}}{ds} = -G \overline{\xi}+T I_{3},\
\displaystyle\frac{dI_{3}}{ds} = -K \overline{\xi} - T \overline{\mu}.
\end{equation}
Consequently,
\begin{eqnarray}
G(\delta,d) &=& \left\langle \overline{\xi}, \displaystyle\frac{d\overline{\xi}}{ds},
I_{3} \right\rangle\nonumber\\
K(\delta,d)&=& \left\langle\displaystyle\frac{d\overline{\xi}}{ds}, I_{3}\right\rangle =
-\left\langle \overline{\xi}, \displaystyle\frac{dI_{3}}{ds}\right\rangle\\T(\delta,d)&=&
\left\langle \overline{\xi}, I_{3}, \displaystyle\frac{dI_{3}}{ds}\right\rangle.\nonumber
\end{eqnarray}
The proofs and the geometrical meanings are similar to those from
Chapter 3.
By means of expression (5.2.11) of versors $\overline{\xi}(s)$ we
deduce:
\begin{eqnarray}
\displaystyle\frac{d\overline{\xi}}{ds}&=& \left[
\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\widetilde{\omega}_{1}}{\delta s}\right) -
\displaystyle\frac{r}{ds}\displaystyle\frac{\widetilde{\omega}_{2}}{\delta s} \right]I_{1} + \left[
\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\widetilde{\omega}_{2}}{\delta s}\right) +
r\displaystyle\frac{\widetilde{\omega}_{1}}{\delta s} \right]I_{2} + \nonumber\\&&+
\left[ \displaystyle\frac{p}{ds}\displaystyle\frac{\widetilde{\omega}_{2}}{\delta s} - \displaystyle\frac{q}{ds}
\displaystyle\frac{\widetilde{\omega}_{1}}{\delta s} \right]I_{3},
\end{eqnarray}
and
\begin{equation}
\displaystyle\frac{dI_{3}}{ds} = \left(q_{1}\displaystyle\frac{\omega_{1}}{ds} +
q_{2}\displaystyle\frac{\omega_{2}}{ds}\right)I_{1} - \left(p_{1}\displaystyle\frac{\omega_{1}}{ds}
+ p_{2}\displaystyle\frac{\omega_{2}}{ds} \right)I_{2}.
\end{equation}
The formulae (5.3.2) lead to the following expressions for {\it
geodesic curvature} $G(\delta,d)$, {\it normal curvature} $K(\delta,d)$
and {\it geodesic torsion} $T(\delta,d)$ of the tangent versor field
$(C,\overline{\xi})$ on the nonholonomic manifold $E_{3}^{2}$:
\begin{equation}
\hspace*{8mm}G(\delta,d) = \displaystyle\frac{\widetilde{\omega}_{1}}{\delta
s}\left[\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\widetilde{\omega}_{2}}{\delta s }\right)
+ \displaystyle\frac{r}{ds}\displaystyle\frac{\widetilde{\omega}_{1}}{\delta s} \right] -
\displaystyle\frac{\widetilde{\omega}_{2}}{\delta
s}\left[\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\widetilde{\omega}_{1}}{\delta s } -
\displaystyle\frac{r}{ds}\displaystyle\frac{\widetilde{\omega}_{2}}{\delta s} \right)\right],
\end{equation}
\begin{equation}
K(\delta,d) = \displaystyle\frac{p\widetilde{\omega}_{2} - q\widetilde{\omega}_{1}}{\delta
s\cdot d s}
\end{equation}
\begin{equation}
T(\delta,d) = \displaystyle\frac{p\widetilde{\omega}_{1} + q\widetilde{\omega}_{2}}{\delta s
ds}.
\end{equation}
All these formulae take very simple forms if we consider the
angles $\alpha = \sphericalangle (\overline{\alpha},I_{1})$, $\beta =
\sphericalangle (\overline{\xi},I_{1})$ since we have
\begin{eqnarray}
&&\displaystyle\frac{\omega_{1}}{ds} = \cos \alpha,\ \displaystyle\frac{\omega_{2}}{ds} = \sin
\alpha,\nonumber\\&& \displaystyle\frac{\widetilde{\omega}_{1}}{\delta s} = \cos \beta,\
\displaystyle\frac{\widetilde{\omega}_{2}}{\delta s} = \sin \beta.
\end{eqnarray}
With notations
\begin{equation}
\hspace*{8mm}G(\delta,d) = G(\beta,\alpha); K(\delta,d) = K(\beta,\alpha), T(\delta,d) = T(\beta,\alpha)
\end{equation}
the formulae (5.3.2) can be written:
\begin{equation}
\hspace*{12mm}G(\beta,\alpha) = \displaystyle\frac{d\overline{\beta}}{ds} +r_{1}\cos \alpha +r_{2}\sin \alpha,
\end{equation}
\begin{equation}
\hspace*{14mm}K(\beta,\alpha) = (p_{1}\cos \alpha + p_{2}\sin \alpha)\sin \beta - (q_{1}\cos \alpha
+q_{2}\sin \alpha)\cos \beta,
\end{equation}
\begin{equation}
\hspace*{14mm}T(\beta,\alpha) = (p_{1}\cos \alpha +p_{2}\sin \alpha)\cos \beta + (q_{1}\cos\alpha +
q_{2}\sin \alpha)\sin \beta.
\end{equation}
Immediate consequences:
Setting
\begin{equation}
T_{m} = p_{1} +q_{2},
\end{equation}
(called the {\it mean torsion} of $E_{3}^{2}$ at point $P\in
E_{3}^{2}$),
\begin{equation}
H = p_{2} - q_{1}
\end{equation}
(called the {\it mean curvature} of $E_{3}^{2}$ at point $P\in
E_{3}^{2}$), from (5.3.11) and (5.3.12) we have:
\begin{eqnarray}
K(\beta,\alpha) - K(\alpha,\beta)&=& T_{m}\cos (\alpha-\beta),\nonumber\\T(\beta,\alpha)-
T(\alpha,\beta) &=& H \cos (\alpha-\beta).
\end{eqnarray}
\begin{theorem}
The following properties hold:
1. Assuming $\beta-\alpha\neq \pm \displaystyle\frac{\pi}{2}$, the normal curvature
$K(\beta,\alpha)$ is symmetric with respect to the versor field
$(C,\overline{\xi}), (C,\overline{\alpha})$ at every point $P\in
E_{3}^{2}$, if and only if the mean torsion $T_{m}$ of $E_{3}^{2}$ vanishes
$(E_{3}^{2}$ is integrable$)$.
2. For $\beta-\alpha \neq \pm \displaystyle\frac{\pi}{2}$, the geodesic torsion
$T(\beta,\alpha)$ is symmetric with respect to the versor fields
$(C,\overline{\xi})$, $(C,\overline{\alpha})$, if and only if the
nonholonomic manifold $E_{3}^{2}$ has null mean curvature
$($$E_{3}^{2}$ is minimal$)$.\end{theorem}
\section{Parallelism, conjugation and orthonormal conjugation}
\setcounter{theorem}{0}\setcounter{equation}{0}
The notion of conjugation of versor field $(C,\overline{\xi})$
with tangent field $(C,\overline{\alpha})$ on the nonholonomic
manifold $E_{3}^{2}$ is that studied for the associated Myller
configuration $\mathfrak{M}_t$,
So that $(C,\overline{\xi})$ are conjugated with
$(C,\overline{\alpha})$ on $E_{3}^{2}$ if and only if $K(\delta,d)=0$ i.e.
\begin{equation}
(p_{1}\omega_{1} +p_{2}\omega_{2})\widetilde{\omega}_{2} - (q_{1}\omega_{1}
+q_{2}\omega_{2})\widetilde{\omega}_{1} = 0
\end{equation}
or, by means of (5.3.11):
\begin{equation}
(p_{1}\cos\alpha + p_{2}\sin \alpha)\sin \beta - (q_{1}\cos\alpha +q_{2}\sin
\alpha)\cos \beta = 0.
\end{equation}
All propositions established in section 3.3, Chapter 3, can be
applied.
The notion of orthogonal conjugation of $(C, \overline{\xi})$ with
$(C,\overline{\alpha})$ is given by $T(\delta,d) = 0$ or by
\begin{equation}
(p_{1}\omega_{1} + p_{2}\omega_{2})\widetilde{\omega}_{1} + (q_{1}\omega_{1} +
q_{2}\omega_{2})\widetilde{\omega}_{2} = 0
\end{equation}
or, by means of (5.3.11), it is characterized by
\begin{equation}
(p_{1}\cos \alpha + p_{2}\sin \alpha)\cos \beta + (q_{1}\cos \alpha + q_{2}\sin
\alpha)\sin \beta = 0.
\end{equation}
The relation of conjugation is symmetric iff $E_{3}^{2}$ is
integrable $(T_{m} = 0)$ and that of orthogonal conjugation is
symmetric iff $E_{3}^{2}$ is minimal (i.e. $H=0$).
Now we study the case of tangent versor field $(C,\overline{\xi})$
parallel along $C$, in $E_{3}^{2}$. Applying the theory of
parallelism in $\mathfrak{M}_{t}(C,\overline{\xi},\pi)$, from
Chapter 3 we obtain:
\begin{theorem}
The versors $(C,\overline{\xi})$, tangent to the manifold
$E_{3}^{2}$ along the curve $C$ is parallel in the Levi-Civita
sense if and only if the following system of equations holds
\begin{equation}
\displaystyle\frac{\widetilde{\omega}_{1}}{\delta
s}\left[\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\widetilde{\omega}_{2}}{\delta s }\right)
+\displaystyle\frac{r}{ds}\displaystyle\frac{\widetilde{\omega}_{1}}{\delta s} \right] -
\displaystyle\frac{\widetilde{\omega}_{2}}{\delta
s}\left[\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\widetilde{\omega}_{1}}{\delta s }\right)
- \displaystyle\frac{r}{ds}\displaystyle\frac{\widetilde{\omega}_{2}}{\delta s} \right] = 0.
\end{equation}
\end{theorem}
But, the previous system is equivalent to the system:
\begin{eqnarray}
\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\widetilde{\omega}_{1}}{\delta s}\right) -
\displaystyle\frac{r}{ds}\displaystyle\frac{\widetilde{\omega}_{2}}{\delta s} &=&
h(s)\displaystyle\frac{\widetilde{\omega}_{1}}{\delta
s}\nonumber\\\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\widetilde{\omega}_{2}}{\delta
s}\right) +\displaystyle\frac{r}{ds}\displaystyle\frac{\widetilde{\omega}_{1}}{\delta s} &=&
h(s)\displaystyle\frac{\widetilde{\omega}_{2}}{\delta s},\;\; \forall h(s)\neq 0.
\end{eqnarray}
Since, the tangent vector field $h(s)\overline{\xi}(s)$ has the
same direction with the versor $\overline{\xi}(s)$, from (5.2.11) we
obtain $G(\widetilde{\delta},d) = h^{2}G(\delta,d)$, $\widetilde{\delta}$
being the direction of tangent vector $h(s)\overline{\xi}$.
Therefore, the equations $G(\delta,d)=0$ is invariant with respect to
the applications $\overline{\xi}(s)\to h(s)\overline{\xi}(s)$.
Thus the equations (5.4.3) or (5.4.4) characterize the parallelism of
directions $(C, h(s) \overline{\xi}(s))$ tangent to $E_{3}^{2}$
along $C$.
All properties of the parallelism of versors $(C,\xi)$ studied for
the configuration $\mathfrak{M}_{t}$ in sections 2.8, Chapter 2 are
applied here.
Theorem of existence and uniqueness of parallel of tangent versors
$(C,\overline{\xi})$ on $E_{3}^{2}$ can be formulated exactly as
for this notion in $\mathfrak{M}_{t}$.
But a such kind of theorem can be obtained by means of the
following equation, given by (5.3.10):
\begin{equation}
G(\beta,\alpha) \equiv \displaystyle\frac{d\beta}{ds} + r_{1}(s)\cos \alpha +r_{2}(s)\sin \alpha
= 0.
\end{equation}
The parallelism of vector field $(C,\overline{V})$ tangents to
$E_{3}^{2}$ along the curve $C$ can be studied using the form
$$
\overline{V}(s)= \|\overline{V}(s)\|\overline{\xi}(s)
$$
where $\overline{\xi}$ is the versor of $\overline{V}(s)$.
Also, we can start from the definition of Levi-Civita parallelism
of tangent vector field $(C,\overline{V})$, expressed by the
property
\begin{equation}
\displaystyle\frac{d\overline{V}}{ds} = h(s)I_{3}.
\end{equation}
Setting
\begin{equation}
\overline{V}(s) = V^{1}(s)I_{1} + V^{2}(s)I_{2}
\end{equation}
and using (5.4.8) we obtain the system of differential equations
\begin{equation}
\displaystyle\frac{dV^{1}}{ds} - \displaystyle\frac{r}{ds}V^{2}=0,\ \
\displaystyle\frac{dV^{2}}{ds}+\displaystyle\frac{r}{ds}V^{1}=0.
\end{equation}
All properties of parallelism in Levi-Civita sense of tangent
vectors $(C,\overline{V})$ studied in section 2.8, Chapter 2 for
Myller configuration are valid. For instance
\begin{theorem}
The Levi-Civita parallelism of tangent vectors $(C,\overline{V})$
in the manifold $E_{3}^{2}$ preserves the lengths and angle of
vectors.
\end{theorem}
The concurrence in Myller sense of tangent vector fields
$(C,\overline{\xi})$ is characterized by Theorem 2.9.2 Chapter 2, by
the following equations
\begin{equation}
\displaystyle\frac{d}{ds}\left(\displaystyle\frac{c_{2}}{G}\right) = c_{1},
\end{equation}
where $c_{1} = \langle \overline{\alpha},\overline{\xi}\rangle$,
$c_{2} = \langle \overline{\alpha}, \overline{\mu}\rangle,$ $G =
G(\delta,d)\neq 0.$
Consequence: the concurrence of tangent versor field
$(C,\overline{\xi})$ in $E_{3}^{2}$ for $G\neq 0$ is characterized
by
\begin{equation}
\displaystyle\frac{d}{ds}\left[\displaystyle\frac{\widetilde{\omega}_{1}\omega_{2} -
\widetilde{\omega}_{2}\omega_{1}}{\delta s\cdot ds}\cdot \displaystyle\frac{1}{G}\right] =
\displaystyle\frac{\widetilde{\omega}_{1}\omega_{1} + \widetilde{\omega}_{2}\omega_{2}}{\delta sds}
\end{equation}
or by:
\begin{equation}
\displaystyle\frac{d}{ds}\left[\displaystyle\frac{1}{G}\sin (\alpha - \beta)\right] =\cos (\alpha-\beta).
\end{equation}
The properties of concurrence in Myller sense can be taken from
section 2.8, Chapter 2.
\section{Theory of curves in $E_{3}^{2}$}
\setcounter{theorem}{0}\setcounter{equation}{0}
To a curve $C$ in the nonholonomic manifold $E_{3}^{2}$ one can
uniquely associate a Myller configuration $\mathfrak{M}_{t} =
\mathfrak{M}_{t}(C, \overline{\alpha}, \pi)$ where $\pi(s)$ is the
orthogonal to the normal versor $I_{3}(s)$ of $E_{3}^{2}$ at every
point $P\in C.$
Thus the geometry of curves in $E_{3}^{2}$ is the geometry of
associated Myller configurations $\mathfrak{M}_{t}$. It is
obtained as a particular case taking $\overline{\xi}(s) =
\overline{\alpha}(s)$ from the previous sections of this chapter.
The Darboux frame of the curve $C$, in $E_{3}^{2}$ is given by
$$
\cal{R}_{D} = \{P(s); \overline{\alpha}(s), \overline{\mu}^{*}(s),
I_{3}(s)\},\;\; \overline{\mu}^{*} = I_{3}\times \overline{\alpha}
$$
and the fundamental equations of $C$ in $E_{3}^{2}$ are as
follows:
\begin{equation}
\displaystyle\frac{d\overline{r}}{ds} = \overline{\alpha}(s)
\end{equation}
and
\begin{eqnarray}
\displaystyle\frac{d\overline{\alpha}}{ds} &=& \kappa_{g}(s)\overline{\mu}^{*} +
\kappa_{n}(s)I_{3},\\\displaystyle\frac{d\overline{\mu}^{*}}{ds} &=&
-\kappa_{g}(s)\overline{\alpha} +\tau_{g}(s)I_{3},\\\displaystyle\frac{dI_{3}}{ds}&=&
-\kappa_{n}(s)\overline{\alpha} - \tau_{g}(s)\overline{\mu}^{*}.
\end{eqnarray}
The invariant $\kappa_{g}(s)$ is the {\it geodesic curvature} of $C$
at point $P\in C,$ $\kappa_{n}(s)$ is the {\it normal curvature} of
curve $C$ at $P\in C$ and $\tau_{g}(s)$ it {\it geodesic torsion}
of $C$ at $P\in C$ in $E_{3}^{2}$.
The geometrical interpretations of these invariants are exactly
those described in the section 3.2, Chapter 3. Also, a fundamental
theorem can be enounced as in section 3.2, Theorem 3.2.2, Chapter
3.
The calculus of $\kappa_{g}$, $\kappa_{n}$ and $\tau_{g}$ is obtained by
the formulae (5.3.2), for $\overline{\xi}=\overline{\alpha}$.
We have:
\begin{eqnarray}
\kappa_{g}&=&\left\langle \overline{\alpha}, \displaystyle\frac{d\overline{\alpha}}{ds},
I_{3}\right\rangle,\\\kappa_{n}&=&
\left\langle\displaystyle\frac{d\overline{\alpha}}{ds},I_{3}\right\rangle =
-\left\langle\overline{\alpha},\displaystyle\frac{dI_{3}}{ds}\right\rangle,\\\tau_{g} &=&
\left\langle\overline{\alpha}, I_{3}, \displaystyle\frac{dI_{3}}{ds}\right\rangle.
\end{eqnarray}
By using the expressions (3.1.4), Ch. 3 we
obtain
\begin{equation}
\kappa_{g} = \kappa \sin \varphi^{*},\; \kappa_{n}=\kappa\cos \varphi^{*},\; \tau_{g} =
\tau+\displaystyle\frac{d\varphi^{*}}{ds}
\end{equation}
with $\varphi^{*} = \sphericalangle (\overline{\alpha}_2,I_3)$, $\overline{\alpha}_{2}$ being the versor of
principal normal of curve $C$ at point $P$.
A theorem of Meusnier can be formulates as in section 3.1, Chapter
3.
The line $C$ is {\it geodesic} (or {\it autoparallel}) {\it line} of the
nonholonomic manifold $E_{3}^{2}$ if $\kappa_{g}(s) = 0$, $\forall
s\in (s_{1},s_{2})$.
\begin{theorem}
Along a geodesic $C$ of the nonholonomic manifold $E_{3}^{2}$ the
normal curvature $\kappa_{n}$ is equal to $\pm \kappa$ and geodesic
torsion $\tau_{g}$ is equal to the torsion $\tau$ of $C$.
The differential equations of geodesics are as follows
\begin{equation}
\kappa_{g} \equiv
\displaystyle\frac{\omega_{1}}{ds}\left[\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\omega_{2}}{ds}\right)
+ r_{1} \right] -
\displaystyle\frac{\omega_{3}}{ds}\left[\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\omega_{1}}{ds}\right)
- r_{2} \right]=0.
\end{equation}
\end{theorem}
This equations are equivalent to the system of differential equations
\begin{eqnarray}
\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\omega_{1}}{ds}\right) -r_{2}(s)&=&
h(s)\displaystyle\frac{\omega_{1}}{ds},\\\displaystyle\frac{d}{ds}\left(\displaystyle\frac{\omega_{2}}{ds}\right)
+ r_{1}(s) &=& h(s)\displaystyle\frac{\omega_{2}}{ds},
\end{eqnarray}
where $h(s)\neq 0$ is an arbitrary smooth function.
If we consider $\sigma = \sphericalangle (I_{3}, \overline{\nu}_{3})$,
where $\overline{\nu}_{3}$ the versor of binormal is of the
field $(C,I_{3})$ and $\chi_{2}$ is the torsion of $(C, I_{3})$,
then the $\kappa_{g}(s)$, the geodesic curvature of the field
$(C,\overline{\alpha})$ is obtained by the formulas of Gh. Gheorghiev:
\begin{equation}
\kappa_{g} = \displaystyle\frac{d\sigma}{ds}+\chi_{2}.
\end{equation}
It follows that: the geodesic lines of the nonholonmic manifold
$E_{3}^{2}$ are characterized by the differential equations:
\begin{equation}
\displaystyle\frac{d\sigma}{ds} + \chi_{2}(s) = 0. \end{equation}
By means of this equations we can prove a theorem of existence of
uniqueness of geodesic on the nonholonomic manifold $E_{3}^{2}$,
when $\sigma_{0} = \sphericalangle (I_{3}(s_{0}),
\overline{\nu}_{3}(s_{0}))$ is a priori given.
\section{The fundamental forms of $E_{3}^{2}$}
\setcounter{theorem}{0}\setcounter{equation}{0}
The first, second and third fundamental forms $\phi, \psi$ and
$\chi$ of the nonholonomic manifold $E_{3}^{2}$ are defined as in
the theory of surfaces in the Euclidean space $E_{3}.$
The tangent vector $d\overline{r}$ to a curve $C$ at point $P\in
C$ in $E_{3}^{2}$ is given by:
\begin{equation}
d\overline{r} = \omega_{1}(s)I_{1} + \omega_{2}(s)I_{2},\ \omega_3=0.
\end{equation}
The first fundamental form of $E_{3}^{2}$ at point $P\in
E_{3}^{2}$ is defined by:
\begin{equation}
\phi = \langle d\overline{r},d\overline{r}\rangle =
(\omega_{1}(s))^{2} + (\omega_{2}(s))^{2},\ \omega_3=0.
\end{equation}
Clearly, $\phi$ has geometric meaning and it is a quadratic positive
defined form.
The arclength of $C$ is determined by the formula (5.2.10) and the
arclength element $ds$ is given by:
\begin{equation}
ds^{2} = \phi = (\omega^{1})^{2} + (\omega^{2})^{2},\ \omega_3=0.
\end{equation}
The angle of two tangent vectors $d\overline{r}$ and $\delta
\overline{r} =\widetilde{\omega}_{1} I_{1} + \widetilde{\omega}_{2}I_{2}$
is expressed by
\begin{equation}
\cos \theta = \displaystyle\frac{\langle \delta \overline{r}, d\overline{r}\rangle}{\delta
sds} = \displaystyle\frac{\widetilde{\omega}_{1}\omega_{1} +
\widetilde{\omega}_{2}\omega_{2}}{\sqrt{(\omega_{1})^{2}+(\omega_{2})^{2}}\sqrt{(\widetilde{\omega}_{1})^{2}
+ (\widetilde{\omega}_{2})^{2} }}.
\end{equation}
The second fundamental from $\psi$ of $E_{3}^{2}$ at point $P$ is
defined by
\begin{equation}
\psi =-\langle d\overline{r}, dI_{3}\rangle = p\omega_{2} - q \omega_{1} =
p_{2}\omega_{2}^{2} +(p_{1} - q_{2})\omega_{1}\omega_{2} - q_{1}\omega_{1}^{2}.
\end{equation}
The form $\psi$ has a geometric meaning. It is not symmetric.
The third fundamental form $\chi$ of $E_{3}^{2}$ at point $P$ is
defined by
\begin{equation}
\chi = \langle d\overline{r},\overline{\theta}\rangle,\ \omega_3=0,
\end{equation}
where $\overline{\theta}$ is given by (5.1.7).
As a consequence, we have
\begin{equation}
\chi = p\omega_{1}+q\omega_{2} = p_{1}\omega_{1}^{2}
+(p_{2}+q_{1})\omega_{1}\omega_{2} + q_{2}\omega_{2}^{2}
\end{equation}
$\chi$ has a geometrical meaning and it is not symmetric.
One can introduce a fourth fundamental form of $E_{3}^{2}$, by
$$
\Theta = \langle dI_{3}, dI_{3}\rangle = p^{2} + q^{2}\; (mod\;
\omega_{3}).
$$
But $\Theta$ linearly depends by the forms $\phi, \psi, \chi$.
Indeed, we have
$$
\Theta = M\psi - K_{t}\phi - T_{m}\chi,
$$
where $M$ is mean curvature, $T_{m}$ is mean torsion and $K_{t}$
is total curvature of $E_{3}^{2}$ at point $P.$ The expression of
$K_{t}$ is
\begin{equation}
K_{t} = p_{1}q_{2} - p_{2}q_{1}.
\end{equation}
The formulae (5.5.6) and (5.5.7) and the fundamental forms $\phi, \psi, \chi$
allow to express the normal curvature and geodesic torsion of a
curve $C$ at a point $P\in C$ as follows
\begin{equation}
\kappa_{n} = \displaystyle\frac{\psi}{\phi} = \displaystyle\frac{p\omega_{2} - q\omega_{1}}{\omega_{1}^{2}
+\omega_{2}^{2}} = \displaystyle\frac{p_{2}\omega_{2}^{2} + (p_{1} - q_{2})\omega_{1}\omega_{2}
- q_{1}\omega_{1}^{2}}{\omega_{1}^{2} + \omega_{2}^{2}}
\end{equation}
and
\begin{equation}
\tau_{g} = \displaystyle\frac{\chi}{\phi} = \displaystyle\frac{p\omega_{1} +q\omega_{2}}{\omega_{1}^{2}
+ \omega_{2}^{2}} = \displaystyle\frac{p_{1}\omega_{1}^{2} + (p_{2} +q_{1})\omega_{1}\omega_{2}
+ q_{2}\omega_{2}^{2}}{\omega_{1}^{2} + \omega_{2}^{2}}.
\end{equation}
The cases when $\kappa_{n} = 0$ and $\tau_{g} =0$ are important. They
will be investigated in the next section.
\section{Asymptotic lines. Curvature lines}
\setcounter{theorem}{0}\setcounter{equation}{0}
An asymptotic tangent direction $d\overline{r}$ to $E^2_3$ at a point $P\in
E_{3}^{2}$ is defined by the property $\kappa_{n}(s) = 0.$ A line $C$
in $E_{3}^{2}$ for which the tangent directions $d\overline{r}$
are asymptotic directions is called an {\it asymptotic line} of the
nonholonomic manifold $E_{3}^{2}$.
The asymptotic directions are characterized by the following
equation of degree 2.
\begin{equation}
p_{2}\omega_{2}^{2} + (p_{1} - q_{2})\omega_{1}\omega_{2} - q_{1}\omega_{1}^{2}
=0.
\end{equation}
The realisant of this equation is the invariant
\begin{equation}
K_{g} = K_{t} - \displaystyle\frac{1}{4}T_{m}^{2},
\end{equation}
called the {\it gaussian curvature} of $E_{3}^{2}$ at point $P\in
E_{3}^{2}$.
We have
\begin{theorem}
At every point $P\in E_{3}^{2}$ there are two asymptotic
directions:
- real if $K_{g}<0$
- imaginary if $K_{g}>0$
- coincident if $K_{g}=0$
\end{theorem}
The point $P\in E_{3}^{2}$ is called {\it planar} if the
asymptotic directions of $E_{3}^{2}$ at $P$ are nondeterminated.
A planar point is characterized by the equations
\begin{equation}
p_{1} - q_{2} = 0,\;\; p_{2} = q_{1} = 0.
\end{equation}
The point $P\in E_{3}^{2}$ is called {\it elliptic} if the
asymptotic direction of $E_{3}^{2}$ at $P$ are imaginary and $P$
is a {\it hyperbolic} point if the asymptotic directions of
$E_{3}^{2}$ at $P$ are real.
Of course, if $P$ is a hyperbolic point of $E_{3}^{2}$ then,
exists two asymptotic line through the point $P$, tangent to the
asymptotic directions, solutions of the equations (5.7.1).
The curvature direction $d\overline{r}$ at a point $P\in E_{3}$ is
defined by the property $\tau_{g}(s) = 0.$ A line $C$ in
$E_{3}^{2}$ for which the tangent directions $d\overline{r}$ are
the curvature directions is called a {\it curvature line} of the
nonholonomic manifold $E_{3}^{2}$.
The curvature directions are characterized by the following second
order equations.
\begin{equation}
p_{1}\omega_{1}^{2} + (p_{2}+q_{1})\omega_{1}\omega_{2} + q_{2}\omega_{2}^{2} =0.
\end{equation}
The realisant of this equations is
\begin{equation}
T_{t} = K_{t} - \displaystyle\frac{1}{4}M^{2}.
\end{equation}
We have
\begin{theorem}
At every point $P\in E_{3}^{2}$ there are two curvature directions
real, if $T_{t}<0$
imaginary, if $T_{t} >0$
coincident, if $T_{t} = 0$
\end{theorem}
The curvature lines on $E_{3}^{2}$ are obtained by integrating the
equations (5.7.1) in the case $T_{t}\leq 0$.
\begin{remark}
In the case of surfaces $(T_{m}=0)$, the curvature lines coincides
with the lines of extremal normal curvature.
\end{remark}
\section{The extremal values of $\kappa_{n}$. Euler formula. Dupin indicatrix}
\setcounter{theorem}{0}\setcounter{equation}{0}
The extremal values at a point $P\in E_{3}^{2}$, of the normal
curvature $\kappa_{n} = \displaystyle\frac{\psi}{\phi}$ when $(\omega_{1},\omega_{2})$ are
variables are given by $\phi \displaystyle\frac{\partial \psi}{\partial \omega_{i}} -
\psi\displaystyle\frac{\partial\phi}{\partial \omega_{i}} = 0,$ $(i=1,2)$ which are equivalent
to the equations
\begin{equation}
\displaystyle\frac{\partial\psi}{\partial w_{i}} - \kappa_{n}\displaystyle\frac{\partial\phi}{\partial\omega_{i}} = 0,\;\; (i=1,2).
\end{equation}
Taking into account the form (5.6.5) of $\psi$ and (5.6.3) of $\phi$,
the system (5.8.1) can be written:
\begin{eqnarray}
-(q_{1} + \kappa_{n})\omega_{1} + \displaystyle\frac{1}{2}(p_{1}
-q_{2})\omega_{2}&=&0\\\displaystyle\frac{1}{2}(p_{1} - q_{1})\omega_{1}
+(p_{2}-\kappa_{n})\omega_{2} &=& 0.
\end{eqnarray}
It is a homogeneous and linear system in $(\omega_{1},\omega_{2})$-which
gives the directions $(\omega_{1},\omega_{2})$ of extremal values of
normal curvature.
But, the previous system has solutions if and only if the following equations hold
$$
\left|\begin{array}{ccc} -(q_{1} +\kappa_{n})&\displaystyle\frac{1}{2}(p_{1} -
q_{2})\\\displaystyle\frac{1}{2}(p_{1}-q_{2})&p_{2} - \kappa_{n}
\end{array}\right|=0
$$
equivalent to the following equations of second order in $\kappa_{n}$:
\begin{equation}
\kappa_{n}^{2} - (p_{2} - q_{1})\kappa_{n} +p_{1}q_{2} - p_{2}q_{1} -
\displaystyle\frac{1}{4}(p_{1} + q_{2})^{2} = 0.
\end{equation}
This equation has two real solutions $\displaystyle\frac{1}{R_{1}}$ and
$\displaystyle\frac{1}{R_{2}}$ because its realisant is
\begin{equation}
\Delta = (p_{2} +q_{1})^{2} + (p_{1} -q_{2})^{2}.
\end{equation}
We have $\Delta\geq 0.$ Therefore the solutions
$\displaystyle\frac{1}{R_{1}},\displaystyle\frac{1}{R_{2}}$ are real, different or equal,
$\displaystyle\frac{1}{R_{1}}, \displaystyle\frac{1}{R_{2}}$ are the extremal values of
normal curvature $\kappa_{n}$.
Let us denote \begin{equation} H = \displaystyle\frac{1}{R_{1}}+\displaystyle\frac{1}{R_{2}}
=p_{2} - q_{1}
\end{equation}
called the mean curvature of the nonholonomic manifold $E_{3}^{2}$
at point $P$, and
\begin{equation}
K_{g} = \displaystyle\frac{1}{R_{1}}\displaystyle\frac{1}{R_{2}} = p_{1}q_{2} - p_{2}q_{1} -
\displaystyle\frac{1}{4}(p_{1} +q_{2}).
\end{equation}
called the Gaussian curvature of $E_{3}^{2}$ at point $P$.
Substituting $\displaystyle\frac{1}{R_{1}},$ $\displaystyle\frac{1}{R_{2}}$ in the equations
(5.8.2) we have the directions of extremal values of the normal
curvature-called the {\it principal directions} of $E_{3}^{2}$ at
point $P\in E_{3}^{2}$. These directions are obtained from (5.8.2)
for $\kappa_{n} = \displaystyle\frac{1}{R_{1}}$, $\kappa_{n}=\displaystyle\frac{1}{R_{2}}$. Thus,
one obtains the following equations which determine the principal
directions:
\begin{equation}
(p_{1} - q_{2})\omega_{1}^{2} +2(p_{2}+q_{1})\omega_{1}\omega_{2} - (p_{1} -
q_{2})\omega_{2}^{2}=0.
\end{equation}
Let $(\omega_{1}, \omega_{2}), (\widetilde{\omega}_{1},\widetilde{\omega}_{2})$
the solution of (5.8.8). Then $d\overline{r} = \omega_{1}I_{1} +
\omega_{2}I_{2}$ and $\delta \overline{r} = \widetilde{\omega}_{1}I_{1} +
\widetilde{\omega}_{2}I_{2}$ are the principal directions on
$E_{3}^{2}$ at point $P.$
The principal directions $d\overline{r},\delta\overline{r}$ of $E_{3}^{2}$ in
every point $P$ are real and orthogonal.
Indeed, because $\Delta>0$, and if $\displaystyle\frac{1}{R_{1}}\neq
\displaystyle\frac{1}{R_{2}}$ we have $\langle \delta \overline{r},
d\overline{r}\rangle = 0$.
The curves on $E_{3}^{2}$ tangent to $\delta \overline{r}$ and
$d\overline{r}$ are called the {\it lines of extremal normal curvature}.
So, we have
\begin{theorem}
At every point $P\in E_{3}^{2}$ there are two real and orthogonal
lines of extremal normal curvature.\end{theorem}
Assuming that the frame $\mathcal{R} = (P; I_{1}, I_{2},I_{3})$
has the vectors $I_{1}, I_{2}$ in the principal directions $\delta
\overline{r}, d\overline{r}$ respectively, thus the equation (5.8.8)
implies
\begin{equation}
p_{1} - q_{2} = 0
\end{equation}
and the extremal values of normal curvature $\kappa_{n}$ are given by
\begin{equation}
\kappa_{n}^{2} - (p_{2}-q_{1})\kappa_{n}-p_{2}q_{1} =0.
\end{equation}
We have
\begin{equation}
\displaystyle\frac{1}{R_{1}} = p_{2},\;\;\displaystyle\frac{1}{R_{2}} = -q_{1}.
\end{equation}
Denoting by $\alpha$ the angle between the versor $I_{1}$ and the
versor $\overline{\xi}$ of an arbitrary direction at $P$, tangent
to $E_{3}^{2}$ we can write the normal curvature $\kappa_{n}$ from
(5.6.8) in the form
\begin{equation}
\kappa_{n} = \displaystyle\frac{1}{R_{1}}\cos^{2}\alpha +
\displaystyle\frac{1}{R_{2}}\sin^{2}\alpha,\;\; \alpha =
\sphericalangle(\overline{\xi},I_{1}).
\end{equation}
The formula (5.8.12) is called the {\it Euler formula} for normal
curvatures on the nonholonomic manifold $E_{3}^{2}$.
Consider the tangent vector $\overrightarrow{PQ} =
|\kappa_{n}|^{-1/2}\overline{\xi}$, i.e.
$$
\overrightarrow{PQ} = |\kappa_{n}|^{-\displaystyle\frac{1}{2}}(\cos \alpha I_{1} +\sin
\alpha I_{2} ).
$$
The cartesian coordinate $(x,y)$ of the point $Q$, with respect to
the frame $(P; I_{1}, I_{2})$ are given by
\begin{equation}
x = |\kappa_{n}|^{-\displaystyle\frac{1}{2}}\cos \alpha, \;\; y =
|\kappa_{n}|^{-\displaystyle\frac{1}{2}}\sin \alpha.
\end{equation}
Thus, eliminating the parameter $\alpha$ from (5.8.12) and (5.8.13) are
gets the geometric locus of the point $Q$ (in the tangent plan
$(P;I_{1},I_{2})$):
\begin{equation}
\displaystyle\frac{x^{2}}{R_{1}} + \displaystyle\frac{y^{2}}{R_{2}} = \pm 1
\end{equation}
called {\it the Dupin$'$s indicatrix} of the normal curvature of
$E_{3}^{2}$ at point $P$.
It is formed by a pair of conics, having the axis in the principal
directions $\delta \overline{r}, d \overline{r}$ and the invariants:
$H$-the mean curvature and $K_{g}$ - the gaussian curvature of
$E_{3}^{2}$ at $P$.
The Dupin indicatrix is formed by two ellipses, one real and
another imaginary, if $K_{g}>0.$ It is formed by a pair of
conjugate hyperbolae if $K_{g}<0$, whose asymptotic lines are
tangent to the asymptotic lines of $E_{3}^{2}$ at point $P$. It
follows that the asymptotic directions of $E_{3}^{2}$ at $P$ are
symmetric with respect to the principal directions of $E_{3}^{2}$
at $P.$
In the case $K_{g} = 0$, the Dupin indicatrix of $E_{3}^{2}$ at
$P$ is formed by a pair of parallel straight lines - one real and
another imaginary.
\section{The extremal values of $\tau_{g}$. Bonnet formula. Bonnet indicatrix}
\setcounter{theorem}{0}\setcounter{equation}{0}
The extremal values of the geodesic torsion $\tau_{g} =
\displaystyle\frac{\chi}{\phi}$ at the point $P$, on $E_{3}^{2}$ can be
obtained following a similar way as in the previous paragraph.
Such that, the extremal values of $\tau_{g}$ are given by the
system of equations
$$
\displaystyle\frac{\partial \chi}{\partial \omega_{i}} - \tau_{g}\displaystyle\frac{\partial \phi}{\partial \omega_{i}} = 0,
(i=1,2).
$$
Or, taking into account the expressions of $\phi = (\omega_{1})^{2}
+ (\omega_{2})^{2}$ and $\chi = p\omega_{1} + q\omega_{2} = (p_{1}\omega_{1}
+p_{2}\omega_{2})\omega_{1} +(q_{1}\omega_{1} + q_{2}\omega_{2})\omega_{2}$, we have:
\begin{eqnarray}
(p_{1}-\tau_{g})\omega_{1} + \displaystyle\frac{1}{2}(p_{2}+q_{1})\omega_{2} =
0\\\displaystyle\frac{1}{2}(p_{2} +q_{1})\omega_{1} + (q_{2}-\tau_{g})\omega_{2} =
0.\nonumber
\end{eqnarray}
But, these imply
\begin{equation}
\left|
\begin{array}{ccc}
p_{1} - \tau_{g}& \displaystyle\frac{1}{2}(p_{2}
+q_{1})\\\displaystyle\frac{1}{2}(p_{2}+q_{1})&q_{2} - \tau_{g}
\end{array}\right|=0
\end{equation}
or expanded:
\begin{equation}
\tau_{g}^{2} - (p_{1}+q_{2})\tau_{g} +(p_{1}q_{2} - p_{2}q_{1} -
\displaystyle\frac{1}{4}(p_{2} - q_{1})^{2}) = 0.
\end{equation}
The solutions $\displaystyle\frac{1}{T_{1}}, \displaystyle\frac{1}{T_{2}}$ of this equations
are the extremal values of geodesic torsion $\tau_{g}$. They are
real. The realisant of (5.9.3) is
\begin{equation}
\Delta_{1} = (p_{1} - q_{2})^{2} +(p_{2}+q_{1})^{2}
\end{equation}
Therefore $\displaystyle\frac{1}{T_{1}},$ $\displaystyle\frac{1}{T_{2}}$ are real, distinct
or coincident.
We denote
\begin{equation}
T_{m} = \displaystyle\frac{1}{T_{1}} + \displaystyle\frac{1}{T_{2}} = p_{1} +q_{2},
\end{equation}
{\it the mean torsion} of $E_{3}^{2}$ at point $P.$ And
\begin{equation}
T_{t} = \displaystyle\frac{1}{T_{1}}\displaystyle\frac{1}{T_{2}} = p_{1}q_{2} - p_{2}q_{1} -
\displaystyle\frac{1}{4}(p_{2} -q_{1})^{2} =K_{g} - \displaystyle\frac{1}{4}H^{2}
\end{equation}
is called {\it the total torsion} of $E_{3}^{2}$ at $P.$
But, as we know the condition of integrability of the Pfaf
equation $\omega_{3}=0$ is
$$
d\omega_{3} = 0,\; \mbox{modulo}\; \omega_{3},
$$
which is equivalent to $T_{m} = p_{1} +q_{2} = 0.$
\begin{theorem}
The nonholonomic manifold $E_{3}^{2}$ of mean torsion $T_{m}$
equal to zero is a family of surfaces in $E_{3}$.
\end{theorem}
The directions $\delta \overline{r} = \widetilde{\omega}_{1}I_{1} +
\widetilde{\omega}_{2}I_{2}$ and $d\overline{r} = \omega_{1}I_{1} +
\omega_{2}I_{2}$ of extremal geodesic torsion, corresponding to the
extremal values $\displaystyle\frac{1}{T_{1}}$ and $\displaystyle\frac{1}{T_{2}}$ of
$\tau_{g}$ are given by the equation obtained from (5.9.2) by
eliminating $\tau_{g}$ i.e.
\begin{equation}
(p_{2} + q_{1})\omega_{1}^{2} - (p_{1}-q_{2})\omega_{1}\omega_{2} -
(p_{2}+q_{1})\omega_{2}^{2} = 0.
\end{equation}
Thus $\delta \overline{r}$ and $d\overline{r}$ are real and orthogonal
at every point $P\in E_{3}^{2}$ for which $\displaystyle\frac{1}{T_{1}}\neq
\displaystyle\frac{1}{T_{2}}$. $\delta \overline{r}$ and $d \overline{r}$ are
called the direction of the {\it extremal geodesic torsion} at
point $P\in E_{3}^{2}.$
\begin{theorem}
At every point $P\in E_{3}^{2}$, where $\displaystyle\frac{1}{T_{1}}\neq
\displaystyle\frac{1}{T_{2}}$ there exist two real and orthogonal directions of
extremal geodesic torsion.
\end{theorem}
Now, assuming that, at point $P$, the frame $\cal{R} = (P;I_{1},
I_{2}, I_{3})$ has the vectors $I_{1}$ and $I_{2}$ in the
directions $\delta \overline{r}$, $d\overline{r}$ of the extremal
geodesic torsion, respectively, from (5.9.7) we deduce
\begin{equation}
p_{2} + q_{1} =0.
\end{equation}
Thus, the equation (5.9.3) takes the form
\begin{equation}
\tau_{g}^{2} - (p_{1}+q_{2})\tau_{g} +p_{1}q_{2} = 0.
\end{equation}
Its solutions are
\begin{equation}
\displaystyle\frac{1}{T_{1}} = p_{1},\;\; \displaystyle\frac{1}{T_{2}} = q_{2}
\end{equation}
and, setting $\beta = \sphericalangle (\overline{\xi}, I_{1})$, from
$\tau_{g} = \displaystyle\frac{\chi}{\phi}$ one obtains [65] the so called {\it
Bonnet formula} giving $\tau_{g}$ for the nonholonomic manifold
$E_{3}^{2}$:
\begin{equation}
\tau_{g}=\displaystyle\frac{\cos^{2}\beta}{T_{1}}+\displaystyle\frac{\sin^{2}\beta}{T_{2}},\; \beta =
\sphericalangle (\overline{\xi},I_{1}).
\end{equation}
If $T_{m} = 0,$ (5.9.11) reduce to the very known formula from
theory of surfaces:
\begin{equation}
\tau_{g} = \displaystyle\frac{1}{T_{1}}\cos 2\beta.
\end{equation}
By means of the formula (5.9.9) we can determine an indicatrix of
geodesic torsions.
In the tangent plane at point $P$ we take the cartesian frame $(P;
I_{1}, I_{2})$, $I_{1}$ having the direction of extremal geodesic
torsion $\delta \overline{r}$ corresponding to $\displaystyle\frac{1}{T_{1}}$ and
$I_{2}$ having the same direction with $d\overline{r}$
corresponding to $\displaystyle\frac{1}{T_{2}}$. Consider the point $Q^{\prime}
\in \pi(s)$ given by
\begin{equation}
\overrightarrow{PQ^{\prime}} = |\tau_{g}|^{-1}\overline{\xi} =
|\tau_{g}|^{-1}(\cos \beta I_{1} + \sin \beta I_{2})
\end{equation}
with $\beta = \sphericalangle (\overline{\xi},I_{1})$. The cartesian
coordinate $(x,y)$ at point $Q^{\prime}$:
$\overrightarrow{PQ^{\prime}} = xI_{1} +y I_{2}$, by means of
(5.9.13) give us
\begin{equation}
x = |\tau_{g}|^{-1}\cos\beta,\;\; y = |\tau_{g}|^{-1}\sin \beta.
\end{equation}
Eliminating the parameter $\beta$ from (5.9.14) and (5.9.11) one obtains
the locus of point $Q^{\prime}$:
\begin{equation}
\displaystyle\frac{x^{2}}{T_{1}} + \displaystyle\frac{y^{2}}{T_{2}} = \pm 1,
\end{equation}
called the {\it indicatrix of Bonnet} for the geodesic torsions at
point $P\in E_{3}^{2}$.
It consists of two conics having the axis the tangents in the
directions of extremal values of geodesic torsion. The invariants
of these conics are $T_{m}$-the mean torsion and $T_{t}$-the total
torsion of $E_{3}^{2}$ at point $P.$
The Bonnet indicatrix is formed by two ellipses, one real and
another imaginary, if and only if the total torsion $T_{t}$ is positive. It
is composed by two conjugated hyperbolas if $T_{t}<0.$ In this
case the curvature lines of $E_{3}^{2}$ at $P$ are tangent to the
asymptotics of these conjugated hyperbolas. Finally, if $T_{t} =
0,$ the Bonnet indicatrix (5.9.15) is formed by two pair of parallel
straightlines, one being real and another imaginary.
\noindent {\bf Remark 5.9.1}
1. If $T_{t}<0,$ the directions of asymptotics of the hyperbolas
$(5.9.15)$ are symmetric to the directions of the extremal geodesic
torsion.
2. The direction of extremal geodesic torsion are the bisectrices
of the principal directions.
3. In the case of surfaces, $T_{m} =0,$ the Bonnet indicatrix is
formed by two equilateral conjugates hyperbolas.
Assuming that the Dupin indicatrix is formed by two conjugate
hyperbolas and the Bonnet indicatrix is formed by two conjugate
hyperbolas, we present here only one hyperbolas from every of this
indicatrices as follows (Fig. 1):
\newpage
\includegraphics{indicatrices.eps}
\centerline{\footnotesize{Fig. 1}}
\section{The circle of normal curvatures and geodesic torsions}
\setcounter{theorem}{0}\setcounter{equation}{0}
Consider a curve $C\subset E_{3}^{2}$ and its tangent versor
$\overline{\alpha}$ at point $P\in C.$ Denoting by $\alpha =
\sphericalangle (\overline{\alpha},I_{1})$ and using the formulae
(5.3.8) we obtain \begin{eqnarray} \kappa_{n}&=& p_{2}\sin^{2}\alpha +
(p_{1} - q_{2})\sin \alpha \cos \alpha -
q_{1}c\cos^{2}\alpha\nonumber\\\tau_{g}&=& p_{1}\cos^{2}\alpha +(p_{2} +
q_{1})sin \alpha \cos \alpha +q_{2}\sin^{2}\alpha.
\end{eqnarray}
Eliminating the parameter $\alpha$ from (5.10.1) we deduce
\begin{equation}
\kappa_{n}^{2} +\tau_{g}^{2} -H \kappa_{n} - T_{m}\tau_{g} - K_{g} = 0
\end{equation}
where $H = p_{2}-q_{1}$ is the mean curvature of $E_{3}^{2}$ at
point $P$, $T_{m} = p_{1} + q_{2}$ is the mean torsion of
$E_{3}^{2}$ at $P$ and $K_{g} = p_{1}q_{2} - p_{2}q_{1}$ is the
Gaussian curvature of $E_{3}^{2}$ at point $P$. Therefore we have
\begin{theorem}
1. The normal curvature $\kappa_{n}$ and the geodesic torsion
$\tau_{g}$ at a point $P\in E_{3}^{2}$ satisfy the equations
$(5.10.2)$.
2. In the plan of variable $(\kappa_{n}, \tau_{g})$ the equation
$(5.10.2)$ represents a circle with the center $\(\displaystyle\frac{H}{2},
\displaystyle\frac{T_{m}}{2}\)$ and of the radius given by
\begin{equation}
R^{2} = \displaystyle\frac{1}{4}(H^{2}+T_{m}^{2})-K_{g}.
\end{equation}
\end{theorem}
Evidently, this circle is real if we have
$$
K_{g} <\displaystyle\frac{1}{4}(H^{2} + T_{m}^{2}).
$$
It is a complex circle if $K_{g}>\displaystyle\frac{1}{4}(H^{2} + T_m^{2})$.
This circle, for the nonholonomic manifolds $E_{3}^{2}$ was
introduced by the author [62]. It has been studied by Izu Vaisman
in [41], [42].
In the case of surfaces, $T_{m} = 0$ and (5.10.3) gives us
$$
R^{2} = \displaystyle\frac{1}{4}H^{2} - K_{g} = \displaystyle\frac{1}{4}\(\displaystyle\frac{1}{R_{1}} +
\displaystyle\frac{1}{R_{2}}\)^{2} -\displaystyle\frac{1}{R_{1}R_{2}}
=\displaystyle\frac{1}{4}\(\displaystyle\frac{1}{R_{1}} - \displaystyle\frac{1}{R_{2}}\)^{2}.
$$
So, for surfaces the circle of normal curvatures and geodesic torsions is real.
If $T_{m} = 0$, the total torsion is of the form
\begin{equation}
T_{t} = -\displaystyle\frac{1}{4}\left(\displaystyle\frac{1}{R_{1}} -
\displaystyle\frac{1}{R_{2}}\right)^{2}.
\end{equation}
This formula was established by Alexandru Myller [35]. He proved
that $T_{t}$ is just ``the curvature of Sophy Germain and Emanuel
Bacaloglu'' ([35]).
\section{The nonholonomic plane and sphere}
\setcounter{theorem}{0}\setcounter{equation}{0}
The notions of nonholonomic plane and nonholonomic sphere have
been defined by Gr. Moisil who expressed the Pfaff equations, provided the existence of these manifolds, and R. Miron [64] studied them. Gh.
Vr\u{a}nceanu, in the book [44] has studied the notion of non
holonomic quadrics.
The Pfaff equation of the nonholonomic plane $\Pi_{3}^{2}$ given
by Gr. Moisil [30], [71] is as follows
\begin{equation}
\hspace*{14mm}\omega\equiv (q_{0}z - r_{0}y +a)dx +(r_{0}x - p_{0}z +b)dy +(p_{0}y -
q_{0}x +c)dz = 0,
\end{equation}
where $p_{0}, q_{0}, r_{0}, a,b,c$ are real constants verifying
the nonholonomy condition: $ap_{0} + bq_{0} + cr_{0}\neq 0$.
While, the nonholonomic sphere $\Sigma_{3}^{2}$ has the equation
given by Gr. Moisil [30]:
\begin{eqnarray}
\hspace*{5mm}\omega&\equiv & [2x (ax+by+cz) - a(x^{2} + y^{2}+z^{2}) + \mu x
+q_{0}z - r_{0}y + h] dx \nonumber\\&&+[2y(ax+by+cz)- b(x^{2} +
y^{2}+z^{2}) + \mu y +r_{0}x - p_{0}z + k]dy \nonumber\\&& +
[2z(ax+by+cz) - c(x^{2} + y^{2}+z^{2}) + \mu z +p_{0}y -
q_{0}x+l]dz.
\end{eqnarray}
$\mu, p_{0}, q_{0}, r_{0}, a,b,c, h,k,l$ being constants which
verify the condition $\omega\wedge d\omega \neq 0.$
In this section we investigate the geometrical properties of these
special manifolds.
1. {\it The nonholonomic plane $\Pi_{3}^{2}$}
The manifold $\Pi_{3}^{2}$ is defined by the property $\psi\equiv
0$, $\psi$ being the second fundamental form (5.6.3) of $E_{3}^{2}$.
From the formula (5.6.3) is follows
\begin{equation}
p_{2} = q_{1} = p_{1} - q_{2} = 0.
\end{equation}
The nonholonomic plane $\Pi_{3}^{2}$ has the invariants $H,
T_{m},\ldots$ given by
\begin{equation}
H=0, T_{m}\neq 0, T_{t} = \displaystyle\frac{1}{4}T_{m}^{2}, K_{t} = T_{t},
K_{g} = 0.
\end{equation}
Conversely, (5.11.4) imply (5.11.3).
Therefore, the properties (5.11.4), characterize the nonholonomic
plane $\Pi_{3}^{2}$.
Thus, the following theorems hold:
\begin{theorem}
1. The following line on $\Pi_{3}^{2}$ are nondetermined:
-The lines tangent to the principal directions.
- The lines tangent to the directions of extremal geodesic
torsion.
- The asymptotic lines.
\end{theorem}
\begin{theorem}
Also, we have
- The lines of curvature coincide with the minimal lines,
$\langle \delta \overline{r}, \delta \overline{r}\rangle=0$. The normal
curvature $\kappa_{n}\equiv 0.$
- The geodesic of $\Pi_{3}^{2}$ are straight lines.
\end{theorem}
Consequences: the nonholonomic manifold $\Pi_{3}^{2}$ is totally
geodesic.
Conversely, a total geodesic manifold $\Pi_{3}^{2}$ is a
nonholonomic plane.
If $T_{m} = 0,$ $\Pi_{3}^{2}$ is a family of ordinary planes from
$E_{3}^{2}.$
\medskip
2. {\it The nonholonomic sphere, $\Sigma_{3}^{2}$}.
A nonholonomic manifold $E_{3}^{2}$ is a nonholonomic sphere
$\Sigma_{3}^{2}$ if the second fundamental from $\psi$ is
proportional to the first fundamental form $\Phi.$
By means of (5.6.3) it follows that $\Sigma_{3}^{2}$ is
characterized by the following conditions
\begin{equation}
p_{1} - q_{2} = 0,\;\; p_{2} +q_{1} = 0.
\end{equation}
It follows that we have
\begin{equation}
\psi = \displaystyle\frac{1}{2}H\phi,\ \chi = \displaystyle\frac{1}{2}T_{m}\phi,\ \Theta^{2} =
K_{t}\phi.
\end{equation}
Thus, we can say
\begin{proposition}
1. The normal curvature $\kappa_{n}$, at a point $P\in \Sigma^{2}_{3}$
is the same $\(=\displaystyle\frac{1}{2}H\)$ in all tangent direction at point
$P$.
2. The geodesic torsion $\tau_{g}$ at point $P\in \Sigma_{3}^{2}$
is the same in all tangent direction at $P.$
\end{proposition}
\begin{proposition}
1. The principal directions at point $P\in E_{3}^{2}$ are
nondetermined.
2. The directions of extremal geodesic torsion at $P$ are non
determinated, too.
\end{proposition}
\begin{proposition}
At every point $P\in \Sigma_{3}^{2}$ the following relations hold:
\begin{equation}
K_{g} = \displaystyle\frac{1}{4}H^{2},\;\; T_{t} = \displaystyle\frac{1}{4}T_{m}^{2},\;\;
K_{t} = \displaystyle\frac{1}{4}(H^{2} +T_{m}^{2}),
\end{equation}
\end{proposition}
Conversely, if the first two relations (5.11.7) are verified at any
point $P\in E_{3}^{2}$, then $E_{3}^{2}$ is a nonholonomic sphere.
Indeed, (5.11.7) are the consequence of (5.11.5). Conversely, from
$K_{g} = \displaystyle\frac{1}{4}H^{2}$, $T_{t} = \displaystyle\frac{1}{4}T_{m}^{2}$ we
deduce (5.11.5).
\begin{proposition}
The nonholonomic sphere $\Sigma_{3}^{2}$ has the properties
\begin{equation}
K_{g}>0, T_{t} >0, K_{t} >0.
\end{equation}
\end{proposition}
\begin{proposition}
1. If $H\neq 0$, the asymptotic lines of $\Sigma_{3}^{2}$ are
imaginary.
2. The curvature lines of $\Sigma_{3}^{2}$, $(T_{m}\neq 0)$ are
imaginary.
\end{proposition}
The geodesic of $\Sigma_{3}^{2}$ are given by the equation (5.5.7)
or by the equation (5.5.11).
\begin{theorem}
The geodesics of nonholonomic sphere $\Sigma_{3}^{2}$ cannot be
the plane curves.
\end{theorem}
\begin{proof}
By means of theorem 5.5.1, along a geodesic $C$ we have $\kappa_{n} =
|\kappa|,$ $\tau_{g} = \tau$, $\kappa$ and $\tau$ being curvature and
torsion of $C$. From (11.6) we get
$$
\kappa_{n} = \displaystyle\frac{1}{2}H,\;\; \tau_{g} = \displaystyle\frac{1}{2}T_{m}.
$$
So, the torsion $\tau$ of $C$ is different from zero. It can not be
a plane curve.
We stop here the theory of nonholonomic manifolds $E_{3}^{2}$ in
the Euclidean space $E_{3}$, remarking that the nonholonomic quadric was investigated by G. Vr\u{a}nceanu and A. Dobrescu [8], [44], [45].
\end{proof}
\newpage
\thispagestyle{empty}
\chapter*{Preface}
In 2010, the Mathematical Seminar of the ``Alexandru Ioan Cuza''
University of Ia\c{s}i comes to its 100th anniversary of
prodigious existence.
The establishing of the Mathematical Seminar by Alexandru Myller
also marked the beginning of the School of Geometry in Ia\c{s}i,
developed in time by prestigious mathematicians of international
fame, Octav Mayer, Gheorghe Vr\^{a}nceanu, Grigore Moisil, Mendel
Haimovici, Ilie Popa, Dimitrie Mangeron, Gheorghe Gheorghiev and
many others.
Among the first paper works, published by Al. Myller and O. Mayer,
those concerning the generalization of the Levi-Civita parallelism
must be specified, because of the relevance for the international
recognition of the academic School of Ia\c{s}i. Later on, through
the effort of two generations of Ia\c{s}i based mathematicians,
these led to a body of theories in the area of differential
geometry of Euclidian, affine and projective spaces.
At the half-centenary of the Mathematical Seminary, in 1960, the
author of the present opuscule synthesized the field$'$s results
and laid it out in the form of a ``whole, superb theory'', as
mentioned by Al. Myller. In the same time period, the book {\it The
geometry of the Myller configurations} was published by the
Technical Publishing House. It represents, as mentioned by Octav
Mayer in the Foreword, ``the most precious tribute ever offered to
Alexandru Myller''. Nowadays, at the 100th anniversary of the
Mathematical Seminary ``Alexandru Myller'' and 150 years after the
enactment, made by Alexandru Ioan Cuza, to set up the University
that carries his name, we are going to pay homage to those two
historical acts in the Romanian culture, through the publishing,
in English, of the volume {\it The Geometry of the Myller
Configurations}, completed with an ample chapter, containing new
results of the Romanian geometers, regarding applications in the
studying of non-holonomic manifolds in the Euclidean space. On
this occasion, one can better notice the undeniable value of the
achievements in the field made by some great Romanian
mathematicians, such as Al. Myller, O. Mayer, Gh. Vr\u{a}nceanu,
Gr. Moisil, M. Haimovici, I. Popa, I. Creang\u{a} and Gh. Gheorghiev.
The initiative for the re-publishing of the book belongs to the
leadership of the ``Alexandru Ioan Cuza'' University of Ia\c{s}i, to the local branch of the Romanian Academy, as
well as NGO Formare Studia Ia\c{s}i. A competent
assistance was offered to me, in order to complete this work, by
my former students and present collaborators, Professors Mihai
Anastasiei and Ioan Buc\u{a}taru, to whom I express my utmost gratitude.
\vspace*{1cm}
\,\,\, Ia\c{s}i, 2010 \hfill Acad. Radu Miron
\cleardoublepage
\chapter*{Preface of the book {\it Geometria configura\c{t}iilor Myller}
written in 1966 by Octav Mayer, former member of the Romanian
Academy {\Large{(Translated from Romanian)}}}
One can bring to the great scientists, who passed away, various homages. Some times, the strong wind
of the progress wipes out the trace of their steps. To not forget them
means to continue their work, connecting them to the living
present. In this sense, this scientific work is the most precious homage
which can be dedicated to Alexandru Myller.
Initiator of a modern education of Mathematics at the University
of Iassy, founder of the Geometry School, which is still
flourishing today, in the third generation, Alexandru Myller was
also a hardworking researcher, well known inside the country and
also abroad due to his papers concerning Integral Equations and
Differential Geometry.
Our Academy elected him as a member and published his scientific
work. The ``A. Humboldt'' University from Berlin awarded him the
title of ``doctor Honoris Causa'' for ``special efforts in creating
an independent Romanian Mathematical School''.
Some of Alexandru Myller$'$s discoveries have penetrated fruitfully the
impetuous torrent of ideas, which have changed the Science of Geometry
during the first decades of the century. Among others, it is the case
of ``Myller configurations''. The reader would be perhaps interested
to find out some more details about how he discovered these configurations.
Let us imagine a surface $S$, on which there is drawn a curve $C$
and, along it, let us circumscribe to $S$ a developing surface
$\Sigma $; then we apply (develop) the surface $\Sigma$ on a
plane $\pi $. The points of the curve $C$, connected to the
respective tangent planes, which after are developed overlap each
other on the plane $\pi $, are going to represent a curve $C'$
into this plane.
In the same way, a series $(d)$ of tangent directions to the
surface $S$ at the points of the curve $C$ becomes developing, on
the plane $\pi $, a series $(d')$ of directions getting out from
the points of the curve $C'$. The directions $(d)$ are parallel on
the surface $S$ if the directions $(d')$ are parallel in the
common sense.
Going from this definition of T. Levi-Civita parallelism (valid in
the Euclidian space), Alexandru Myller arrived to a more general
concept in a sensible process of abstraction. Of course, it was
not possible to leave the curve $C$ aside, neither the
directions $(d)$ whose parallelism was going to be defined
in a more general sense. What was left aside was the surface $S$.
For the surface $\Sigma$ was considered, in a natural way, the enveloping
of the family of planes constructed from the points of the curve
$C$, planes in which are given the directions $(d)$. Keeping
unchanged the remainder of the definition one gets to what Alexandru
Myller called ``parallelism into a family of planes''.
A curve $C$, together with a family of planes on its points and a
family of given directions (in an arbitrary way) in these planes constitutes
what the author called a ``Myller configuration''.
It was considered that this new introduced notion had a central
place into the classical theory of surfaces, giving the
possibility to interpret and link among them many particular
facts. It was obvious that this notion can be successfully applied
in the Geometry of other spaces different from the Euclidian one.
Therefore the foreworded study worthed all the efforts made by the
author, a valuable mathematician from the third generation of the
Ia\c si Geometry School.
By recommending this work, we believe that the reader (who needs
only basic Differential Geometry) will be attracted by the clear
lecture of Radu Miron and also by the beauty of the subject, which
can still be developed further on.
\bigskip
\hfill Octav Mayer, 1966
\newpage
\chapter*{A short biography of\\ Al. Myller}
\begin{figure}[h]
\begin{center}
\includegraphics[width=6cm,height=7cm]{Myller.eps}
\end{center}
\end{figure}
{\it ALEXANDRU MYLLER} was born in Bucharest in 1879, and died in
Ia\c{s}i on the 4th of July 1965. Romanian mathematician. Honorary
Member (27 May 1938) and Honorific Member (12 August 1948) of the
Romanian Academy. High School and University studies (Faculty of
Science) in Bucharest, finished with a bachelor degree in
mathematics.
He was a Professor of Mathematics at the Pedagogical Seminar; he
sustained his PhD thesis {\it Gewohnliche Differential Gleichungen
Hoherer Ordnung}, in G\"{o}ttingen, under the scientific
supervision of David Hilbert.
He worked as a Professor at the Pedagogical Seminar and the School
of Post and Telegraphy (1907 - 1908), then as a lecturer at the
University of Bucharest (1908 - 1910). He is appointed Professor
of Analytical Geometry at the University of Ia\c{s}i (1910 -
1947); from 1947 onward - consultant Professor. In 1910, he sets
up the Mathematical Seminar at the ``Alexandru Ioan Cuza``
University of Ia\c{s}i, which he endows with a library full of
didactic works and specialty journals, and which, nowadays, bears
his name. Creator of the mathematical school of Ia\c{s}i, through
which many reputable Romanian mathematicians have passed, he
conveyed to us research works in fields such as integral
equations, differential geometry and history of mathematics. He
started his scientific activity with papers on the integral
equations theory, including extensions of Hilbert$'$s results, and
then he studied integral equations and self- adjoint of even and
odd order linear differential equations, being the first
mathematician to introduce integral equations with skew symmetric
kernel.
He was the first to apply integral equations to solve problems for
partial differential equations of hyperbolic type. He was also
interested in the differential geometry, discovering a
generalization of the notion of parallelism in the Levi-Civita
sense, and introducing the notion today known as concurrence in
the Myller sense. All these led to ``the differential geometry of
Myller configurations''(R. Miron, 1960). Al. Myller, Gh.
\c{T}i\c{t}eica and O. Mayer have created ``the differential
centro-affine geometry``, which the history of mathematics refers
to as ``purely Romanian creation$!$`` He was also concerned with
problems from the geometry of curves and surfaces in Euclidian
spaces. His research outcomes can be found in the numerous
memoirs, papers and studies, published in Ia\c{s}i and abroad:
Development of an arbitrary function after Bessel$'$s functions
(1909); Parallelism in the Levi-Civita sense in a plane system
(1924); The differential centro-affine geometry of the plane
curves (1933) etc. - within the volume ``Mathematical Writings``
(1959), Academy Publishing House. Alexandru Myller has been an
Honorary Member of the Romanian Academy since 1938. Also, he was a
Doctor Honoris Causa of the Humboldt University of Berlin.
Excerpt from the volume ``Members of the Romanian Academy 1866 -
1999. Dictionary.`` - Dorina N. Rusu, page 360.
\cleardoublepage
\section*{Abstract (Not appropriate in this style!)}%
\else \small
\begin{center}{\bf Abstract\vspace{-.5em}\vspace{\z@}}\end{center}%
\quotation
\fi
}%
}{%
}%
\@ifundefined{endabstract}{\def\endabstract
{\if@twocolumn\else\endquotation\fi}}{}%
\@ifundefined{maketitle}{\def\maketitle#1{}}{}%
\@ifundefined{affiliation}{\def\affiliation#1{}}{}%
\@ifundefined{proof}{\def\noindent\it Proof.\ \rm{\noindent{\bfseries Proof. }}}{}%
\@ifundefined{endproof}{\def\endproof{\mbox{\ \rule{.1in}{.1in}}}}{}%
\@ifundefined{newfield}{\def\newfield#1#2{}}{}%
\@ifundefined{chapter}{\def\chapter#1{\par(Chapter head:)#1\par }%
\newcount\c@chapter}{}%
\@ifundefined{part}{\def\part#1{\par(Part head:)#1\par }}{}%
\@ifundefined{section}{\def\section#1{\par(Section head:)#1\par }}{}%
\@ifundefined{subsection}{\def\subsection#1%
{\par(Subsection head:)#1\par }}{}%
\@ifundefined{subsubsection}{\def\subsubsection#1%
{\par(Subsubsection head:)#1\par }}{}%
\@ifundefined{paragraph}{\def\paragraph#1%
{\par(Subsubsubsection head:)#1\par }}{}%
\@ifundefined{subparagraph}{\def\subparagraph#1%
{\par(Subsubsubsubsection head:)#1\par }}{}%
\@ifundefined{therefore}{\def\therefore{}}{}%
\@ifundefined{backepsilon}{\def\backepsilon{}}{}%
\@ifundefined{yen}{\def\yen{\hbox{\rm\rlap=Y}}}{}%
\@ifundefined{registered}{%
\def\registered{\relax\ifmmode{}\r@gistered
\else$\m@th\r@gistered$\fi}%
\def\r@gistered{^{\ooalign
{\hfil\raise.07ex\hbox{$\scriptstyle\rm\RIfM@\expandafter\text@\else\expandafter\mbox\fi{R}$}\hfil\crcr
\mathhexbox20D}}}}{}%
\@ifundefined{Eth}{\def\Eth{}}{}%
\@ifundefined{eth}{\def\eth{}}{}%
\@ifundefined{Thorn}{\def\Thorn{}}{}%
\@ifundefined{thorn}{\def\thorn{}}{}%
\def\TEXTsymbol#1{\mbox{$#1$}}%
\@ifundefined{degree}{\def\degree{{}^{\circ}}}{}%
\newdimen\theight
\def\Column{%
\vadjust{\setbox\z@=\hbox{\scriptsize\quad\quad tcol}%
\theight=\ht\z@\advance\theight by \dp\z@\advance\theight by \lineskip
\kern -\theight \vbox to \theight{%
\rightline{\rlap{\box\z@}}%
\vss
}%
}%
}%
\def\qed{%
\ifhmode\unskip\nobreak\fi\ifmmode\ifinner\else\hskip5\partial@\fi\fi
\hbox{\hskip5\partial@\vrule width4\partial@ height6\partial@ depth1.5\partial@\hskip\partial@}%
}%
\def\cents{\hbox{\rm\rlap/c}}%
\def\miss{\hbox{\vrule height2\partial@ width 2\partial@ depth\z@}}%
\def\vvert{\Vert
\def\tcol#1{{\baselineskip=6\partial@ \vcenter{#1}} \Column} %
\def\dB{\hbox{{}}
\def\mB#1{\hbox{$#1$}
\def\nB#1{\hbox{#1}
\def\note{$^{\dag}}%
\defLaTeX2e{LaTeX2e}
\def\chkcompat{%
\if@compatibility
\else
\usepackage{latexsym}
\fi
}
\ifx\fmtnameLaTeX2e
\DeclareOldFontCommand{\rm}{\normalfont\rmfamily}{\mathrm}
\DeclareOldFontCommand{\sf}{\normalfont\sffamily}{\mathsf}
\DeclareOldFontCommand{\tt}{\normalfont\ttfamily}{\mathtt}
\DeclareOldFontCommand{\bf}{\normalfont\bfseries}{\mathbf}
\DeclareOldFontCommand{\it}{\normalfont\itshape}{\mathit}
\DeclareOldFontCommand{\sl}{\normalfont\slshape}{\@nomath\sl}
\DeclareOldFontCommand{\sc}{\normalfont\scshape}{\@nomath\sc}
\chkcompat
\fi
\def\alpha{{\Greekmath 010B}}%
\def\beta{{\Greekmath 010C}}%
\def\gamma{{\Greekmath 010D}}%
\def\delta{{\Greekmath 010E}}%
\def\epsilon{{\Greekmath 010F}}%
\def\zeta{{\Greekmath 0110}}%
\def\eta{{\Greekmath 0111}}%
\def\theta{{\Greekmath 0112}}%
\def\iota{{\Greekmath 0113}}%
\def\kappa{{\Greekmath 0114}}%
\def\lambda{{\Greekmath 0115}}%
\def\mu{{\Greekmath 0116}}%
\def\nu{{\Greekmath 0117}}%
\def\xi{{\Greekmath 0118}}%
\def\pi{{\Greekmath 0119}}%
\def\rho{{\Greekmath 011A}}%
\def\sigma{{\Greekmath 011B}}%
\def\tau{{\Greekmath 011C}}%
\def\upsilon{{\Greekmath 011D}}%
\def\phi{{\Greekmath 011E}}%
\def\chi{{\Greekmath 011F}}%
\def\psi{{\Greekmath 0120}}%
\def\omega{{\Greekmath 0121}}%
\def\varepsilon{{\Greekmath 0122}}%
\def\vartheta{{\Greekmath 0123}}%
\def\varpi{{\Greekmath 0124}}%
\def\varrho{{\Greekmath 0125}}%
\def\varsigma{{\Greekmath 0126}}%
\def\varphi{{\Greekmath 0127}}%
\def{\Greekmath 0272}{{\Greekmath 0272}}
\def\FindBoldGroup{%
{\setbox0=\hbox{$\mathbf{x\global\edef\theboldgroup{\the\mathgroup}}$}}%
}
\def\Greekmath#1#2#3#4{%
\if@compatibility
\ifnum\mathgroup=\symbold
\mathchoice{\mbox{\boldmath$\displaystyle\mathchar"#1#2#3#4$}}%
{\mbox{\boldmath$\textstyle\mathchar"#1#2#3#4$}}%
{\mbox{\boldmath$\scriptstyle\mathchar"#1#2#3#4$}}%
{\mbox{\boldmath$\scriptscriptstyle\mathchar"#1#2#3#4$}}%
\else
\mathchar"#1#2#3#4%
\fi
\else
\FindBoldGroup
\ifnum\mathgroup=\theboldgroup
\mathchoice{\mbox{\boldmath$\displaystyle\mathchar"#1#2#3#4$}}%
{\mbox{\boldmath$\textstyle\mathchar"#1#2#3#4$}}%
{\mbox{\boldmath$\scriptstyle\mathchar"#1#2#3#4$}}%
{\mbox{\boldmath$\scriptscriptstyle\mathchar"#1#2#3#4$}}%
\else
\mathchar"#1#2#3#4%
\fi
\fi}
\newif\ifGreekBold \GreekBoldfalse
\let\SAVEPBF=\pbf
\def\pbf{\GreekBoldtrue\SAVEPBF}%
\@ifundefined{theorem}{\newtheorem{theorem}{Theorem}}{}
\@ifundefined{lemma}{\newtheorem{lemma}[theorem]{Lemma}}{}
\@ifundefined{corollary}{\newtheorem{corollary}[theorem]{Corollary}}{}
\@ifundefined{conjecture}{\newtheorem{conjecture}[theorem]{Conjecture}}{}
\@ifundefined{proposition}{\newtheorem{proposition}{Proposition}}{}
\@ifundefined{axiom}{\newtheorem{axiom}{Axiom}}{}
\@ifundefined{remark}{\newtheorem{remark}{Remark}}{}
\@ifundefined{example}{\newtheorem{example}{Example}}{}
\@ifundefined{exercise}{\newtheorem{exercise}{Exercise}}{}
\@ifundefined{definition}{\newtheorem{definition}{Definition}}{}
\@ifundefined{mathletters}{%
\newcounter{equationnumber}
\def\mathletters{%
\addtocounter{equation}{1}
\edef\@currentlabel{\arabic{equation}}%
\setcounter{equationnumber}{\c@equation}
\setcounter{equation}{0}%
\edef\arabic{equation}{\@currentlabel\noexpand\alph{equation}}%
}
\def\endmathletters{%
\setcounter{equation}{\value{equationnumber}}%
}
}{}
\@ifundefined{BibTeX}{%
\def\BibTeX{{\rm B\kern-.05em{\sc i\kern-.025em b}\kern-.08em
T\kern-.1667em\lower.7ex\hbox{E}\kern-.125emX}}}{}%
\@ifundefined{AmS}%
{\def\AmS{{\protect\usefont{OMS}{cmsy}{m}{n}%
A\kern-.1667em\lower.5ex\hbox{M}\kern-.125emS}}}{}%
\@ifundefined{AmSTeX}{\def\AmSTeX{\protect\AmS-\protect\TeX\@}}{}%
\ifx\ds@amstex\relax
\message{amstex already loaded}\makeatother\endinpu
\else
\@ifpackageloaded{amstex}%
{\message{amstex already loaded}\makeatother\endinput}
{}
\@ifpackageloaded{amsgen}%
{\message{amsgen already loaded}\makeatother\endinput}
{}
\fi
\let\DOTSI\relax
\def\RIfM@{\relax\ifmmode}%
\def\FN@{\futurelet\next}%
\newcount\intno@
\def\iint{\DOTSI\intno@\tw@\FN@\ints@}%
\def\iiint{\DOTSI\intno@\thr@@\FN@\ints@}%
\def\iiiint{\DOTSI\intno@4 \FN@\ints@}%
\def\idotsint{\DOTSI\intno@\z@\FN@\ints@}%
\def\ints@{\findlimits@\ints@@}%
\newif\iflimtoken@
\newif\iflimits@
\def\findlimits@{\limtoken@true\ifx\next\limits\limits@true
\else\ifx\next\nolimits\limits@false\else
\limtoken@false\ifx\ilimits@\nolimits\limits@false\else
\ifinner\limits@false\else\limits@true\fi\fi\fi\fi}%
\def\multint@{\int\ifnum\intno@=\z@\intdots@
\else\intkern@\fi
\ifnum\intno@>\tw@\int\intkern@\fi
\ifnum\intno@>\thr@@\int\intkern@\fi
\int
\def\multintlimits@{\intop\ifnum\intno@=\z@\intdots@\else\intkern@\fi
\ifnum\intno@>\tw@\intop\intkern@\fi
\ifnum\intno@>\thr@@\intop\intkern@\fi\intop}%
\def\intic@{%
\mathchoice{\hskip.5em}{\hskip.4em}{\hskip.4em}{\hskip.4em}}%
\def\negintic@{\mathchoice
{\hskip-.5em}{\hskip-.4em}{\hskip-.4em}{\hskip-.4em}}%
\def\ints@@{\iflimtoken@
\def\ints@@@{\iflimits@\negintic@
\mathop{\intic@\multintlimits@}\limits
\else\multint@\nolimits\fi
\eat@
\else
\def\ints@@@{\iflimits@\negintic@
\mathop{\intic@\multintlimits@}\limits\else
\multint@\nolimits\fi}\fi\ints@@@}%
\def\intkern@{\mathchoice{\!\!\!}{\!\!}{\!\!}{\!\!}}%
\def\plaincdots@{\mathinner{\cdotp\cdotp\cdotp}}%
\def\intdots@{\mathchoice{\plaincdots@}%
{{\cdotp}\mkern1.5mu{\cdotp}\mkern1.5mu{\cdotp}}%
{{\cdotp}\mkern1mu{\cdotp}\mkern1mu{\cdotp}}%
{{\cdotp}\mkern1mu{\cdotp}\mkern1mu{\cdotp}}}%
\def\RIfM@{\relax\protect\ifmmode}
\def\RIfM@\expandafter\text@\else\expandafter\mbox\fi{\RIfM@\expandafter\RIfM@\expandafter\text@\else\expandafter\mbox\fi@\else\expandafter\mbox\fi}
\let\nfss@text\RIfM@\expandafter\text@\else\expandafter\mbox\fi
\def\RIfM@\expandafter\text@\else\expandafter\mbox\fi@#1{\mathchoice
{\textdef@\displaystyle\f@size{#1}}%
{\textdef@\textstyle\tf@size{\firstchoice@false #1}}%
{\textdef@\textstyle\sf@size{\firstchoice@false #1}}%
{\textdef@\textstyle \ssf@size{\firstchoice@false #1}}%
\glb@settings}
\def\textdef@#1#2#3{\hbox{{%
\everymath{#1}%
\let\f@size#2\selectfont
#3}}}
\newif\iffirstchoice@
\firstchoice@true
\def\Let@{\relax\iffalse{\fi\let\\=\cr\iffalse}\fi}%
\def\vspace@{\def\vspace##1{\crcr\noalign{\vskip##1\relax}}}%
\def\multilimits@{\bgroup\vspace@\Let@
\baselineskip\fontdimen10 \scriptfont\tw@
\advance\baselineskip\fontdimen12 \scriptfont\tw@
\lineskip\thr@@\fontdimen8 \scriptfont\thr@@
\lineskiplimit\lineskip
\vbox\bgroup\ialign\bgroup\hfil$\m@th\scriptstyle{##}$\hfil\crcr}%
\def\Sb{_\multilimits@}%
\def\endSb{\crcr\egroup\egroup\egroup}%
\def\Sp{^\multilimits@}%
\let\endSp\endSb
\newdimen\ex@
\[email protected]
\def\rightarrowfill@#1{$#1\m@th\mathord-\mkern-6mu\cleaders
\hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill
\mkern-6mu\mathord\rightarrow$}%
\def\leftarrowfill@#1{$#1\m@th\mathord\leftarrow\mkern-6mu\cleaders
\hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill\mkern-6mu\mathord-$}%
\def\leftrightarrowfill@#1{$#1\m@th\mathord\leftarrow
\mkern-6mu\cleaders
\hbox{$#1\mkern-2mu\mathord-\mkern-2mu$}\hfill
\mkern-6mu\mathord\rightarrow$}%
\def\overrightarrow{\mathpalette\overrightarrow@}%
\def\overrightarrow@#1#2{\vbox{\ialign{##\crcr\rightarrowfill@#1\crcr
\noalign{\kern-\ex@\nointerlineskip}$\m@th\hfil#1#2\hfil$\crcr}}}%
\let\overarrow\overrightarrow
\def\overleftarrow{\mathpalette\overleftarrow@}%
\def\overleftarrow@#1#2{\vbox{\ialign{##\crcr\leftarrowfill@#1\crcr
\noalign{\kern-\ex@\nointerlineskip}$\m@th\hfil#1#2\hfil$\crcr}}}%
\def\overleftrightarrow{\mathpalette\overleftrightarrow@}%
\def\overleftrightarrow@#1#2{\vbox{\ialign{##\crcr
\leftrightarrowfill@#1\crcr
\noalign{\kern-\ex@\nointerlineskip}$\m@th\hfil#1#2\hfil$\crcr}}}%
\def\underrightarrow{\mathpalette\underrightarrow@}%
\def\underrightarrow@#1#2{\vtop{\ialign{##\crcr$\m@th\hfil#1#2\hfil
$\crcr\noalign{\nointerlineskip}\rightarrowfill@#1\crcr}}}%
\let\underarrow\underrightarrow
\def\underleftarrow{\mathpalette\underleftarrow@}%
\def\underleftarrow@#1#2{\vtop{\ialign{##\crcr$\m@th\hfil#1#2\hfil
$\crcr\noalign{\nointerlineskip}\leftarrowfill@#1\crcr}}}%
\def\underleftrightarrow{\mathpalette\underleftrightarrow@}%
\def\underleftrightarrow@#1#2{\vtop{\ialign{##\crcr$\m@th
\hfil#1#2\hfil$\crcr
\noalign{\nointerlineskip}\leftrightarrowfill@#1\crcr}}}%
\def\qopnamewl@#1{\mathop{\operator@font#1}\nlimits@}
\let\nlimits@\displaylimits
\def\setboxz@h{\setbox\z@\hbox}
\def\varlim@#1#2{\mathop{\vtop{\ialign{##\crcr
\hfil$#1\m@th\operator@font lim$\hfil\crcr
\noalign{\nointerlineskip}#2#1\crcr
\noalign{\nointerlineskip\kern-\ex@}\crcr}}}}
\def\rightarrowfill@#1{\m@th\setboxz@h{$#1-$}\ht\z@\z@
$#1\copy\z@\mkern-6mu\cleaders
\hbox{$#1\mkern-2mu\box\z@\mkern-2mu$}\hfill
\mkern-6mu\mathord\rightarrow$}
\def\leftarrowfill@#1{\m@th\setboxz@h{$#1-$}\ht\z@\z@
$#1\mathord\leftarrow\mkern-6mu\cleaders
\hbox{$#1\mkern-2mu\copy\z@\mkern-2mu$}\hfill
\mkern-6mu\box\z@$}
\def\qopnamewl@{proj\,lim}{\qopnamewl@{proj\,lim}}
\def\qopnamewl@{inj\,lim}{\qopnamewl@{inj\,lim}}
\def\mathpalette\varlim@\rightarrowfill@{\mathpalette\varlim@\rightarrowfill@}
\def\mathpalette\varlim@\leftarrowfill@{\mathpalette\varlim@\leftarrowfill@}
\def\mathpalette\varliminf@{}{\mathpalette\mathpalette\varliminf@{}@{}}
\def\mathpalette\varliminf@{}@#1{\mathop{\underline{\vrule\@depth.2\ex@\@width\z@
\hbox{$#1\m@th\operator@font lim$}}}}
\def\mathpalette\varlimsup@{}{\mathpalette\mathpalette\varlimsup@{}@{}}
\def\mathpalette\varlimsup@{}@#1{\mathop{\overline
{\hbox{$#1\m@th\operator@font lim$}}}}
\def\tfrac#1#2{{\textstyle {#1 \over #2}}}%
\def\displaystyle\frac#1#2{{\displaystyle {#1 \over #2}}}%
\def\binom#1#2{{#1 \choose #2}}%
\def\tbinom#1#2{{\textstyle {#1 \choose #2}}}%
\def\dbinom#1#2{{\displaystyle {#1 \choose #2}}}%
\def\QATOP#1#2{{#1 \atop #2}}%
\def\QTATOP#1#2{{\textstyle {#1 \atop #2}}}%
\def\QDATOP#1#2{{\displaystyle {#1 \atop #2}}}%
\def\QABOVE#1#2#3{{#2 \above#1 #3}}%
\def\QTABOVE#1#2#3{{\textstyle {#2 \above#1 #3}}}%
\def\QDABOVE#1#2#3{{\displaystyle {#2 \above#1 #3}}}%
\def\QOVERD#1#2#3#4{{#3 \overwithdelims#1#2 #4}}%
\def\QTOVERD#1#2#3#4{{\textstyle {#3 \overwithdelims#1#2 #4}}}%
\def\QDOVERD#1#2#3#4{{\displaystyle {#3 \overwithdelims#1#2 #4}}}%
\def\QATOPD#1#2#3#4{{#3 \atopwithdelims#1#2 #4}}%
\def\QTATOPD#1#2#3#4{{\textstyle {#3 \atopwithdelims#1#2 #4}}}%
\def\QDATOPD#1#2#3#4{{\displaystyle {#3 \atopwithdelims#1#2 #4}}}%
\def\QABOVED#1#2#3#4#5{{#4 \abovewithdelims#1#2#3 #5}}%
\def\QTABOVED#1#2#3#4#5{{\textstyle
{#4 \abovewithdelims#1#2#3 #5}}}%
\def\QDABOVED#1#2#3#4#5{{\displaystyle
{#4 \abovewithdelims#1#2#3 #5}}}%
\def\tint{\mathop{\textstyle \int}}%
\def\tiint{\mathop{\textstyle \iint }}%
\def\tiiint{\mathop{\textstyle \iiint }}%
\def\tiiiint{\mathop{\textstyle \iiiint }}%
\def\tidotsint{\mathop{\textstyle \idotsint }}%
\def\toint{\mathop{\textstyle \oint}}%
\def\tsum{\mathop{\textstyle \sum }}%
\def\tprod{\mathop{\textstyle \prod }}%
\def\tbigcap{\mathop{\textstyle \bigcap }}%
\def\tbigwedge{\mathop{\textstyle \bigwedge }}%
\def\tbigoplus{\mathop{\textstyle \bigoplus }}%
\def\tbigodot{\mathop{\textstyle \bigodot }}%
\def\tbigsqcup{\mathop{\textstyle \bigsqcup }}%
\def\tcoprod{\mathop{\textstyle \coprod }}%
\def\tbigcup{\mathop{\textstyle \bigcup }}%
\def\tbigvee{\mathop{\textstyle \bigvee }}%
\def\tbigotimes{\mathop{\textstyle \bigotimes }}%
\def\tbiguplus{\mathop{\textstyle \biguplus }}%
\def\dint{\mathop{\displaystyle \int}}%
\def\diint{\mathop{\displaystyle \iint }}%
\def\diiint{\mathop{\displaystyle \iiint }}%
\def\diiiint{\mathop{\displaystyle \iiiint }}%
\def\didotsint{\mathop{\displaystyle \idotsint }}%
\def\doint{\mathop{\displaystyle \oint}}%
\def\dsum{\mathop{\displaystyle \sum }}%
\def\dprod{\mathop{\displaystyle \prod }}%
\def\dbigcap{\mathop{\displaystyle \bigcap }}%
\def\dbigwedge{\mathop{\displaystyle \bigwedge }}%
\def\dbigoplus{\mathop{\displaystyle \bigoplus }}%
\def\dbigodot{\mathop{\displaystyle \bigodot }}%
\def\dbigsqcup{\mathop{\displaystyle \bigsqcup }}%
\def\dcoprod{\mathop{\displaystyle \coprod }}%
\def\dbigcup{\mathop{\displaystyle \bigcup }}%
\def\dbigvee{\mathop{\displaystyle \bigvee }}%
\def\dbigotimes{\mathop{\displaystyle \bigotimes }}%
\def\dbiguplus{\mathop{\displaystyle \biguplus }}%
\def\stackunder#1#2{\mathrel{\mathop{#2}\limits_{#1}}}%
\begingroup \catcode `|=0 \catcode `[= 1
\catcode`]=2 \catcode `\{=12 \catcode `\}=12
\catcode`\\=12
|gdef|@alignverbatim#1\end{align}[#1|end[align]]
|gdef|@salignverbatim#1\end{align*}[#1|end[align*]]
|gdef|@alignatverbatim#1\end{alignat}[#1|end[alignat]]
|gdef|@salignatverbatim#1\end{alignat*}[#1|end[alignat*]]
|gdef|@xalignatverbatim#1\end{xalignat}[#1|end[xalignat]]
|gdef|@sxalignatverbatim#1\end{xalignat*}[#1|end[xalignat*]]
|gdef|@gatherverbatim#1\end{gather}[#1|end[gather]]
|gdef|@sgatherverbatim#1\end{gather*}[#1|end[gather*]]
|gdef|@gatherverbatim#1\end{gather}[#1|end[gather]]
|gdef|@sgatherverbatim#1\end{gather*}[#1|end[gather*]]
|gdef|@multilineverbatim#1\end{multiline}[#1|end[multiline]]
|gdef|@smultilineverbatim#1\end{multiline*}[#1|end[multiline*]]
|gdef|@arraxverbatim#1\end{arrax}[#1|end[arrax]]
|gdef|@sarraxverbatim#1\end{arrax*}[#1|end[arrax*]]
|gdef|@tabulaxverbatim#1\end{tabulax}[#1|end[tabulax]]
|gdef|@stabulaxverbatim#1\end{tabulax*}[#1|end[tabulax*]]
|endgroup
\def\align{\@verbatim \frenchspacing\@vobeyspaces \@alignverbatim
You are using the "align" environment in a style in which it is not defined.}
\let\endalign=\endtrivlist
\@namedef{align*}{\@verbatim\@salignverbatim
You are using the "align*" environment in a style in which it is not defined.}
\expandafter\let\csname endalign*\endcsname =\endtrivlist
\def\alignat{\@verbatim \frenchspacing\@vobeyspaces \@alignatverbatim
You are using the "alignat" environment in a style in which it is not defined.}
\let\endalignat=\endtrivlist
\@namedef{alignat*}{\@verbatim\@salignatverbatim
You are using the "alignat*" environment in a style in which it is not defined.}
\expandafter\let\csname endalignat*\endcsname =\endtrivlist
\def\xalignat{\@verbatim \frenchspacing\@vobeyspaces \@xalignatverbatim
You are using the "xalignat" environment in a style in which it is not defined.}
\let\endxalignat=\endtrivlist
\@namedef{xalignat*}{\@verbatim\@sxalignatverbatim
You are using the "xalignat*" environment in a style in which it is not defined.}
\expandafter\let\csname endxalignat*\endcsname =\endtrivlist
\def\gather{\@verbatim \frenchspacing\@vobeyspaces \@gatherverbatim
You are using the "gather" environment in a style in which it is not defined.}
\let\endgather=\endtrivlist
\@namedef{gather*}{\@verbatim\@sgatherverbatim
You are using the "gather*" environment in a style in which it is not defined.}
\expandafter\let\csname endgather*\endcsname =\endtrivlist
\def\multiline{\@verbatim \frenchspacing\@vobeyspaces \@multilineverbatim
You are using the "multiline" environment in a style in which it is not defined.}
\let\endmultiline=\endtrivlist
\@namedef{multiline*}{\@verbatim\@smultilineverbatim
You are using the "multiline*" environment in a style in which it is not defined.}
\expandafter\let\csname endmultiline*\endcsname =\endtrivlist
\def\arrax{\@verbatim \frenchspacing\@vobeyspaces \@arraxverbatim
You are using a type of "array" construct that is only allowed in AmS-LaTeX.}
\let\endarrax=\endtrivlist
\def\tabulax{\@verbatim \frenchspacing\@vobeyspaces \@tabulaxverbatim
You are using a type of "tabular" construct that is only allowed in AmS-LaTeX.}
\let\endtabulax=\endtrivlist
\@namedef{arrax*}{\@verbatim\@sarraxverbatim
You are using a type of "array*" construct that is only allowed in AmS-LaTeX.}
\expandafter\let\csname endarrax*\endcsname =\endtrivlist
\@namedef{tabulax*}{\@verbatim\@stabulaxverbatim
You are using a type of "tabular*" construct that is only allowed in AmS-LaTeX.}
\expandafter\let\csname endtabulax*\endcsname =\endtrivlist
\def\@@eqncr{\let\@tempa\relax
\ifcase\@eqcnt \def\@tempa{& & &}\or \def\@tempa{& &}%
\else \def\@tempa{&}\fi
\@tempa
\if@eqnsw
\iftag@
\@taggnum
\else
\@eqnnum\stepcounter{equation}%
\fi
\fi
\global\@ifnextchar*{\@tagstar}{\@tag}@false
\global\@eqnswtrue
\global\@eqcnt\z@\cr}
\def\endequation{%
\ifmmode\ifinner
\iftag@
\addtocounter{equation}{-1}
$\hfil
\displaywidth\linewidth\@taggnum\egroup \endtrivlist
\global\@ifnextchar*{\@tagstar}{\@tag}@false
\global\@ignoretrue
\else
$\hfil
\displaywidth\linewidth\@eqnnum\egroup \endtrivlist
\global\@ifnextchar*{\@tagstar}{\@tag}@false
\global\@ignoretrue
\fi
\else
\iftag@
\addtocounter{equation}{-1}
\eqno \hbox{\@taggnum}
\global\@ifnextchar*{\@tagstar}{\@tag}@false%
$$\global\@ignoretrue
\else
\eqno \hbox{\@eqnnum
$$\global\@ignoretrue
\fi
\fi\fi
}
\newif\iftag@ \@ifnextchar*{\@tagstar}{\@tag}@false
\def\@ifnextchar*{\@tagstar}{\@tag}{\@ifnextchar*{\@tagstar}{\@tag}}
\def\@tag#1{%
\global\@ifnextchar*{\@tagstar}{\@tag}@true
\global\def\@taggnum{(#1)}}
\def\@tagstar*#1{%
\global\@ifnextchar*{\@tagstar}{\@tag}@true
\global\def\@taggnum{#1}%
}
\makeatother
\endinput
|
\section{Introduction}
\label{sec:intro}
The study of molecular biophysics began with Fr$\ddot{\mathrm{o}}$lich's
hypotheses which assumes that an excitation from an atomic vibration related to
biological activity \cite{takeno}. Based on this idea, Davydov has developed a
quantum theory of protein to understand the mechanisms of energy transport in
molecular protein, in particular alpha helix protein. The mechanism is an
excitation energy of an amide-I is stabilized by its vibration in a combined
excitation which propagate as a soliton \cite{scott,scott2}. Most of studies in
this field have been done at zero temperature, and little attention has been
given at physiological temperatures.
The studies of Davydov soliton in physiological temperature is realized by the Davydov soliton interacting with thermal bath. While zero temperature calculation allows the existence of a soliton states in the proteins, there is an important question whether these states are stable at biological temperatures \cite{cruzeiro1,cruzeiro2}. Because the temperature effect may exchange heat energy with surrounding aqueous medium. The measurement of infrared absorption and Raman's scattering of an crystallineacetanilide $(CH_3CONHC_6H_5)_x$ at low temperature
showed a new band closing to the amide-I band \cite{careri1,careri2}. The result is interpreted as a signature of Davydov's soliton. The experiment using femtosecond IR spectroscopy realizing a band of the amide-I from acetanilide (ACN) and N-methylacetanide (NMA) shows the dependencies of absorption spectrum on temperature. At high temperature, the absorption spectrum shifts to higher frequency \cite{edler2}. Theoretical prediction using first order perturbation methods based on Davydov model gives soliton life time at the order of $O(10^{-12})$ s. This is very short time for biological processes in room temperature \cite{cottingham}.
Using standard Davydov model, some numerical calculations showed that soliton is stable at 310K \cite{cruzeiro1}. The calculation based on trial function by Kapor et.al. showed that soliton is stable at 300K \cite{kapor1}. The result has also been confirmed in \cite{kapor2}.
From those results, it is important to study Davydov model involving the contact
with thermal bath. The behavior of the system has attracted many interests in
the last three decades. The interaction of a system with its environment is
given by the dissipation effect in quantum system \cite{weiss,rigo,perceval}.
However, the dissipation effect leads to a serious problem for quantization
procedure due to the broken Heisenberg's relation. The most appropriate theory
to resolve this problem is the Linblad formulation of master equation
\cite{rigo,perceval}.
The first application of Linblad formulation of master equation to the protein
model has been done by Cuevas et.al.
\cite{cuevas}. They used Davydov-Scott monomer model and showed that at 10K the
quantum effect of amide-I vibration can not be neglected. Also at room
temperature the
semi classical approach might be a good approximation compared to the
corresponding full quantum system. However the study was focused on the
dynamical aspect of the system. Since the real world is affected by thermal
fluctuation, it is important to study such system from statistical mechanics
point of view.
In this paper the thermodynamic properties of Davydov-Scott
monomer is investigated using Lindblad formulation and the
partition function is calculated using path integral method
\cite{feynmann}. The path integral method is a powerful tool to
investigate the properties of nonlinear dynamical systems with
retarded interactions \cite{mazoli}. In Sec. \ref{sec:hdm} the
Hamiltonian of the system under consideration is described as a
coupled harmonics oscillation \ie the amide-I and
amide-site oscillators. In addition to the original Davydov-Scott monomer, the
higher order of amide-site displacement inducing the anharmonic oscillation
effect of amide-site is also taken into account. This is motivated by the fact
that the excitation and relaxation of collective modes of a protein are
generally achieved via anharmonic interactions with other normal modes through
energy exchange \cite{xie}. In Sec. \ref{sec:pi} the partition function is
calculated using path integral approach to obtain the thermodynamic specific
heat of the system which is then discussed in Sec. \ref{sec:tpdc}.
\section{Hamiltonian of Davydov-Scott monomer with anharmonic oscillation effect
in Lindblad formulation of master equation}
\label{sec:hdm}
Considering the Davydov-Scott monomer, the excitation of the amide-I is
described by the coordinate ($x$) and momentum ($p$) operators. On the other
hand, the displacement and momentum operators of amide-site are expressed by $Q$
and $P$. The hamiltonian for Davydov-Scott monomer can then be written in the
full quantum approach as
follow,
\be
H = \frac{p^2}{2 m} + \frac{1}{2} m \o^2 x^2 - \frac{1}{4} \delta x^4 +
\frac{P^2}{2M} + V(Q)+ \chi \frac{\pd V}{\pd Q} x \; ,
\label{eq:mono1}
\ee
where $\o$ is the intrinsic frequency of amide-I oscilation, $\chi$ measures
the coupling between the amide-I excitation and the amide site vibration, $m$
($M$) is the amide-I (amide-site) mass and $\delta$ is the anharmonic
coefficient. Here the amide-site potential can generally be expanded around the
origin,
\bea
V(Q) & = & V_0 + \left. \frac{\pd V}{\pd Q} \right|_{Q=0} Q + \left.
\frac{1}{2}\frac{\pd^2 V}{\pd Q^2} \right|_{Q=0} Q^2
\nonumber\\
& & + \left. \frac{1}{6}\frac{\pd^3 V}{\pd Q^3} \right|_{Q=0} Q^3
+ \left. \frac{1}{24}\frac{\pd^4 V}{\pd Q^4} \right|_{Q=0} Q^4 + \cdots \;
.
\label{eq:poten1}
\eea
The first term in Eq. (\ref{eq:poten1}) is a constant and can be scaled out. On
the other hand, the equilibrium condition implies that the term of
$Q$ vanishes due to $-\nabla V(Q)|_{Q=0} = 0$. The third term is the harmonic
oscillator one, while the remaining higher order terms represent the anharmonic
contributions.
The original Davydov-Scott model assumes that the anharmonicity of amide-site is
not important compared to the amide-I \cite{pouthier}. The nonlinear effects
appear only through exciton and amide-site coupling and higher order potential
of excitons. However, recent experiment in biopolymer showed that the
anharmonicity of polymer amide-site might be important \cite{go,xie,petr}. The
solid state experiments showed that
at high temperature (room temperature), specific heat is not a constant
anymore, but it is increasing. This fact can only be explained by taking into
account the anharmonic oscillator term \cite{kittel}.
In Eq. (\ref{eq:poten1}), the terms $Q^3, Q^5, \cdots$ represent
the asymmetry of atom mutual repulsion, while the terms $Q^4, Q^6,
\cdots$ represent the softening of vibration at large amplitudes
\cite{kittel}. Moreover, concerning the fact that the Davydov-Scott polymer has
an identic monomer, it is plausible to assume that the asymmetry of atom
mutual repulsion can be ignored. This yields,
\be
V(Q) = \frac{1}{2} \, \kappa Q^2 - \frac{1}{4} \lambda \, Q^4 \; ,
\label{eq:poten2}
\ee
where $\kappa/2 = 1/2 \, ({\pd^2 V}/{\pd Q^2})|_{Q=0}$ and $1/{24} \, ({\pd^4
V}/{\pd Q^4})|_{Q=0} = -\lambda/4$. Again, the second term induces the
anharmonic oscillation effect in the system.
Unfortunately the term leads to a severe problem of describing
the damping in open quantum systems, and it has been discussed for a long time.
One of the known models dealing with this problem is the one dimensional damped
harmonic oscillator (known as the Caldirola-Kanai Hamiltonian) \cite{um}. In
this model the momentum and coordinate operators are multiplied by $e^{-\gamma
t}$ and $e^{\gamma t}$, where $\gamma$ is a damping factor.
In statistical mechanics, the behavior of an open system within system-plus-bath
can be modeled by the density matrix formalism $\rho$. The equation of density
matrix with hamiltonian $H$ and environment operator $R$ satisfy particular
master equation \cite{weiss}. It is usually restricted to weak system-bath
interaction. The density operator gives the probability for the expected
outcomes of measurements on the system. However, this formulation does not
preserve density operator properties, that is hermiticity, unit trace, and
positivity. The open quantum system theory which preserves density matrix
properties can be realized using Lindblad formulation. In this theory, the
interaction between Hamiltonian and thermal bath is realized by introducing some
operators called Lindblad operators. The operators obey the master equation
\cite{lindblad},
\be
\frac{\pd \rho}{\pd t} =
-\frac{i}{\hbar}[H,\rho]+\frac{i}{2\hbar}
\sum_{j} \left( [L_j, \rho L_j^\dag] + [L_j, \rho L_j^\dag] \right) \; ,\
\label{eq:lindblad}
\ee
where $L_j$ are the Lindblad operators. This choice is not unique and not
necessarily Hermitian. Since $L$ must be the first order in $Q$ and $P$
\cite{perceval}. The operators $H$ and $L$ denote the internal dynamics and
environmental effects of the system. Throughout the paper, the Lindblad
operators are put,
\bea
L_1 &=& \sqrt{\gamma(1+\nu)}
\nonumber \\
&& \times \left( \sqrt{\frac{M \Omega}{2\hbar}}Q+i\sqrt{\frac{1}{2M\hbar
\Omega}}P
+ \frac{\chi}{\hbar \Omega} x \right) \; ,
\label{eq:operator1}\\
L_2 &=& \sqrt{\gamma\nu}
\left( \sqrt{\frac{M\Omega}{2\hbar}}Q-i\sqrt{\frac{1}{2M \hbar \Omega}}P
+ \frac{\chi}{\hbar \Omega} x \right) \; ,\
\label{eq:operator2}
\eea
where $\gamma$ is a damping parameter related to the intensity of thermal bath,
$\Omega=\sqrt{\kappa/M}$, $k_B$ is the Boltzmann constant, $T$ is temperature
and $\nu = (e^{\hbar \Omega/k_BT}-1)^{-1}$ is the Bose-Einstein distribution.
Substituting Eqs. (\ref{eq:operator1}) and (\ref{eq:operator2}) into Lindblad
master equation in Eq. (\ref{eq:lindblad}),
\bea
\frac{\pd \rho}{\pd t} & = &
-\frac{i}{\hbar}[H,\rho]
-\frac{i\delta_1}{2\hbar}[Q,P\rho+\rho P]
\nonumber \\
&& - \frac{i \delta_2}{2\hbar}[x,P\rho+\rho P]
-\frac{\delta_3}{2\hbar} \left[ Q,[Q,\rho] \right]
\nonumber \\
&&
-\frac{\delta_4}{2\hbar} \left[x,[Q,\rho] \right]
-\frac{\delta_5}{2\hbar} \left[x,[x,\rho] \right]
-\frac{\delta_6}{2\hbar} \left[P,[P,\rho] \right] \; ,\
\label{eq:lindblad22}
\eea
where $\delta_i$'s are the coefficient in the Lindblad operators,
\bea
\delta_1 & = & \frac{\gamma}{2 \hbar} (1 + 2 \nu) \; , \\
\delta_2 & = & \frac{\gamma\chi}{4 \hbar \Omega} \sqrt{\frac{\omega}{\Omega}}
(1 + 2 \nu) \; , \\
\delta_3 & = & \frac{\gamma M \Omega}{2 \hbar}(1+2 \nu) \; , \\
\delta_4 & = & \frac{\gamma\chi}{2 \hbar \Omega} \sqrt{M m \Omega \omega}
(1+2 \nu) \; , \\
\delta_5 & = & \frac{\gamma\chi^2 m \omega}{4 \hbar \Omega^2} (1 + 2 \nu) \;
, \\
\delta_6 & = & \frac{\gamma}{2 \hbar M \Omega} (1 + 2 \nu) \; .
\label{eq:deltanya}
\eea
$\delta_{1,2}$ are the frictional damping rate, while
$\delta_{3,4,5,6}$ are the quantum mechanical diffusion coefficients
\cite{palchikov}. This is the underlying model in the paper.
\section{Path integral calculation of partition function}
\label{sec:pi}
In order to solve the master equation in Eq. (\ref{eq:lindblad22}), the
resolution of a set of differential equations among the matrix elements of
density operator with respect to a specific basis is required. This is usually
provided by the eigen states of the system hamiltonian \cite{nakazato}. But in
the statistical mechanics it is not necessary to solve the master equation.
Instead one can calculate the partition function $\z$ using, for instance,
path integral method.
\subsection{Partition function}
The density matrix can be obtained by performing a transformation $t \rightarrow
\tau = -it = \beta \hbar$ with $\beta = 1/(k_B T)$ \cite{feynmann}.
We assume that the quantum mechanical diffusion is dominant than
the frictional damping rate such that it can be ignored. Then, the Lindblad
master equation of the unnormalized $\rho$ can be rewritten as,
\be
-\frac{\pd \rho}{\pd \beta}= H \rho + \triangle \rho \; ,
\label{eq:thermo1}
\ee
where
\be
\triangle \rho =
\frac{i\delta_3 }{2\hbar} Q^2 \rho + \frac{i\delta_4}{2\hbar} xQ\rho
+ \frac{i\delta_5}{2\hbar} x^2 \rho + \frac{i\delta_6}{2\hbar} P^2 \rho \; .
\label{eq:difusi}
\ee
It should be remarked here that, using this equation one can confirm the
Linblad operators in Eqs. (\ref{eq:operator1}) and (\ref{eq:operator2}) lead
to the right equilibrium. The proof is given in App. \ref{app:es}.
The partition function corresponding to the master equation is given by
\cite{feynmann},
\be
\z = \int \id [x(t)] e^{-\frac{1}{\hbar}S(x(t))} \; ,
\label{eq:par}
\ee
where $S(x(t)) = \int_0^{\beta \hbar}(T+V) \d t'$ is the Euclidean action
corresponding to the equation.
Our interest is on the anharmonic amide-site interactions. Choosing the
amide-site potential as Eq. (\ref{eq:poten2}) and
taking into account only the first order of potential in the
interaction, the action becomes \cite{feynmann},
\bea
\s_{xQ} & = & \int_0^\tau \d t \left(
\frac{1}{2} m \dot{x}^2 +\frac{1}{2}m \omega^2 x^2 - \frac{1}{2}\delta x^4 +
\chi x Q
\right. \nonumber \\
&& \left.
+ \frac{1}{2}M \dot{Q}^2 + \frac{1}{2} \kappa Q^2 - \frac{i\delta_3}{2}
Q^2- \frac{1}{4}\lambda Q^4
\right.
\nonumber\\
&& \left. - \frac{i \delta_4}{2} x Q
-\frac{i\delta_5}{2}x^2 - \frac{i\delta_6}{2}\dot{Q}^2 \right) \; .
\label{eq:inter1}
\eea
This yields,
\bea
\s_{xQ} & = & \int_0^\tau \d t \left( \frac{1}{2} m \dot{x}^2 + \frac{1}{2}
\overline{k} x^2
- \frac{1}{2} \delta x^4 + \tilde{\chi} x Q \right. \nonumber \\
&& \left.
+ \frac{1}{2} \overline{M} \dot{Q}^2 + \frac{1}{2} \tilde{\kappa} Q^2
- \frac{1}{4}\lambda Q^4 \right) \; ,
\label{eq:inter3}
\eea
where $\tilde{\kappa}=\kappa-i\delta_3$ ,
$\overline{k}=k-i\delta_5$, $\overline{M}=M-i\delta_6$,
$\tilde{\chi}=\chi-i/2\delta_4$ and $k=m\omega^2$. The amide-site
is assumed to be more rigid than the amide-I. So the quantum
fluctuation is dominated by the amide-I to enable us to use the
Gaussian approximation. Making use of the Gaussian approximation,
only the classical path of $Q$ contributes to the interaction term
\cite{levit}. Therefore the partition function becomes,
\be
\z= \z_x \z_Q \; ,
\label{par1}
\ee
where
\bea
\z_x = \int \id x \, \ex^{-\s_x/\hbar} \; ,\\
\z_Q= \int \id Q \, \ex^{-\s_Q/\hbar} \; ,
\label{eq:par2}
\eea
and the actions are,
\bea
\s_{x}&=& \int_0^\tau \d t \left( \frac{1}{2}m \dot{x}^2
+\frac{1}{2}\overline{k}
x^2 - \frac{1}{2}\delta x^4
+ \tilde{\chi} x \overline{Q} \right) \; , \\
\s_{Q}&=& \int_0^\tau \d t \left( \frac{1}{2}\overline{M} \dot{Q}^2
+\frac{1}{2}\tilde{\kappa} Q^2
- \frac{1}{4}\lambda Q^4 \right) \; .
\label{eq:par3}
\eea
\subsection{Partition function for the amide-site}
Further, one should solve the partition function of the amide-site ($\z_Q$). We
use Gaussian approximation to solve the partition
function. Under this approximation the general path can be
expressed in the usual way as $Q = \overline{Q} + \breve{Q}$, where
$\overline{Q}$ is the classical path and $\breve{Q}$ is the quantum
path \cite{feynmann,levit}. Expanding the action in Taylor series
$\s_Q$ becomes \cite{levit},
\be
\s_Q = \s_Q^\mathrm{cl} +\frac{1}{1!}\delta(\s_Q)^\mathrm{cl}
+\frac{1}{2!}\delta^2(\s_Q)^\mathrm{cl} + \cdots \; .
\label{eq:sQ1}
\ee
Since the classical path satisfies the variational principles, $\delta\s_Q = 0$,
and taking only the second order,
\be
\z_Q = \int \id Q \, \exp \left[ \frac{1}{\hbar} \left( \s_Q^\mathrm{cl} +
\frac{1}{2!} \delta^2 (\s_Q^\mathrm{cl}) \right) \right] \; .
\label{eq:sQ2}
\ee
This yields,
\bea
\z_Q & = & \ex^{-\s_Q^\mathrm{cl}/\hbar} \int \id \breve{Q} \, \exp \left\{
-\frac{1}{\hbar}\int \d t\frac{1}{2!} \left[ \left( \frac{\pd^2
\l^\mathrm{cl}_{\overline{Q}}}{\pd \overline{Q}^2} \right) \breve{Q}^2
\right. \right. \nonumber \\
&& \left.\left.
+ 2 \left( \frac{\pd^2
\l^\mathrm{cl}_{\overline{Q}}}{\pd\overline{Q}\dot{\overline{Q}}} \right)
\breve{Q} \dot{\breve{Q}}
+ \left( \frac{\pd^2 \l^\mathrm{cl}_{\overline{Q}}}{\pd
\dot{\overline{Q}}^2} \right) \dot{\breve{Q}}^2
\right] \right\} \; ,
\label{eq:sQ3}
\eea
where,
\be
\l^\mathrm{cl}_{\overline{Q}} = \frac{1}{2} \overline{M} \dot{\overline{Q}}^2
+ \frac{1}{2}
\tilde{\kappa} \overline{Q}^2
- \frac{1}{4} \lambda \overline{Q}^4 \; .\
\label{eq:sQ4}
\ee
Calculating the second variation, and substituting the result into $\z_Q$ in Eq.
(\ref{eq:sQ3}) one has,
\bea
\z_Q & = & \ex^{-\s_Q^\mathrm{cl}/\hbar} \, \z_{Q_0} \nonumber \\
& = & \exp \left[ -\frac{1}{\hbar} \int_{0}^\tau \d t
\left( \frac{1}{2} \overline{M} \dot{\overline{Q}}^2 +
\frac{1}{2}\tilde{\kappa}
\overline{Q}^2 -
\frac{1}{4}\lambda \overline{Q}^4 \right)
\right] \nonumber\\
&&
\times \int \id \breve{Q} \, \exp \left[
-\frac{1}{\hbar} \int_{0}^\tau \d t
\right. \nonumber \\
&& \left.
\times \left( \frac{1}{2} \overline{M} \dot{\breve{Q}}^2
+ \frac{1}{2}\tilde{\kappa} \breve{Q}^2 - \frac{3}{2}\lambda\overline{Q}^2
\breve{Q}^2 \right)
\right] \; .
\label{eq:sQ5}
\eea
The equation of motion for $\overline{Q}$ in Euclidean coordinate is given by,
\be
\frac{d ^2\overline{Q}}{d t^2} - \overline{\Omega}^2\overline{Q} +
\frac{\lambda}{\overline{M}} \overline{Q}^3 = 0 \; ,
\label{eq:eomQ}
\ee
where $\overline{\Omega}^2 = {\tilde{\kappa}}/\overline{M}$. In this paper
the solution is taken to have the form of,
\be
\overline{Q}= \overline{Q}_0 \, \mathrm{sech}(\overline{\Omega}t)\; ,
\label{eq:solutionsQ}
\ee
which leads to $\overline{Q}_0= \overline{\Omega}\sqrt{{(2 M)}/\lambda}$. This
choice is motivated by the fact that the Davydov model is the self-trapping of
its energy, \ie it should be localized. Substituting this
solution into classical action and using the identity $1 - \mathrm{sech}^2 x =
\tanh^2 x$, one obtains,
\be
\s_Q^\mathrm{cl} = A_1 \tanh \left(\overline{\Omega} \hbar \beta \right)
+ A_2 \tanh^3 \left( \overline{\Omega}\hbar\beta \right) \; ,
\label{eq:sQklasik}
\ee where, \bea
A_1 & = &
\frac{1}{2}\overline{M}\overline{\Omega}\overline{Q}_0^2-\frac{1}{4}\frac{
\lambda}{\overline{\Omega}
}\overline{Q}_0^4 \; , \\
A_2 & = &
\frac{1}{12}\frac{\lambda}{\overline{\Omega}}\overline{Q}_0^4-\frac{1}{2}
\overline{M}\overline{
\Omega}\overline{Q}_0^2 \; .
\label{eq:coefsQklasik}
\eea
Now the problem is turned into solving the prefactor in the path integral,
\bea
\z_{Q_0} & = & \int \id \breve{Q} \, \exp \left[ -\frac{1}{\hbar}
\int_{0}^\tau \d t
\left( \frac{1}{2} \overline{M} \dot{\breve{Q}}^2
\right.\right. \nonumber \\
&& \left.\left.
+ \frac{1}{2}\kappa \breve{Q}^2 - \frac{3}{2}\lambda\overline{Q}^2
\breve{Q}^2 \right) \right] \; .
\label{eq:prefactorsQ}
\eea
Using semi-classical approximation it can be rewritten as \cite{ranfagui},
\be
\mathcal{Z}_{Q_0}= \frac{1}{\sqrt{2\pi \hbar}} \left. \left(\frac{\delta^2
\s_{\breve{Q}_0}}{\delta \breve{Q}^2} \right)^{1/2} \right|_{\overline{Q}}
\label{eq:prefactorsQ2} \; ,
\ee
and its second order variation is given by,
\bea
\left. \frac{\delta^2 \s_{\breve{Q}_0}}{\delta \breve{Q}^2}
\right|_{\overline{Q}} & = &
\frac{\tau \sqrt{\overline{M}\tilde{\kappa} }}{2\pi \hbar} \int^\tau_0 \d t
\left( \frac{\pd^2 \l^\mathrm{cl}_{\overline{Q}}}{\pd
\dot{\overline{Q}}^2}\ddot{\overline{Q}}^2
\right. \nonumber \\
&& \left.
+ 2\frac{\pd^2 \l^\mathrm{cl}_{\overline{Q}}}{\pd \dot{\overline{Q}}
\pd \overline{Q}}\overline{Q}\dot{\overline{Q}} + \frac{\pd^2
\l^\mathrm{cl}_{\overline{Q}}}{\pd \overline{Q}^2}\dot{\overline{Q}}^2 \right)
\label{eq:prefactorsQ3} \; .
\eea
The result is,
\be
\left. \frac{\delta^2 \s_{\breve{Q}_0}}{\delta \breve{Q}^2}
\right|_{\overline{Q}}
=
\frac{\tau \sqrt{\overline{M}\tilde{\kappa} }}{2\pi \hbar} \int^\tau_0 \d t
\left[ \overline{M}\ddot{\overline{Q}}^2
+ \left( - \tilde{\kappa} + 3 \lambda \overline{Q}^2 \right)
\dot{\overline{Q}}^2
\right]
\label{eq:prefactors4} \; ,
\ee
and by substituting $\overline{Q}$ in Eq. (\ref{eq:solutionsQ}),
\bea
\left. \frac{\delta^2 \s_{\breve{Q}_0}}{\delta \breve{Q}^2}
\right|_{\overline{Q}}
& = &
\frac{\beta \sqrt{\overline{M}\tilde{\kappa} }}{2\pi}
\left[ \Lambda_1 \, \mathrm{sech}^4(\overline{\Omega}\hbar \beta
)\tanh(\overline{\Omega}\hbar \beta)
\right. \nonumber \\
&& \left.
\times \left( \cosh(2\overline{\Omega}\hbar \beta) - 3 \right)
+ \Lambda_2 \tanh^3(\overline{\Omega}\hbar \beta )
\right. \nonumber\\
&& \left.
+ \Lambda_3 \mathrm{sech}^2 (\overline{\Omega} \hbar\beta) \tanh^3
(\overline{\Omega}\hbar \beta )
\right. \nonumber \\
&& \left.
\times (4 + \cosh(2 \overline{\Omega}\hbar \beta ) \right] \; ,
\label{eq:prefactorsQ6}
\eea
where,
\bea
\Lambda_1 & = & \frac{1}{2}\overline{M}\overline{Q}_0^2 \; , \\
\Lambda_2 & = &
-\frac{1}{3}\tilde{\kappa}\overline{\Omega}\overline{Q}_0^2\; , \\
\Lambda_3 & = & \frac{1}{5}\lambda \overline{\Omega}\overline{Q}_0^4\; .
\label{eq:coefprefactorsQ}
\eea
Finally, the complete partition function for the amide-site is
obtained, \bea
\z_Q & = &
\frac{1}{2\pi}\sqrt{\frac{\beta}{\hbar}}(\overline{M}\tilde{\kappa})^{1/4}
\left\{ \Lambda_1 \, \mathrm{sech}^4(\overline{\Omega}\hbar \beta
)\tanh(\overline{\Omega}\hbar \beta)
\right. \nonumber \\
&& \left.
\times (\cosh(2\overline{\Omega}\hbar \beta)-3)
- \Lambda_2 \tanh^3(\overline{\Omega}\hbar \beta )\right. \nonumber \\
&& \left.
+ \Lambda_3 \, \mathrm{sech}^2(\overline{\Omega}
\hbar\beta)\tanh^3(\overline{\Omega}\hbar \beta )
(4+\cosh(2\overline{\Omega}\hbar \beta))\right\}^{1/2}
\nonumber \\
&&
\times \exp \left[- \frac{1}{\hbar} (A_1 \tanh(\overline{\Omega} \hbar \beta)
+ A_2 \tanh^3(\overline{\Omega}\hbar\beta)) \right] \; .
\label{eq:partisiQ}
\eea
\subsection{Partition function for the amide-I}
For the action of amide-I, $\s_x$, the partition function is,
\bea
\z_x & = & \int \id x \, \exp \left[ -\frac{1}{\hbar} \int \d t
\left(\frac{1}{2}m \dot{x}^2 +\frac{1}{2}\overline{k} x^2
+ \tilde{\chi} x \overline{Q}
\right.\right. \nonumber \\
&& \left. \left.
- \frac{1}{2}\delta x^4 \right) \right] \; .
\label{eq:sX}
\eea
Dividing $x$ into classical path $\overline{x}$ and quantum path $\breve{x}$,
\ie
$x=\overline{x}+\breve{x}$, and again using the Gaussian approximation, the
classical
path is,
\be
\s_x^\mathrm{cl}= \int \d t \left( \frac{1}{2}m \dot{\overline{x}}^2
+\frac{1}{2}\overline{k} \overline{x}^2
- \frac{1}{2}\delta \overline{x}^4 + \tilde{\chi}
\overline{x}\overline{Q} \right) \; .
\label{eq:sXklasik1}
\ee
This can be solved by determining the classical path which is the solution of
following equation,
\be
m\ddot{\overline{x}} - \overline{k} \overline{x} + 2\delta \overline{x}^3 =
\tilde{\chi} \overline{Q}_0 \, \mathrm{sech} (\overline{\Omega} t ) \; .
\label{eq:eomsXklasik}
\ee
However this is hard to be solved analytically. Instead one
can use the perturbation method, \ie calculating the solutions
order by order $\overline{x} = \overline{x}^0 + \varepsilon \overline{x}^1 +
\cdots$. Note that the inhomogeneous term is assumed being
generated from the leading order. Substituting this expansion into
Eq. (\ref{eq:eomsXklasik}) up to the leading orders one has,
\be
m\ddot{\overline{x}}^0 - \overline{k} \overline{x}^0 + 2\delta
\overline{x}^{03} = 0 \; ,
\label{eq:eomsXklasik0}
\ee
for the lowest order and,
\be
m\ddot{\overline{x}}^1 - \overline{k} \overline{x}^1 + 6\delta
\overline{x}^{02}\overline{x}^{1} =
\tilde{\chi} \overline{Q}_0 \, \mathrm{sech}( \overline{\Omega} t ) \; ,\
\label{eq:eomsXklasik02}
\ee
for the first order. Similar to the solution in the amide-site
coordinate, since $\delta < \overline{k}$, the solution for the zeroth
order is,
\be
\overline{x}^0= \overline{X}_{0} \, \mathrm{sech}(\overline{\omega} t) \; ,
\label{eq:solusisXklasik0}
\ee
where $\overline{\omega}=\overline{k}/m$,
$\overline{X}_{0}=\overline{\omega}\sqrt{m/{\delta}}$ after transforming $t
\rightarrow it$. In this solution $\delta$ must not be zero. For
the first order, the solution is $x^1=x^1_h+x^1_p$ and $x^1_h$
satisfies,
\be
\ddot{\overline{x}}_h^1 - \overline{\omega}^2 \overline{x}_h^1 +
6\frac{\delta}{m}
\overline{X}_{0}^2 \, \mathrm{sech}^2(\overline{\omega}t)\overline{x}_h^{1} = 0
\; .
\label{eq:eomsXklasik1h}
\ee Performing a transformation $\tau = \tanh(\overline{\omega} t)$ one
gets the associated Legendre equation,
\be
\frac{\d}{\d\tau} \left[ (1-\tau^2) \frac{\d\overline{x}_h^1}{\d\tau} \right]
+ \left[ l ( l + 1 ) + \frac{n^2}{1 - \tau^2} \right] \overline{x}_h^1 = 0 \; ,
\label{eq:solhomox}
\ee where $n^2=\overline{\omega}$ and $l(l+1)= 6
{\delta}/{(m\overline{\omega})} \overline{X}_{0}^2$. The solution is $x_h^1
= x_{h_0}^1P_l^n(\tanh(\overline{\omega} t))$ where $P_l^n(\tau)$ is
the associated Legendre function. Particularly its solution
satisfies,
\be
m\ddot{\overline{x}}_p^1 - m \overline{\omega}^2 \overline{x}_p^1
+ 6\delta \overline{X}_{0}^2 sech^2(\overline{\omega} t)\overline{x}_p^{1} =
\tilde{\chi}
\overline{Q}_0 \, \mathrm{sech}( \overline{\Omega} t ) \; .
\label{eq:eomsxklasik1p}
\ee
The solution of this equation can be written using Green function
$G(\tau,\tau')$ as follow,
\be
\overline{x}_p^{1} = \frac{\tilde{\chi} \overline{Q}_0}{m} \int \d\tau' \,
G(\tau,\tau')
\mathrm{sech}
\left[ \frac{\overline{\Omega} }{\overline{\omega}} \tanh^{-1}(\tau') \right]
\; ,\
\label{eq:solkhususx}
\ee
and the Green function is governed by,
\bea
&& \frac{\d}{\d\tau} \left[(1-\tau^2) \frac{\d G(\tau,\tau')}{\d\tau} \right]
+ \left[ l (l + 1) - \frac{m^2}{1 - \tau^2} \right] G(\tau,\tau')
\nonumber \\
&& = -\delta(\tau-\tau') \; .
\label{eq:green}
\eea
The Green function is given by \cite{filho}, $G(\tau,\tau')=(-1)^n P_l^n(\tau_<)
Q_l^n(\tau_>)$
with $Q_l^n(\tau)$ is the associated Legendre functions of the second kind. Its
complete solution is,
\bea
\overline{x} & = & \overline{X}_{0} \, \mathrm{sech}(\overline{\omega} t) +
\varepsilon
\left[ x_{h_0}^1P_l^n(\tanh(\overline{\omega} t))
\right. \nonumber \\
& & \left.
+ \frac{\tilde{\chi} \overline{Q}_0}{m} \int \d\tau' \, G(\tau,\tau')
\mathrm{sech}
\left( \frac{\overline{\Omega} }{\overline{\omega}} \tanh^{-1}(\tau') \right)
\right]
\; .\
\label{eq:completesol}
\eea
Substituting this result into Eq. (\ref{eq:sXklasik1}) one obtains the classical
action.
On the other hand, the classical action up to the first order is given by,
\bea
\s_x^\mathrm{cl} & = & \Delta_1 \tanh(\overline{\omega} \hbar \beta) +
\Delta_2
\tanh^3(\overline{\omega} \hbar \beta)
\nonumber\\
&& + \Delta_3 \int_0^{\beta \hbar} \d t \, \mathrm{sech}(\overline{\omega}
t) \,
\mathrm{sech}(\overline{\Omega}t)\\
&& + \varepsilon (F_1 + F_2 + F_3 + F_4 +F_5 + F_6 + F_7 + F_8)
\nonumber \; ,\
\label{eq:sXklasik2}
\eea
where,
\bea
\Delta_1 & = & \frac{1}{2}m \overline{\omega} \overline{X}_{0}^2
+\frac{\delta}{3\overline{\omega}}\overline{X}_{0}^4 \; ,\\
\Delta_2 & = & -\frac{m}{2} \overline{\omega} \overline{X}_{0}^2 +
\frac{\delta\overline{X}_{0}^4}{6 \overline{\omega}} \; ,\\
\Delta_3 & = & -\frac{1}{2}\tilde{\chi} \overline{X}_{0}\overline{Q}_{0} \; ,
\label{eq:koefsXklasik2nya}
\eea
and $F_i$'s are given in App. \ref{app:1}.
Performing the same procedure as done in the previous subsection, one should
consider the prefactor of $\mathcal{Z}_x$,
that is,
\be
\mathcal{Z}_{x^0}= \frac{1}{\sqrt{2\pi \hbar}} \left. \left( \frac{\delta^2
\s_{\breve{x}^0}}{\delta \breve{x}^2} \right)^{1/2} \right|_{\overline{x}}
\label{eq:prefactorsX1} \; ,\
\ee
where the second order variation is given by,
\bea
\left.\frac{\delta^2 \s_{\breve{x}^0}}{\delta \breve{x}^2}
\right|_{\overline{x}} &
= &
\frac{\beta \sqrt{m\overline{\omega}^2}}{2\pi} \int^\tau_0 \d t
\left( \frac{\pd^2 \l^\mathrm{cl}_{\overline{x}}}{\pd
\dot{\overline{x}}^2}\ddot{\overline{x}}^2
\right. \nonumber \\
&&
\left.
+ 2\frac{\pd^2 \l^\mathrm{cl}_{\overline{x}}}{\pd \dot{\overline{x}}\pd
\overline{x}}\overline{x}\dot{\overline{x}}
+ \frac{\pd^2 \l^\mathrm{cl}_{\overline{x}}}{\pd
\overline{x}^2}\dot{\overline{x}}^2 \right)
\label{eq:prefactorsX2} \; ,
\eea
and,
\be
\l^\mathrm{cl}_{\overline{x}} = \frac{1}{2}m \dot{\overline{x}}^2
- \frac{1}{2}m \overline{\omega}^2 \overline{x}^2+ \frac{1}{2}\delta
\overline{x}^4 + \tilde{\chi} \overline{x} \overline{Q} \; , \ee which yields,
\bea
\left. \frac{\delta^2 \s_{\breve{x}^0}}{\delta \breve{x}^2}
\right|_{\overline{x}}
& = &
\frac{\beta \sqrt{m\overline{\omega}^2}}{2\pi} \int^{\beta \hbar}_0 \d t \,
\nonumber \\
&& \times \left[ m\ddot{\overline{x}}^2
- (m\overline{\omega}^2 + 3 \delta \overline{x}^2)\dot{\overline{x}}^2
\right]
\label{eq:prefactorsX3} \; .\
\eea
Substituting Eq. (\ref{eq:completesol}) into Eq. (\ref{eq:prefactorsX3}) and
keeping only the first order,
\bea
&& \left.\frac{\delta^2 \s_{\breve{x}^0}}{\delta
\breve{x}^2}\right|_{\overline{x}} =
\frac{\beta \sqrt{m\overline{\omega}^2}}{2\pi} \times
\nonumber\\
&&
\left[ \Gamma_1 \, \mathrm{sech}^4(\overline{\omega}\hbar \beta)
\tanh(\overline{\omega}\hbar \beta)(\cosh(2\overline{\omega}\hbar \beta)-3)
\right.\nonumber\\
&& \left.
+ \Gamma_2 \tanh^3(\overline{\omega} \hbar \beta)
\right.\nonumber\\
&& \left.
+ \Gamma_3 (4+\cosh(2\overline{\omega} \hbar \beta))
\mathrm{sech}^2(\overline{\omega}
\hbar \beta) \tanh^3(\overline{\omega} \hbar \beta)
\right.\nonumber\\
&& \left.
+ \varepsilon (G_1 + G_2 + G_3 + G_4 + G_5 \right.\nonumber\\
&& \left.+ G_6 + G_7 + G_8 +G_9 + G_{10}) \right] \; ,\
\label{eq:prefactorsX4} \; .\
\eea
where,
\bea
\Gamma_1 & = & -\frac{m}{2}\overline{\omega}^3\overline{X}_0^2 \; , \\
\Gamma_2 & = & \frac{m \overline{\omega}^3 \overline{X}_0^2}{3} \; , \\
\Gamma_3 & = & \frac{\delta \overline{\omega} \overline{X}_0^4}{5} \; ,
\label{eq:koefprefactorX4}
\eea
and $G_i$'s are given in the App. \ref{app:2}..
Hence, the partition function $\z_x$ is obtained,
\bea
&& \z_x = \left( \frac{\beta \sqrt{m\overline{\omega}^2}}{4\pi^2 \hbar}
\right)^{1/2} \times
\nonumber\\
&& \left[ -\Gamma_1 \, \mathrm{sech}^4(\overline{\omega}\hbar \beta)
\tanh(\overline{\omega}\hbar \beta)(\cosh(2\overline{\omega}\hbar \beta)-3)
\right. \nonumber\\
&& \left. + \Gamma_2 \tanh^3(\overline{\omega} \hbar \beta)
\right. \nonumber\\
&& \left. + \Gamma_3 (4+\cosh(2\overline{\omega} \hbar \beta)) \,
\mathrm{sech}^2(\overline{\omega} \hbar \beta)\tanh^3(\overline{\omega} \hbar
\beta)
\right. \nonumber\\
&& \left. + \varepsilon (G_1 + G_2 + G_3 + G_4 + G_5
\right. \nonumber\\
&& \left. + G_6 + G_7 + G_8 +G_9 + G_{10}) \right]^{1/2}
\nonumber\\
&&
\times \exp \left\{ - \frac{1}{\hbar} \left[
\Delta_1 \tanh(\overline{\omega} \hbar \beta)
+ \Delta_2 \tanh^3(\overline{\omega} \hbar \beta)
\right.\right. \nonumber \\
&& \left.\left.
+ \Delta_3 \int_0^{\beta \hbar} \d t \, \mathrm{sech}(\overline{\omega} t)
\,
\mathrm{sech}(\overline{\Omega}t)
\right. \right. \nonumber\\
&& \left.\left. + \varepsilon (F_1 + F_2 + F_3 + F_4 +F_5 + F_6 + F_7 + F_8)
\right] \right\} \; .\
\label{eq:partisiX}
\eea
Finally, the complete partition function density of Davydov-Scott
monomer in thermal bath, $\z = \z_x \, \z_Q$, is obtained from
Eqs. (\ref{eq:partisiQ}) and (\ref{eq:partisiX}).
\section{Thermodynamical properties of Davydov-Scott monomer}
\label{sec:tpdc}
\begin{figure}[b]
\includegraphics[width=0.45\textwidth]{specific1.eps}
\caption{The temperature dependence of normalized specific heat for various
values of $\gamma$ with $\lambda = 1$.}
\label{fig:heat}
\end{figure}
From physical point of view, the Davydov-Scott monomer is a harmonic oscillator
coupled to a quantum excitation. Using Euler-Lagrange equation, one can derive
the appropriate equation of motion (EOM) from the action in Eq.
(\ref{eq:inter1}). Although solving of the EOM is very interesting and
attractive, but it has a little physical significant due to unobservable
individual molecular motion of the Davydov-Scott monomer. Therefore in this
paper let us consider thermodynamic observable as specific heat \cite{feynmann},
\be
C = k_B \beta^2 \frac{\pd^2 \ln (\z)}{\pd \beta^2}\; .\
\label{eq:thermo}
\ee
Some previous works considering similar model as the Hamiltonian in
Eq. ({\ref{eq:mono1}}) to study the Davydov-Scott monomer in
thermal bath, for example the semiclassical approach in \cite{cuevas}. It has
been argued that using Lindblad formulation
the semiclassical limit is a good approximation to the corresponding full
quantum treatment at biological temperatures in the highly underdamped and
harmonic limits. In the semiclassical approximation, the coupling between
Davydov-Scott monomer with thermal bath is described by Langevin equation of the
amide-site
displacement characterized by $\gamma$ and $\Omega$ \cite{cruzeiro2}. If the
stochastic force represents the thermal bath is zero, the equation is reduced
into the damped harmonic oscillator. There are three regions regarding the
values of $\gamma$ and $\Omega$, that is $\gamma < 2\Omega$
for the underdamped condition, $\gamma = 2 \Omega$ for the critical damped
condition and $\gamma > 2 \Omega$ for the overdamped condition. In the Lindblad
formulation the damping coefficient $\gamma$ represents the relaxation time due
to interaction with the environment. The higher values of $\gamma$ corresponds
to the shorter relaxation time. The previous work by Cuevas \etal has
established that the semiclassical approximation is equivalent to the full
quantum approach (for biological temperature) as long as $\gamma \ll 2 \Omega$.
Otherwise, the oscilation frequency of the observable would be different
\cite{cuevas}.
In this paper, the analysis is done for the above three criterions. The values
of parameters used throughout numerical calculation are $M = 6.3 \times
10^{-26}$ kg, $m = 7.3 \times 10^{-26}$ kg,
$\kappa = 10$ Nm$^{-1}$ and $\omega = 1660$ cm$^{-1}$ \cite{cuevas,sinkala},
while $\lambda = 9$. The behavior of specific heat in term of temperature is
shown in Fig. \ref{fig:heat} for various damping parameter $\gamma$, Figs.
\ref{fig:heat1}$-$\ref{fig:heat3} for various strength of anharmonic oscillation
in three damping conditions and Fig. \ref{fig:heat4} for various $\gamma$ in the
underdamped case. The results are similar with the previous ones obtained in the
calculation of a system with anharmonic amide-site \cite{zoli}.
\begin{figure}[t]
\includegraphics[width=0.45\textwidth]{specific2.eps}
\caption{The temperature dependence of normalized specific heat for various
values of $\lambda$ for the underdamped condition with $\gamma = 0.1 \Omega$.}
\label{fig:heat1}
\end{figure}
\begin{figure}[b]
\includegraphics[width=0.45\textwidth]{specific3.eps}
\caption{The temperature dependence of normalized specific heat for various
values of $\lambda$ for the critical damped condition with $\gamma = 2
\Omega$.}
\label{fig:heat2}
\end{figure}
In the present case the damping coefficients are represented by the coefficients
$\delta_1 \sim \delta_6$. In particular, $\delta_1$ appears in the kinetic
terms, and can then be interpreted as the 'effective mass' of amide-site
vibration. Further, $\delta_3$ appears in the harmonic potential as the
'effective elastic constant'. Hence it can be argued that the
environment effects to the amide-site vibration occur through the
kinetic term and the harmonic potential. The coefficient $\delta_4$ represents
the strength of the interaction between amide-I.and the system.
\begin{figure}[t]
\includegraphics[width=0.45\textwidth]{specific4.eps}
\caption{The temperature dependence of normalized specific heat for various
values of $\lambda$ for the overdamped condition with $\gamma = 2.5 \Omega$.}
\label{fig:heat3}
\end{figure}
\begin{figure}[b]
\includegraphics[width=0.45\textwidth]{specific5.eps}
\caption{The temperature dependence of normalized specific heat for the
underdamped condition with various values of $\gamma < 2 \Omega$ and $\lambda =
1$.}
\label{fig:heat4}
\end{figure}
From the figures, the specific heat asymptotically approaches to zero at low
temperature and to infinity at high temperature. Actually $\phi^4$ potential
oscillator interacting with electron and also the anharmonic oscillator give
similar profiles \cite{schwarcz}. Large environment effect causes the
Davydov-Scott monomer to increase the energy and then the temperature to achieve
the equilibrium state. Correspondingly the vibration frequency is also affected
by the damping coefficients since $\omega=\sqrt{\tilde{\kappa}/\overline{M}} =
\sqrt{{(\kappa + i \delta_3)}/{(M + i \delta_2)}}$.
These results indicate that the interaction between Davydov-Scott monomer and
thermal bath depend on the strength of the coupling of system and environment.
Recent study of the open quantum system also shows that the canonical
equilibrium state of an open quantum system depends explicitly on the
system-bath coupling strength \cite{ingold,campisi}.
In particular, from Fig. \ref{fig:heat1} the anomaly of
specific heat that becomes negative for certain parameter sets at high
temperature region is observed. The same phenomena have been pointed out by
Ingold \etal \cite{ingold,ingold2}. In the current case, the anomaly especially
appears for the underdamped condition as shown in Figs. \ref{fig:heat1} and
\ref{fig:heat4}. It is also found that the negative specific heat is restored at
large $\lambda$ and $\gamma$, \ie for large anharmonic oscillation and
intensity of thermal bath.
It should also be remarked that one cannot take $\lambda=0$ (no oscilation
effect) since all results are obtained from Eq. (\ref{eq:solutionsQ}) which is
a special case with $\lambda \ne 0$ condition.
\section{Summary}
\label{sec:su}
The interaction of Davydov-Scott monomer with thermal bath is investigated
using the Lindblad formulation of master equation. In contrast with
previous work by Cuevas \etal \cite{cuevas}, the anharmonic oscillation term of
amide-site is taken into account. Adopting similar Lindblad operators used in
\cite{cuevas}, the master equation of the system is obtained. Instead of solving
the equation of motion, the thermodynamic partition function and in particular
specific heat are calculated using the path integral methods.
It is shown that the coupling with the environment contributes to the kinetic
term, the harmonic potential of amide-site vibration and the anharmonic term of
amide-I.
The anomaly of specific heat that becomes negative for certain parameter sets at
high temperature region is observed as pointed out by Ingold \etal in the case
of pure open quantum systems \cite{ingold,ingold2}. However, it is found that
the negative specific heat is restored for large anharmonic oscillation effect.
In contrast to these results, Ingold \etal have found that the anomaly occurs at
low temperature region. This discrepancy can be explained as the consequences of
different approaches adopted to model the interaction between the system and the
thermal bath. Ingold \etal has represented the interaction in a set of harmonic
oscillators which becomes the coupled harmonic oscillator at classical
approximation. In contrary, in the present approach the thermal bath
is represented in a set of Lindblad operators which becomes the underdamped
harmonic oscillator at classical approximation.
From the figures, it can in general be concluded that the anharmonic oscillation
contributes constructively to the specific heat.
\begin{acknowledgments}
AS thanks the Group for Theoretical and Computational Physics
LIPI for warm hospitality during the work. This work is funded by the Indonesia
Ministry of Research and Technology and the Riset Kompetitif LIPI in fiscal year
2010 under Contract no. 11.04/SK/KPPI/II/2010. FPZ is supported by Riset KK
2010 Institut Teknologi Bandung.
\end{acknowledgments}
|
\section{Introduction}
Flux pinning is one of the most important factors that governs the critical current density, $J_c$, in a type-II superconductor. The critical current density is determined by the balance of two opposing forces acting on the magnetic flux lines: the pinning force due to the spatial variation of the condensation energy and the Lorentz force exerted by the transport current. Energy is dissipated whenever flux lines move. Traditionally, one distinguishes two regimes of dissipation: flux creep when the pinning force dominates \cite{ander} and flux flow when the Lorentz force dominates \cite{ybkim,tinkham}. A complete knowledge on the dynamics of the magnetic flux lines is, therefore, required in order to understand whether a system is potentially attractive candidate for technological applications. In cuprate superconductors, due to high anisotropy, short coherence length and high operating temperature, the pinning energy of the vortices is weak, leading to strong fluctuation and motion of the vortices. Because of the weak pinning, the energy dissipation caused by the thermally activated flux motion is present even at a finite current density, which limits the maximum value of $J_c$, and hence a variety of applications of the cuprates.
In this perspective, the pnictide superconductors are important for studying the vortex dynamics. The phenomenally high upper critical field $H_{c2}$ \cite{bhoi1,jaron,lee} and lower anisotropy observed in the pnictide superconductors indicate encouraging potential applications \cite{oka}. Among different families of pnictide superconductors, the growth of $A$Fe$_2$As$_2$ (122 system, where $A$ = Ba, Ca, Sr) single crystals is relatively easier \cite{milton,saha,prozo}. Eventually, the study of vortex dynamics in these systems has been extensively carried out \cite{prozo,kim,yang1,yama1,shen,kope}. The vortex dynamics in 122 single crystals has been analyzed using weak collective-pinning$-$collective flux-creep model. In the (Ba,K)Fe$_2$As$_2$ single crystal, a vortex-glass to vortex-liquid transition has been observed \cite{kim}. One of the prominent features in pnictide superconductors is the observation of a fishtail or second peak effect in the $M(H)$ curve \cite{prozo,kim,yang1,yama1,shen,kope}. Analysis of the magnetic relaxation data in Ba(Fe$_{0.93}$Co$_{0.07}$)$_2$As$_2$ single crystal shows that the field at which the fishtail magnetization is maximum, a crossover in the vortex dynamics from the collective to the plastic creep occurs \cite{prozo}. However, due to the scarcity of $Re$FeAsO$_{1-x}$F$_x$ (1111 system, where $Re$ = La, Ce, Pr, Nd, Sm, Gd, etc.) single crystals, vortex dynamics has been investigated mainly in polycrystalline samples. The study of vortex dynamics in polycrystalline SmFeAsO$_{0.9}$F$_{0.1}$ \cite{yang} and NdFeAsO$_{0.9}$F$_{0.1}$ \cite{prozo1} samples suggests a collective vortex pinning and creep mechanism in this system. In the present work, we have studied magnetization $M(T,H)$ in order to evaluate the critical state parameters related to the vortex dynamics and to construct the mixed-state phase diagram of the PrFeAsO$_{0.60}$F$_{0.12}$ superconductor.
\section{Sample characterization}
\begin{figure}[h]
\includegraphics[width=0.5\textwidth]{Fig1.pdf}\\
\caption{(a) Field-cooled and zero-field cooled dc susceptibilities at $H$ = 10 Oe and (b) real part of the ac susceptibility at $H$= 3 Oe and 100 Hz of the PrFeAsO$_{0.60}$F$_{0.12}$ sample. The slight paramagnetic susceptibility at the normal state has been subtracted from the above curves. (c) Low magnification and (d) high magnification scanning electron microscope images of the PrFeAsO$_{0.60}$F$_{0.12}$ sample. The high magnification image shows platelike crystallites of different sizes.}\label{Fig.1}
\end{figure}
We have synthesized oxygen deficient and fluorine-doped sample of nominal composition PrFeAsO$_{0.60}$F$_{0.12}$ by standard solid state reaction method. The details of the sample preparation have been discussed in our earlier reports \cite{bhoi}. The phase purity of the sample was determined by powder x-ray diffraction method with Cu K$\alpha$ radiation and details of the structural analysis are reported in Ref.\cite{bhoi}. For further characterization, we have measured the low-field magnetization of the sample using a SQUID magnetometer. The temperature dependence of zero-field-cooled (ZFC) and field-cooled (FC) dc magnetic susceptibilities at $H$ = 10 Oe are shown in figure 1(a). Compared to NdFeAsO$_{1-x}$F$_x$ \cite{prozo1,pissa} and other reported polycrystalline samples \cite{yama,yang,ren,yama2} the superconducting transition is rather sharp ($\Delta T_c \sim$ 4.5 K), indicative of better quality of the sample. Both the ZFC and FC susceptibilities start to deviate from the normal behavior below 48 K due to the appearance of a diamagnetic signal, which is close to the zero-resistance temperature $T_{c}^{0}$ \cite{bhoi}. At 4 K, the shielding and Meissner fractions are calculated to be 88\% and 35\% respectively, showing the bulk nature of superconductivity in the sample. We have also determined the shielding fraction of the sample $\sim$ 100\% at 4 K from the real part of the ac susceptibility curve [figure 1(b)]. In order to determine the average grain size in the sample, scanning electron microscope (SEM) (Quanta 200, FEG) images of the polished surface were obtained. SEM images of the polished surface of the PrFeAsO$_{0.60}$F$_{0.12}$ sample are shown in figure 1(c) and 1(d). The image in figure 1(c) reveals that the sample is porous and the conglomerate particle size varies between 20 and 50 $\mu$m. High magnification image of the particles shows plate-like crystallites with a size of 2-30 $\mu$m [figure 1(d)]. It may be mentioned that the average grain size in this sample is larger than that in LaFeAsO$_{0.89}$F$_{0.11}$, NdFeAsO$_{1-\delta}$ and SmFeAsO$_{0.85}$F$_{0.15}$ polycrystalline samples, where the grains are of the order of few hundreds of nanometer \cite{yama, moor,sena}. The energy dispersive x-ray (EDX) analysis was used to determine the chemical composition of the grains with different morphology. From the EDX spectra, we found that the oxygen content is about 0.75 and, the cations and fluorine contents are close to that of nominal composition. Examining the composition at several points on the surface of the sample, we did not find any local inhomogeneity. This result suggests that the grains are chemically homogeneous at least within the limit of SEM-EDX analysis. Hereafter, in the text we shall only mention the nominal composition of the sample. dc magnetization measurements at moderate and high fields were carried out using a Quantum Design superconducting quantum interference device with magnetic field up to 7 T. The magnetic hysteresis loops were measured using a 14 T vibrating sample magnetometer (Quantum Design). The magnetic hysteresis loops were measured at fixed temperatures with a constant field sweeping rate of 80 Oe s$^{-1}$.
\section{Results and Discussion}
\begin{figure}[t]
\includegraphics[width=0.5\textwidth]{Fig2.pdf}\\
\caption{color online) Temperature dependence of the zero-field-cooled and field-cooled magnetization curves of the PrFeAsO$_{0.60}$F$_{0.12}$ sample under various applied magnetic fields. Left panel: for $H \leq$ 1 T. Right panel: for $H \geq$ 2 T. Upward arrows indicate the shoulder-like feature of the magnetization curve. Downward arrows indicate the superconducting onset transition temperature.}\label{Fig.2}
\end{figure}
Figure 2 shows the temperature dependence of ZFC magnetization ($M_{zfc}$) and FC magnetization ($M_{fc}$) of the PrFeAsO$_{0.60}$F$_{0.12}$ sample under different applied magnetic fields. For $H$ = 0.05 T, the $M(T)$ curve exhibits a usual superconducting behavior with the diamagnetic onset temperature $\sim$ 48 K. However, with the increase of magnetic field the difference between the ZFC and FC magnetizations reduces, and the magnetization becomes positive for fields above $H \simeq 1$ T as the paramagnetic component of the magnetization arising from the Pr ion overcomes the diamagnetic signal of the superconducting electrons. In contrast to the low-field data [figure 1(b)], the $M_{zfc}(T)$ curve for applied fields $H$$\geq$0.05 T displays a shoulder-like feature (shown in the figure with the upward arrow mark) indicating the onset of intergrain superconductivity which is expected in a granular superconductor. The granular behavior is related to the weak links of the grain boundaries. Weak links impose lower global (intergrain) critical current density in comparison to the intragrain critical current density. With the increase of magnetic field, the shoulder shifts towards the lower temperature region. This behavior is related to the magnetic flux penetration inside the individual grains and is determined by the crossover from intragrain to intergrain superconductivity. Similar granular behavior has also been observed in LaFeAsO$_{1-x}$F$_x$ \cite{yama}, SmFeAsO$_{1-x}$F$_x$ and NdFeAsO$_{1-x}$F$_x$ oxypnictides \cite{sena,yama2}.
In order to check that the paramagnetic component of the magnetization comes actually from the Pr$^{3+}$ ion, we have analyzed the temperature dependence of dc magnetic susceptibility measured at 500 Oe field in the normal state as shown in figure 3. The susceptibility curve can be fitted with the Curie-Weiss law, $\chi(T) = \chi(0) + C/(T+\theta)$ in the temperature region 90 K to 300 K, where $\chi(0)$ is the temperature independent Pauli susceptibility, $C$ = N$\mu_{eff}^2/3k_B$, the Curie constant, N is the number of Pr ions in the compound and $\theta$ is the Curie-Weiss temperature. The calculated effective moment $\mu_{eff}$ of the Pr$^{3+}$ ion is equal to 3.8 $\mu_B$ which is comparable with the derived magnetic moments 3.75 $\mu_B$ \cite{rotundu} and 3.64 $\mu_B$ \cite{mcguire} for the PrFeAsO samples. The small discrepancy from the effective moment of the free Pr$^{3+}$ ion (3.58 $\mu_B$) could be attributed to the possible crystalline electric field environment of the Pr$^{3+}$ ion \cite{rotundu}. Thus, we can ascribe the origin of paramagnetic background in the $M(T)$ data to the magnetic moment of the Pr ion. That is, the paramagnetic contribution to the total magnetization is intrinsic, not impurity related. In NdFeAsO$_{0.94}$F$_{0.06}$ sample too, large paramagnetic contribution to the $M(T,H)$ data has been observed because of the large effective magnetic moment of the Nd$^{3+}$ ion (3.62 $\mu_B$) \cite{tarantini}. The difference in the $M(T,H)$ curve from those of the La- and Sm-based Fe-pnictides is due to the large magnetic moment of the free Pr and Nd ions compared to those of La (= 0 $\mu_B$) and Sm ($\sim$ 0.8 $\mu_B$) ions.
\begin{figure}[t]
\includegraphics[width=0.4\textwidth]{Fig3.pdf}\\
\caption{Temperature dependence of the dc susceptibility of PrFeAsO$_{0.60}$F$_{0.12}$ sample in the normal state measured at $H$= 500 Oe. The solid line is the result of the fit to Curie-Weiss law.}\label{Fig.3}
\end{figure}
Figure 4 presents the magnetization hysteresis $M(H)$ loops of the PrFeAsO$_{0.60}$F$_{0.12}$ sample at different temperatures in the field range -14 T $\leq$ H $\leq$ 14 T. One can clearly see that the superconducting hysteresis loop arises from the flux gradient produced by the pinning of flux lines. The observed $M(H)$ curve is the sum of a superconducting irreversible signal ($M_s$) and a paramagnetic component ($M_p$). In the presence of flux pinning, the sign of the superconducting magnetization depends on the direction of field sweep, thus resulting in a hysteresis loop between ascending and descending branches of the magnetization. The symmetric nature of the hysteresis curve suggests the bulk pinning of the vortices, not the surface barrier, dominates the magnetization of the sample. Even a 14 T magnetic field is not sufficient to achieve the closure of the width of the $M(H)$ loop at 25 K, indicating a large irreversible field of the superconductor. With increasing temperature, the irreversible diamagnetic loop shrinks and the paramagnetic component decreases significantly. Taking the mean value of the upper and lower hysteresis branches of $M(H)$ as the paramagnetic component $M_p$ of the total magnetization, we deduce the superconducting magnetization due to the critical state of pinned vortices as $\Delta M$ = $(M^{+}- M^{-})$/2, where $M^{+}(M^{-})$ is the branch of the magnetization for ${dH}/{dt}<0$ (${dH}/{dt}>0$).
\begin{figure}
\includegraphics[width=0.4\textwidth]{Fig4.pdf}\\
\caption{(color online) $M(H)$ loop of the PrFeAsO$_{0.60}$F$_{0.12}$ sample at 5, 7 and 10 K in the magnetic field range -14 T $\leq H \leq$ 14 T.Inset: $M(H)$ curve at 15, 20 and 25 K measured up to 14 T.}\label{Fig.4}
\end{figure}
In a polycrystalline sample, the gap $\Delta M$ in the magnetization loop can be split into an intergranular (global) part and an intragranular (local) part \cite{claim}, $\Delta M$ = $\Delta M_G$ + $\Delta M_L$. Here, $\Delta M_G$ is the magnetic moment originating from the intergranular current density $J_G$ which flows across the whole sample dimension and is related to the grain-boundary Josephson junction current. The second term presents the sum over all the magnetic moments of the grains and is determined by the intragranular current density $J_L$ circulating within the grains. The intragranular current density is determined by the pinning of flux lines in the individual grains. A bulk polycrystalline superconductor can be described as a percolative network of weak links. The critical current is determined by the interconnection of individual superconducting regions, interconnections that take the form of normal or insulating tunnel barriers at grain or subgrain boundaries. The low value of the resulting transport critical current density is rapidly suppressed by an increase of the magnetic field or of the temperature. In low magnetic field region, the gap $\Delta M$ in magnetization loop is mainly caused by the intergranular current, but in the high-field region $\Delta M$ results largely due to the intragranular current. This has been confirmed from the magnetization loop measurements on the bulk sample and on powder prepared from the same bulk sample in cuprates and pnictides \cite{muller,chen}. Chen \emph{et. al.} \cite{chen} measured the magnetization on the bulk and powdered SmFeAsO$_{0.80}$F$_{0.20}$ samples at different temperatures and magnetic fields. In the high-field region, the magnetization of the bulk and powdered samples does not show any significant difference indicating that the gap in the magnetization loops is mainly due to the intragrain supercurrent instead of the current across grain boundaries. But, at zero field ($H$ = 0), the gap in the bulk sample is significantly larger than that for the powder sample, reflecting the contribution from the grain boundary supercurrent in the bulk sample. As the field increases from $H$ = 0 to a critical value, $\Delta M$ for the powder becomes slightly larger than that for the bulk due to the introduction of extra defects into the grains of the powder sample which act as additional pinning centers. The difference in magnetization between the two samples reduces rapidly with further increase of magnetic field. This indicates that the evaluation of the intragrain current from the gap in the magnetization loop of the powder sample will be overestimated in the low-field region. Excluding the low-field region close to $H$=0, we can assume that the gap $\Delta M$ in our bulk sample appears mainly due to the intragrain current density.
Assuming that the current is flowing within the grains, we can evaluate the intragrain current density by using the Bean formula $J_L = \frac{30\Delta M}{\langle R\rangle}$, where $\langle R\rangle$ is the average grain size \cite{bean}. Figure 5 shows the magnetic field dependence of the critical current density over the temperature range 5-35 K for an estimated average grain size of 5 $\mu$m as revealed by the SEM images. The estimated intragrain $J_L$ is 5$\times 10^5$ A cm$^{-2}$ at 5 K and 5 T which is in agreement with the result obtained from the magneto-optical imaging method (3$\times 10^5$ A cm$^{-2}$) in PrFeAsO$_{1-y}$ single crystal \cite{beek}. The value of intragrain $J_L$ for the present sample (5$\times10^5$ A cm$^{-2}$ at $H$ = 5 T and 5 K) is higher than that reported for other polycrystalline 1111 samples \cite{gao,wang} but comparable with the in-plane current density of 122 single crystal ($2.6\times 10^5$ A cm$^{-2}$ at 5 K and $H$ = 0) \cite{prozo}. The reported intragrain $J_L$ at $H$ =0 is $6 \times10^4$ A/cm$^{2}$ for SmFeAsO$_{0.7}$F$_{0.3}$, $2\times10^5$ A/cm$^{2}$ for SmFeAsO$_{0.65}$F$_{0.35}$ \cite{gao} and $2\times10^5 - 10^6$ A/cm$^{2}$ for NdFeAsO$_{0.82}$F$_{0.18}$ \cite{wang} polycrystalline samples.
\begin{figure}
\includegraphics[width=0.4\textwidth]{Fig5.pdf}\\
\caption{(color online) (a) Field dependence of the intragrain critical current density of PrFeAsO$_{0.60}$F$_{0.12}$ sample over the temperature range 5-35 K. (b) The $J_c(H)$ curve is plotted in double logarithmic scale for some selected temperatures. The solid lines show the power law $H^{-5/8}$ plot to the curve. The characteristic field $H^*$, $H_{on}$ and $H_p$ are indicated by arrows.}\label{Fig.5}
\end{figure}
In order to understand the pinning mechanism and to evaluate the parameters related to vortex dynamics, we have investigated the field dependence of the critical current density at different temperatures. A clear nonmonotonous field dependence of the intragrain critical current density is observed as shown in the figure 5(b) on a double logarithmic plot. [Figure 5(b) shows the field dependence of critical current density on a double logarithmic plot.] Up to a field $H^*$, the critical current density is almost independent of $H$ and then decreases rapidly with a power-law $H^{-5/8}$. In the intermediate field the critical current density saturates to a constant value $J^{SV}_c$. With further increase of the magnetic field above $H_{on}$, $J_c$ starts to increase and reaches the maximum value at $H_p$. In the plateau region, the single vortex pinning dominates. In the region between $H^*$ and $H_{on}$, the rapid decrease of critical current is due to the intervortex interaction. It has been claimed in Ref \cite{salem} that a crossover from single to collective pinning occurs above $H_{on}$. The constant value, $J^{SV}_c$, of the critical current density in the intermediate field region has been attributed to weak collective pinning of vortex lines by the small scale fluctuations of the local dopant atom density or oxygen vacancy \cite{beek}. The pinning mechanism is identified as being due to mean free path variations in the vortex core, where the dopant atoms acts as the effective quasiparticle scatterers. In this region, the critical current $J^{SV}_c$ is given as \cite{beek,beek1}
\begin{equation}
J^{SV}_c \simeq j_0 \left[\frac{0.1n_dD_v^4}{\varepsilon_\lambda\xi_{ab}}\left(\frac{\xi_0}{\xi_{ab}}\right)^2\right]^{2/3},
\end{equation}
where $j_0$ = 4$\epsilon_0$/$\sqrt3\Phi_0\xi_{ab}$ is the depairing current density, $n_d$ is the defect density, $D_v$ is the effective ion radius, $\varepsilon_{\lambda}$ = $\lambda_{ab}/\lambda_c$ is the low field anisotropy ratio of penetration depths, $\xi_{ab}$ is the inplane coherence length, $\epsilon_0$ = $\Phi_0^2$/$4\pi\mu_0\lambda_{ab}^2$ is the typical vortex energy scale, $\Phi_0$ is the flux quantum and $\xi_0 \simeq 1.35\xi(0)$ as the temperature independent Bardeen-Cooper-Schrieffer coherence length. The low temperature value of the $J^{SV}_c$ = 5$\times 10^5$ Acm$^{-2}$ can be reproduced if we assume $D_v \sim 1.35 \AA$ as the average ionic radius for oxygen and fluorine ion, with defect density $n_d\sim$ 3.5$\times 10^{27}$ m$^{-3}$. This value of $n_d$ corresponds to the 0.25 times oxygen vacancies in our sample. As our sample is simultaneously oxygen deficient and fluorine doped, this seems in consistent with the doping (25\% sum of oxygen vacancies and fluorine doping) in our sample. The calculated value of $n_d$ is 2.5 times larger than that of PrFeAsO$_{0.9}$ and NdFeAsO$_{0.9}$F$_{0.1}$ single crystals.
The field dependence of the critical current density, a plateau in the low-field followed by a power law decrease $J_c \propto H^{-5/8}$, is compatible with the theory of strong pinning \cite{beek1}. Similar field dependence of $J_c$ has also been observed in the case of PrFeAsO$_{1-y}$ single crystals \cite{beek}. In the presence of strong pins of size larger than the coherence length, it has been shown that \cite{beek1}
\begin{equation}
J_c(0)=\frac{\pi^{1/2}n_i^{1/2}j_0}{\varepsilon_{\lambda}}\left(\frac{f_{p,s}\xi_{ab}}{\epsilon_0}\right)^{3/2} (H < H^*),
\end{equation}
\begin{equation}
J_c(H)\approx\frac{2n_ij_0}{\varepsilon_{\lambda}^{5/4}\xi_{ab}^{1/2}}\left(\frac{f_{p,s}\xi_{ab}}{\epsilon_0}\right)^{9/4}\left(\frac{\Phi_0}{H}
\right)^{5/8} (H>H^*)
\end{equation}
where $n_i$ is the density of strong pins and $f_{p,s}$ is the elementary pinning force of a strong pin. The crossover field $H^*$=0.74$\varepsilon_{\lambda}^{-2}\Phi_0(n_i/\xi_{ab})^{4/5}(f_{p,s}\xi_{ab}/\epsilon_0)^{6/5}$ is determined as that above which the so called vortex trapping area of a single pin is limited by intervortex interactions. Using equation (2) and the power-law decrease of equation (3) we can determine $f_{p,s}$ from the ratio of [$dJ_c(H)/dH^{-5/8}$]/[$J_c(0)$]$^2$. Using the experimentally determined value of [$dJ_c(H)/dH^{-5/8}$]/[$J_c(0)$]$^2$ at 10 K and $\xi$ = 1 nm from our earlier work \cite{bhoi} along with the reported values of $\varepsilon_{\lambda}$ = 0.4, $j_0$ = 2$\times 10^{12}$ Am$^{-2}$ and $\epsilon_0$ = 3.2 $\times 10^{-12}$ Jm$^{-1}$ \cite{oka}, we have estimated the value of $f_{p,s} \approx$ 4$\times10^{-13}$ N for the present compound, which is two times larger than that for PrFeAsO$_{1-y}$ single crystal \cite{beek}. The density of pinning centers as estimated from $H^*$ is $n_i \approx$ 3$\times 10^{21}$ m$^{-3}$. van der Beek \emph{et al.} \cite{beek} attributed the source of strong pinning in PrFeAsO$_{1-y}$ and NdFeAsO$_{0.9}$F$_{0.1}$ single crystals to the extended (nm sized) pointlike inclusions or precipitates and to the spatial variations in the doping level on the scale of several dozen to 100 nm. They have also shown experimentally that in underdoped PrFeAsO, $T_c$ but also $J_c$ increase with an increasing density of oxygen vacancies, which corresponds to the introduction of more pinning centers. The larger values of pinning force, density of strong pins and higher $T_c$ in the present sample with higher doping level are consistent with their prediction.
The nonmonotonic behavior with a well pronounced second peak in magnetization represents the so-called "fishtail" effect. This phenomenon has been observed in conventional as well as high-temperature superconductors, but its origin is attributed to various mechanisms. The fishtail effect of comparable magnitude has also been observed recently in several iron arsenide superconductors \cite{prozo,kim,yang1,yama1,shen,kope,sena,beek,chen}. Most theoretical approaches agree that the temperature dependent field $H_p$, at which the second peak has its maximum, is related to vortex pinning and corresponds to a crossover between two different regimes of the vortex lattice. Various mechanisms have been proposed to explain the fishtail effect, foremost among them are a change in dynamics of the vortex lattice \cite{civa}, a change in flux creep behavior \cite{abu}, a change in its elastic properties \cite{blat} and a order-disorder transition of the vortex lattice due to the presence of topological defects \cite{kir,mik}. The peak effect in type-II superconductors was initially attributed by Pippard to the fact that the elastic constants of the vortex lattice vanish more rapidly than the pinning force as $H$ goes below $H_{c2}$ \cite{pippard}. Larkin and Ovchinnikov noted that this is exacerbated by the gradual softening of the vortex lattice because of the nonlocality of its tilt modulus \cite{larkin}. Both the above proposals were meant for the low-$T_c$ conventional superconductors where the peak effect occurs near the $H_{c2}(T)$ line. While for the present case the peak occurs for applied field well below the $H_{c2}(T)$ line which excludes the above scenario. The other models proposed to explain the second peak effect includes (i) a first order phase transition from an ordered "elastically pinned" low-field vortex phase, the so-called Bragg-glass, to a high-field disordered phase characterized by the presence of topological defects \cite{kir,mik} and (ii) a crossover from collective to plastic pinning. The former scenario has been verified in the case of high-temperature superconductors YBa$_2$Cu$_3$O$_{7-\delta}$ and Bi$_2$Sr$_2$CaCu$_2$O$_8$ \cite{koka,rass,beek3}, in the cubic superconductor (Ba,K)BiO$_3$ \cite{klein}, in NbSe$_2$ \cite{bhat}, in MgB$_2$ \cite{klein1} and recently, in PrFeAsO$_{1-y}$ and NdFeAsO$_{0.9}$F$_{0.1}$ single crystals \cite{beek}. However, discrepancies in the explanation of the second peak effect do exist in several cases. The peak effect in YBa$_2$Cu$_3$O$_{7-x}$ \cite{abu} and Nd$_{1.85}$Ce$_{0.15}$CuO$_{4-x}$ \cite{gill} has been ascribed to the crossover from collective to plastic pinning also. Majority of studies in 122 pnictide single crystals \cite{prozo,shen,salem} suggest that the peak effect is due to the crossover from collective to plastic pinning at $H$ = $H_p$.\\
\begin{figure}
\includegraphics[width=0.4\textwidth]{Fig6.pdf}\\
\caption{(color online) Temperature dependence of the calculated single-vortex to bundle pinning crossover field $H_{SV}$ and order-disorder field $H_{OD}$ at different temperatures with onset field $H_{on}$ data.}\label{Fig.6}
\end{figure}
In case of a crossover in vortex dynamics, the onset field $H_{on}$ should coincide with the single-vortex to bundle pinning crossover field $H_{SV}\sim 40H_{c2}\left(\frac{J^{SV}_c}{J_0}\right)$. However, if it is due to the order-disorder transition of the vortex lattice, then $H_{on}$ should coincide with the order-disorder transition $H_{OD}$ line. In the regime of single-vortex pining, the $H_{OD}$ is given by the following equation,
\begin{equation}
Ab_{SV}^{3/5}b_{OD}^{2/5}\left[1+\frac{F_T(t)}{b_{SV}^{1/2}(1-b_{OD})^{3/2}}\right]=2\pi c_L^2,
\end{equation}
where $b_{OD}\equiv H_{OD}/H_{c2}$, $b_{SV}\equiv H_{SV}/H_{c2}$, $c_L \sim$0.1 is the Lindemann number, A is a numerical constant, $t = T/T_c$, and $F_T(t) = 2t\left(\frac{Gi}{1-t^2}\right)^{1/2}$. Using $Gi$=0.01, $A$=4 and the experimental value of $J^{SV}_c$ at different temperatures, we have evaluated the $H_{SV}$ and $H_{OD}$ lines as shown in figure 6. Figure shows that $H_{on}$ data lie close to the $H_{OD}$ as compared to the $H_{SV}$ line which indicates the appearance of peak effect in this compound likely due to the order-disorder transition. A complete understanding of the second peak effect in pnictide superconductors needs further investigation.
\section{phase diagram}
In figure 7, we have drawn the $H$-$T$ phase diagram for the PrFeAsO$_{0.60}$F$_{0.12}$ superconductor as compiled from the magnetization and resistivity data \cite{bhoi1}. The magnetic upper critical field ($H_m$) determined from the superconducting onset transition of the $M$-$T$ curves under various applied fields is shown in the figure. For the sake of completeness, the resistive upper critical field $H_{c2}$ determined in our earlier measurement \cite{bhoi1} is also plotted in the phase diagram. The $H_m(T)$ curve appears closer to the resistive $H_{c2}(T)$ curve for 10$\%\rho_n$ criterion but far away from the $H_{c2}(T)$ curves for 90$\%\rho_n$ and 50$\%\rho_n$ criteria. It may be noted that the resistive transition reflects the net connectivity from one end to the other end of the sample through the superconducting domains while the magnetic measurement mainly reflects the bulk superconductivity when most of the domains turn superconducting. In this respect, it is more plausible that the $H_m(T)$ line will appear closer to the 10$\%\rho_n$ $H_{c2}(T)$ line rather than 90$\%\rho_n$ or 50$\%\rho_n$ $H_{c2}(T)$ line.
\begin{figure}
\includegraphics[width=0.4\textwidth]{Fig7.pdf}\\
\caption{(color online) The vortex phase diagram of the PrFeAsO$_{0.60}$F$_{0.12}$ superconductor in the $H$-$T$ plane
determined from the magnetization and resistivity data. The solid symbols describe $H_{c2}$ determined from 90$\%\rho_n$, 50$\%\rho_n$ and 10$\%\rho_n$ criteria. The solid lines are fit to the experimental $H_{on}$, $H_p$ and $H_{irr}$ data using Eq. (5) described in the text.}\label{Fig.7}
\end{figure}
A rough estimation of the irreversibility field $H_{irr}$ can be made from the $M(T)$ diagram where $M_{fc}(T)$ and $M_{zfc}(T)$ curves begin to diverge. Compared to the $H_m(T)$ line, the $H_{irr}(T)$ line shifts toward the lower temperature and lower field regions at a faster rate, indicating that there is a considerable gap between the upper critical field and the irreversibility field. Similar to cuprate superconductors, the $H_p(T)$ line is quite far from the $H_{c2}(T)$ line. In the phase diagram, we have also shown $H_{on}(T)$ and $H_{p}(T)$ curves. Normally, these characteristic fields, viz., $H_{irr}$, $H_{on}$ and $H_{p}$ follow a power-law expression of the form
\begin{equation}
H_x(T) = H_x(0)(1-T/T_c)^{n}.
\end{equation}
We observe that $H_{irr}(T)$, $H_{on}(T)$ and $H_p(T)$ data can be fitted well by the above expression with exponent $n \sim$ 1.7, 2.4 and 1.6, respectively. The values of the prefactors are: $H_{irr}(0)$ = 31.9 T, $H_{on}(0)$ = 4.9 T and $H_p(0) = 12.5$ T. Such a temperature dependence of these characteristics fields has been observed in high-$T_c$ cuprates \cite{baily}, (Ba,K)BiO$_3$ \cite{blanc} and SmFeAsO$_{0.8}$F$_{0.2}$ \cite{chen}. The $H$-$T$ phase diagram can be broadly divided into three regions: (i) vortex liquid phase above $H_{irr}$, where magnetic flux lines are not pinned due to strong thermal fluctuations; (ii) the disordered phase between $H_p$ and $H_{irr}$, and (iii) the elastic ordered phase below $H_p$ lines.
\section{conclusion}
In summary, we have analyzed the temperature and field dependence of magnetization of the PrFeAsO$_{0.60}$F$_{0.12}$ superconductor. The magnetization of the compound can be understood as a sum of the superconducting irreversible signal, $M_s$ originating from the pinning flux lines and a paramagnetic component, $M_p$ arising from the magnetic moment of Pr$^{+3}$ ion. The critical current density exhibits a second peak at a magnetic field well below $H_{c2}(T)$ line. The critical current density is almost independent of $H$ up to a field $H^*$ and then decreases rapidly with a power law $H^{-5/8}$ which is the characteristic of strong pinning. The evaluated elementary pinning force, $f_{p,s}$ and the density of pinning centers, $n_i$ are larger than those for PrFeAsO$_{0.9}$ single crystal. This indicates that the oxygen deficiency is useful to increase the critical current density. The $H$-$T$ phase diagram for the present system has been obtained using the magnetization and resistivity data.
\section{Acknowledgement}
The authors would like to thank A. Pal and S. Banerjee for technical help during sample preparation and measurements. A. Banerjee would also like to thank DST, India for financial assistance for the 14 T vibrating sample magnetometer facility at CSR, Indore.
|
\section{Introduction}
The discovery of frequent itemsets in transactional datasets is
a fundamental primitive that arises in the mining of
association rules and in many other scenarios
\cite{HanK01,TanSK06}. In its original formulation, the problem
requires that given a dataset ${\cal D}$ of transactions over a set of items
${\cal I}$, and a support threshold $s$, all itemsets $X \subseteq {\cal I}$ with
support at least $s$ in ${\cal D}$ (i.e., contained in at least $s$ transactions) be
returned. These high-support itemsets are referred to as
\emph{frequent itemsets}.
Since the pioneering paper by Agrawal et al.~\cite{AgrawalIS93}, a
vast literature has flourished, addressing variants of the problem,
studying foundational issues, and presenting novel algorithmic
strategies or clever implementations of known strategies (see, e.g.,
\cite{Fimi03,Fimi04}), but many problems remain open
\cite{HanCXY07}. In particular, assessing the significance of the
discovered itemsets, or equivalently, flagging statistically
significant discoveries with a limited number of false positive
outcomes, is still poorly understood and remains one of the most
challenging problems in this area.
The classical framework requires that the user decide what is
significant by specifying the support threshold $s$. Unless specific
domain knowledge is available, the choice of such a threshold is often
arbitrary \cite{HanK01,TanSK06} and may lead to a large number of
spurious discoveries (false positives) that would undermine the
success of subsequent analysis.
In this paper, we develop a rigorous and efficient novel approach for
identifying frequent itemsets featuring both a global and a pointwise
guarantee on their statistical significance. Specifically, we flag as
significant a population of itemsets extracted with respect to a
certain threshold, if some global characteristics of the population
deviate considerably from what would be expected if the dataset were
generated randomly with no correlations between items. Also, we make sure that
a large fraction of the itemsets belonging to the returned population
are individually significant by enforcing a small False Discovery Rate
(FDR)~\cite{BenjaminiH95} for the population.
\subsection{The model} \label{model}
As mentioned above, the significance of a discovery in our framework
is assessed based on its deviation from what would be expected in a
random dataset in which individual items are placed in transactions
independently.
Formally, let ${\cal D}$ be a dataset of $t$ transactions on a set ${\cal I}$ of
$n$ items, where each transaction is a subset of ${\cal I}$. Let $n(i)$ be
the number of transactions that contain item $i$ and let $f_i=n(i)/t$
be the {\emph{frequency}} of item $i$ in the dataset. The
{\emph{support}} of an itemset $X \subseteq {\cal I}$ is defined as the
number of transactions that contain $X$.
Following~\cite{SilversteinBM98}, the dataset ${\cal D}$ is associated with
a probability space of datasets, all featuring the same number of
transactions $t$ on the same set of items ${\cal I}$ as ${\cal D}$, and in which
item $i$ is included in any given transaction with probability $f_i$,
independently of all other items and all other transactions. A similar
model is used in \cite{PurdomVGG04} and \cite{SayrafiVGP05} to
evaluate the running time of algorithms for frequent itemsets mining.
For a fixed integer $k \geq 1$, among all possible $n\choose k$
itemsets of size $k$ (\emph{$k$-itemsets}) we are interested in
identifying statistically significant ones, that is, those $k$-itemsets
whose supports are significantly higher, in a statistical sense, than
their expected supports in a dataset drawn at random from the
aforementioned probability space.
An alternative probability space of datasets, proposed in
\cite{GionisMMT06}, considers all arrangements of $n$ items into $m$
transactions which match the exact item frequencies and transaction
lengths as ${\cal D}$. Conceivably, the technique of this paper could be
adapted to this latter model as well.
\subsection{Multi-hypothesis testing} \label{sec:multi-hp}
In a simple statistical test, a null hypothesis $H_0$ is tested against
an alternative hypothesis $H_1$. A test consists of a rejection
(critical) region $C$ such that, if the statistic (outcome) of the
experiment is in $C$, then the null hypothesis is rejected, and otherwise the
null hypothesis is not rejected. The {\it significance level} of a
test, $\alpha=\Pr$(Type I error), is the probability of rejecting $H_0$
when it is true (false positive). The {\it power} of the test,
$1-\Pr$(Type II error), is the probability of correctly rejecting the
null hypothesis. The {\it $p$-value} of a test is the probability of
obtaining an outcome at least as extreme as the one that was actually
observed, under the assumption that $H_0$ is true.
In a multi-hypothesis statistical test, the outcome of an experiment
is used to test simultaneously a number of hypotheses. For example, in
the context of frequent itemsets, if we seek significant $k$-itemsets,
we are in principle testing $n \choose k$ null hypotheses simultaneously,
where each null hypothesis corresponds to the support of a given
itemset not being statistically significant. In the context of
multi-hypothesis testing, the significance level cannot be assessed by
considering each individual hypothesis in isolation. To demonstrate
the importance of correcting for multiplicity of hypotheses, consider
a simple real dataset of 1,000,000 transactions over 1,000 items, each
with frequency 1/1,000. Assume that we observed that a pair of items
$(i,j)$ appears in at least 7 transactions. Is the support of this
pair statistically significant? To evaluate the significance of this
discovery we consider a random dataset where each item is included in
each transaction with probability 1/1,000, independent of all items.
The probability that the pair $(i,j)$ is included in a given
transaction is 1/1,000,000, thus the expected number of transactions
that include this pair is 1. A simple calculation shows that the
probability that $(i,j)$ appears in at least 7 transactions is about
0.0001. Thus, it seems that the support of $(i,j)$ in the real
dataset is statistically significant. However, each of the 499,500
pairs of items has probability 0.0001 to appear in at least 7
transactions in the random dataset. Thus, even under the assumption
that items are placed independently in transactions, the expected
number of pairs with support at least 7 is about 50. If there were
only about 50 pairs with support at least 7, returning the pair
$(i,j)$ as a statistically significant itemset would likely be a false
discovery since its frequency would be better explained by random
fluctuations in observed data. On the other hand, assume that the
real dataset contains 300 disjoint pairs each with support at least 7.
By the Chernoff bound \cite{mu05}, the probability of that event in
the random dataset is less than $2^{-300}$. Thus, it is very likely
that the support of most of these pairs would be statistically
significant. A discovery process that does not return these pairs will
result in a large number of false negative errors. Our goal is to
design a rigorous methodology which is able to distinguish between
these two scenarios.
A natural generalization of the significance level to multi-hypothesis
testing is the {\it Family Wise Error Rate (FWER)}, which is the
probability of incurring at least one Type I error in any of the
individual tests. If we have $m$ simultaneous tests and we want to
bound the FWER by $\alpha$, then the Bonferroni method tests each null
hypothesis with significance level $\alpha/m$. While controlling the
FWER, this method is too conservative in that the power of the test is
too low, giving many false negatives. There are a number of techniques
that improve on the Bonferroni method, but for large numbers of
hypotheses all of these techniques lead to tests with low power (see
\cite{Dudoit03} for a good review).
The {\it False Discovery Rate (FDR)} was suggested by Benjamini and
Hochberg~\cite{BenjaminiH95} as an alternative, less conservative
approach to control errors in multiple tests. Let $V$ be the number of
Type I errors in the individual tests, and let $R$ be the total number
of null hypotheses rejected by the multiple test. Then we define FDR
to be the expected ratio of erroneous rejections among all rejections,
namely FDR $= E[V/R]$, with $V/R=0$ when $R=0$. Designing a
statistical test that controls for FDR is not simple, since the FDR is
a function of two random variables that depend both on the set of null
hypotheses and the set of alternative hypotheses. Building on the
work of \cite{BenjaminiH95}, Benjamini and Yekutieli~\cite{BY01}
developed a general technique for controlling the FDR in any
multi-hypothesis test (see Theorem~\ref{by-test} in
Section~\ref{sec:chernoff}).
\subsection{Our Results}\label{sec:ourresults}
We address the classical problem of mining frequent itemsets with
respect to a certain minimum support threshold, and provide a rigorous
methodology to establish a threshold that guarantees, in a statistical
sense, that the returned family of frequent itemsets contains
significant ones with a limited FDR. Our methodology crucially relies
on the following Poisson approximation result, which is the main
theoretical contribution of the paper.
Consider a dataset ${\cal D}$ of $t$ transactions on a set ${\cal I}$ of $n$ items
and let $\hat{{\cal D}}$ be a corresponding random dataset according to
the random model described in Section~\ref{model}. Let $Q_{k,s}$
be the observed number of $k$-itemsets with support at least $s$
in ${\cal D}$, and let $\hat{Q}_{k,s}$ be the corresponding
random variable for $\hat{\cal D}$. We show that there exists a
minimum support value $s_{\min}$ (which depends on the parameters of
${\cal D}$ and on $k$), such that for all $s \ge s_{\min}$ the distribution of
$\hat{Q}_{k,s}$ is well approximated by a Poisson distribution. Our
result is based on a novel application of the Chen-Stein Poisson
approximation method \cite{ArratiaGG90}.
The minimum support $s_{\min}$ provides the grounds to devise a
rigorous method for establishing a support threshold for mining
significant itemsets, both reducing the overall complexity and
improving the accuracy of the discovery process. Specifically, for a
fixed itemset size $k$, we test a small number of support thresholds
$s \geq s_{\min}$, and, for each such threshold, we measure the
$p$-value corresponding to the null hypothesis $H_0$ that the observed
value $Q_{k,s}$ comes from a Poisson distribution of suitable
expectation. From the tests we can determine a threshold $s^*$ such
that, with user-defined significance level $\alpha$, the number of
$k$-itemsets with support at least $s^*$ is not sampled from a Poisson
distribution and is therefore statistically significant. Observe that
the statistical significance of the number of itemsets with support at
least $s^*$ does not imply necessarily that each of the itemsets is
significant. However, our test is also able to guarantee a
user-defined upper bound $\beta$ on the FDR among all discoveries. We
remark that our approach works for any fixed itemset size $k$, unlike
traditional frequent itemset mining, where itemsets of all sizes are
extracted for a given threshold.
To grasp the intuition behind the above approach, recall that a
Poisson distribution models the number of occurrences among a large
set of possible events, where the probability of each event is
small. In the context of frequent itemset mining, the Poisson
approximation holds when the probability that an individual itemset
has support at least $s_{\min}$ in $\hat{{\cal D}}$ is small, and thus the
existence of such an event in ${\cal D}$ is likely to be statistically
significant. We stress that our technique discovers statistically
significant itemsets among those of relatively high support. In fact,
if the expected supports of individual itemsets vary in a large range,
there may exist itemsets with very low expected supports in $\hat{{\cal D}}$
which may have statistically significant supports in ${\cal D}$. These
itemsets would not be discovered by our strategy. However, any mining
strategy aiming at discovering significant, low-support itemsets is
likely to incur high costs due to the large (possibly exponential)
number of candidates to be examined, although only a few of them would turn
out to be significant.
We validate our theoretical results by mining significant frequent
itemsets from a number of real datasets that are standard benchmarks
in this field. Also, we compare the performance of our methodology to
a standard multi-hypothesis approach based on \cite{BY01}, and provide
evidence that the latter often returns fewer significant itemsets,
which indicates that our method has considerably higher power.
\subsection{Related Work}
A number of works have explored various notions of significant
itemsets and have proposed methods for their discovery. Below, we review those
most relevant to our approach and refer the reader to
\cite[Section~3]{HanCXY07} for further references.
Aggarwal and Yu \cite{AggarwalY98} relate the significance of an itemset
$X$ to the
quantity $((1-v(X))/(1-\mathbf{E}[v(X)])) \cdot (\mathbf{E}[v(X)]/v(X))$, where $v(X)$
represents the fraction of transactions containing some but not all of
the items of $X$, and $\mathbf{E}[v(X)]$ represents the expectation of $v(X)$
in a random dataset where items occur in transactions
independently. This ratio provides an empirical measure of the
correlation among the items of $X$ that, according
to the authors, is more effective than absolute support. In
\cite{SrikantA96-1,DuMouchel99,DuMouchelP01}, the significance of an
itemset is measured as the ratio $R$ between its actual support and
its expected support in a random dataset. In order to make this
measure more accurate for small supports,
\cite{DuMouchel99,DuMouchelP01} propose smoothing the ratio $R$ using
an empirical Bayesian approach. Bayesian analysis is also employed in
\cite{SilberschatzT96} to derive subjective measures of significance
of patterns (e.g., itemsets) based on how strongly they ``shake'' a
system of established beliefs. In \cite{JaroszewiczS05}, the
significance of an itemset is defined as the absolute difference
between the support of the itemset in the dataset, and the estimate
of this support made from a Bayesian network with parameters derived from the dataset.
A statistical approach for identifying significant itemsets is
presented in \cite{SilversteinBM98}, where the measure of interest for
an itemset is defined as the degree of dependence among its
constituent items, which is assessed through a $\chi^2$ test.
Unfortunately, as reported in~\cite{DuMouchel99,DuMouchelP01}, there
are technical flaws in the applications of the statistical test in
\cite{SilversteinBM98}. Nevertheless, this work
pioneered the quest for a rigorous framework for addressing the
discovery of significant itemsets.
A common drawback of the aforementioned works is that they assess the
significance of each itemset \emph{in isolation}, rather than taking
into account the \emph{global} characteristics of the dataset from
which they are extracted. As argued before, if the number of itemsets
considered by the analysis is large, even in a purely random dataset
some of them are likely to be flagged as significant if considered in
isolation. A few works attempt at accounting for the global structure
of the dataset in the context of frequent itemset mining. The authors
of \cite{GionisMMT06} propose an approach based on Markov chains to
generate a random dataset that has identical transaction lengths and
identical frequencies of the individual items as the given real
dataset. The work suggests comparing the outcomes of a number of data
mining tasks, frequent itemset mining among the others, in the real
and the randomly generated datasets in order to assess whether the
real datasets embody any significant global structure. However, such
an assessment is carried out in a purely qualitative fashion without
rigorous statistical grounding.
The problem of spurious discoveries in the mining of significant patterns
is studied in \cite{BoltonHA02}. The paper is concerned with the
discovery of significant pairs of items, where significance is
measured through the $p$-value, that is, the probability of occurrence
of the observed support in a random dataset. Significant pairs are
those whose $p$-values are below a certain threshold that can be
suitably chosen to bound the FWER, or to bound the FDR. The authors
compare the relative power of the two metrics through experimental results,
but do not provide
methods to set a meaningful support threshold, which is the most
prominent feature of our approach.
Beyond frequent itemset mining, the issue of significance has also
been addressed in the realm of discovering association rules. In
\cite{HamalainenN08}, the authors provide a variation of the
well-known Apriori strategy for the efficient discovery of a subset
$\cal A$ of association rules with $p$-value below a given cutoff
value, while the results in \cite{MegiddoS98} provide the means of
evaluating the FDR in $\cal A$. The FDR metric is also employed in
\cite{ZhangPT04} for the discovery of significant
quantitative rules, a variation of association rules. None of these
works is able to establish support thresholds such that the returned
discoveries feature small FDR.
\subsection{Benchmark datasets}
\sloppy In order to validate the methodology, a number of experiments,
whose results are reported in Section~\ref{sec:experiments}, have been
performed on datasets which are standard benchmarks in the context of
frequent itemsets mining. The main characteristics of the datasets we
use are summarized in Table~\ref{tab:datasets}. A description of the
datasets can be found in the FIMI Repository ({\tt
http://fimi.cs.helsinki.fi/data/}), where they are available for
download.
\begin{table}
\begin{center}
\begin{tabular}{lcccc}
\hline
Dataset & $n$ & $\left[ f_{\min};f_{\max}\right]$ & $m$ & $t$\\
\hline
Retail & 16470 & $[$1.13e-05 ; 0.57$]$ & 10.3 & 88162 \\
Kosarak & 41270 & $[$1.01e-06 ; 0.61$]$ & 8.1 & 990002 \\
Bms1 & 497 & $[$1.68e-05 ; 0.06$]$ & 2.5 & 59602 \\
Bms2 & 3340 & $[$1.29e-05 ; 0.05$]$ & 5.6 & 77512 \\
Bmspos & 1657 & $[$1.94e-06 ; 0.60$]$ & 7.5 & 515597 \\
Pumsb$^*$ & 2088 & $[$2.04e-05 ; 0.79$]$ & 50.5 & 49046 \\
\hline
\end{tabular}
\end{center}
\caption{\boldmath Parameters of the benchmark
datasets: $n$ is the number of items; $[f_{\min},f_{\max}]$ is the
range of frequencies of the individual items; $m$ is the average
transaction length; and $t$ is the number of transactions.}
\label{tab:datasets}
\end{table}
\subsection{Organization of the Paper}\label{subsec:roadmap}
The rest of the paper is structured as follows.
Section~\ref{sec:poisson} presents the Poisson approximation
result for the random variable $\hat{Q}_{k,s}$. The methodology
for establishing the support threshold $s^*$ is
presented in Section~\ref{sec:methodology}, and experimental
results are reported in Section~\ref{sec:experiments}.
Section~\ref{sec:conclusions} ends the paper with some concluding
remarks.
\section{Poisson Approximation Result}\label{sec:poisson}
The Chen-Stein method \cite{ArratiaGG90} is a powerful tool for
bounding the error in approximating probabilities associated with a
sequence of dependent events by a Poisson distribution. To apply the
method to our case, we fix parameters $k$ and $s$, and define a
collection of ${n \choose k}$ Bernoulli random variables
$\{Z_X~|~X\subset {\cal I},~|X|=k\}$, such that $Z_X=1$ if the
$k$-itemset $X$ appears in at least $s$ transactions in the random
dataset $\hat{\cal D}$, and $Z_X=0$ otherwise. Also, let $p_X =
\Pr(Z_X = 1)$. We are interested in the distribution of
$\hat{Q}_{k,s}=\sum_{X:|X|=k} Z_X$.
For each set $X$ we define the \emph{neighborhood set} of $X$,
\[
I(X)=\{X'~|~X\cap X' \ne \emptyset, |X'| = |X|\}.
\]
If $Y \not\in I(X)$ then $Z_Y$ and $Z_X$ are independent.
The following theorem is a straightforward adaptation
of \cite[Theorem 1]{ArratiaGG90} to our case.
\begin{theorem}\label{arratia}
Let $U$ be a Poisson random variable such that
$\mathbf{E}[U]=\mathbf{E}[\hat{Q}_{k,s}]=\lambda<\infty$. The variation distance
between the distributions ${\cal L}(\hat{Q}_{k,s})$ of $\hat{Q}_{k,s}$
and ${\cal L}(U)$ of $U$ is such that
\begin{align*}
\left\| {\cal L}(\hat{Q}_{k,s})-{\cal L}(U) \right\|
&= \sup_A |\Pr(\hat{Q}_{k,s} \in A)-\Pr(U \in A)| \\
&\leq b_1+b_2,
\end{align*}
where
\[
b_1 = \sum_{X:|X|=k} \sum_{Y\in I(X)} p_X p_Y
\]
and
\[
b_2 = \sum_{X:|X|=k} \sum_{X\neq Y\in I(X)} \mathbf{E}[Z_X Z_Y].
\]
\end{theorem}
We can derive analytic bounds for $b_1$ and $b_2$ in many situations.
Specifically, suppose that we generate $t$ transactions in the
following way. For each item $x$, we sample a random variable $R_x
\in [0,1]$ independently from some distribution $R$. Conditioned on
the $R_x$'s, each item $x$ occurs independently in each transaction
with probability $R_x$. In what follows, we provide specific bounds
for this situation that depend on the moment $\mathbf{E}[R^{2s}]$ of the
random variable $R$.
As a warm-up, we first consider the specific case where each $R_x$ is
a fixed value $p = \gamma/n$ for some constant $\gamma$ for all $x$.
That is, each item appears in each transaction with a fixed
probability $p$, and the expected number of items per transaction is
constant. The more general case follows the same approach, albeit
with a few more technical difficulties.
\begin{theorem} \label{thm:poissapproxsimple}
Consider an asymptotic regime where as $n \to \infty$, we have $k,s =
O(1)$ with $s \ge 2$, each item appears in each transaction with probability
$p = \gamma/n$ for some constant $\gamma$,
and $t = O(n^c)$ for some positive constant $0 < c \leq (k-1)(1-1/s)$.
Let $U$ be a Poisson random variable such that
$\mathbf{E}[U]=\mathbf{E}[\hat{Q}_{k,s}]=\lambda<\infty$.
Then the variation distance between the distributions ${\cal
L}(\hat{Q}_{k,s})$ of $\hat{Q}_{k,s}$ and
${\cal L}(U)$ of $U$ satisfies
\iffalse
\begin{align*}
\left\| {\cal L}(\hat{Q}_{k,s})-{\cal L}(U) \right\|
&= \sup_A |\Pr(\hat{Q}_{k,s} \in A)-\Pr(U \in A)|\\
&\le b_1 + b_2,
\end{align*}
where
\begin{align*}
b_1 &= \left( \binom{n}{k}^2 - \binom{n}{k}\binom{n-k}{k} \right)\\&\quad {} \times
\left( \sum_{i=s}^t \binom{t}{i} \left(\frac{\gamma}{n}\right)^{ki} \left(1-\left(\frac{\gamma}{n}\right)^{k}\right)^{t-i} \right)^2\\
&= \Theta(n^{2cs + 2k(1-s)-1})
\end{align*}
and
\begin{align*}
b_2 &\le \sum_{g=1}^{k-1} \binom{n}{g ; k-g; k-g} \left(\frac{\gamma}{n}\right)^{2ks}\\&\qquad\qquad {} \times \sum_{i=0}^s \binom{t}{i;s-i;s-i} \left(\frac{n}{\gamma}\right)^{gi}\\
&= O(n^{2k(1-s) + s(k-1+c)-k+1}).
\end{align*}
\fi
\[
\left\| {\cal L}(\hat{Q}_{k,s})-{\cal L}(U) \right\| = O(1/n^{2s-2}).
\]
\end{theorem}
\begin{proof}
For a given set $X$ of $k$ items, let $p_{X,i}$ be the probability
that $S$ appears in exactly $i$ transactions, so that $p_X =
\sum_{i=s}^t p_{X,i}$ and
\[
p_{X,i} = \binom{t}{i} \left(\frac{\gamma}{n}\right)^{ki} \left(1-\left(\frac{\gamma}{n}\right)^k\right)^{t-i}.
\]
Applying Theorem~\ref{arratia} gives
\[
\left\|{\cal L}(\hat{Q}_{k,s})-{\cal L}(U)\right\| \leq b_1 + b_2
\]
where
\[
b_1 = \sum_{X:|X|=k} \sum_{Y\in I(S)} p_X p_Y
\]
and
\[
b_2 = \sum_{X:|S|=k} \sum_{Y\neq X\in I(S)} \mathbf{E}[Z_X Z_Y].
\]
We now evaluate $b_1$ and $b_2$. A direct calculation easily gives
the value for $b_1$ given in the statement of the theorem. For the
asymptotic analysis, we write
\begin{align*}
&\left(\binom{n}{k}^2 - \binom{n}{k}\binom{n-k}{k}\right)\\
&=\binom{n}{k}^2\left(1 - \frac{\binom{n-k}{k}}{\binom{n}{k}}\right)\\
&=\binom{n}{k}^2\left(1 - \prod_{i=0}^{k-1} \frac{n-k-i}{n-i}\right)\\
&=\Theta(n^k)^2 \cdot \Theta(1/n)
= \Theta(n^{2k-1})
\end{align*}
and
\begin{align*}
p_{X,s} &= \binom{t}{s} \left(\frac{\gamma}{n}\right)^{ks} \left(1-\left(\frac{\gamma}{n}\right)^{k}\right)^{t-s}\\
&=\Theta(t^s) \cdot \Theta(n^{-ks}) \cdot (1 + o(1))
=\Theta\left(t^{s} n^{-ks}\right),
\end{align*}
where we have used the fact that $t = o(n^k)$ to obtain the
asymptotics for the third term. Also, we note that for any $1 \le i
< t $
\[
\frac{p_{X,i+1}}{p_{X,i}}
= \frac{t-i}{i+1} \left(\frac{\gamma}{n}\right)^k\left(1-\left(\frac{\gamma}{n}\right)^k\right)^{-1}
\]
and so
\[
\max_{i \in \{s,s+1,\ldots,t-1\}} \frac{p_{X,i+1}}{p_{X,i}} = O(tn^{-k}) = O(1/n).
\]
Using a geometric series, it follows that
\[
p_X = \sum_{i=s}^t p_{X,i} = p_{X,s}(1 + o(1)) = \Theta\left(t^{s} n^{-ks}\right).
\]
Thus, we obtain
\begin{align*}
b_1 &= \Theta(n^{2k-1})\cdot\Theta\left(t^{s} n^{-ks}\right)^2\\
&= \Theta(t^{2s}n^{2k(1-s)-1})
= \Theta(n^{2cs + 2k(1-s)-1}).
\end{align*}
We now turn our attention to $b_2$. Consider sets $X \ne Y$ of $k$
items, let $g = |X \cap Y|$, and suppose that $g > 0$. Then if $Z_X
Z_Y = 1$, there exist disjoint subsets $A,B,C \in \{1,\ldots,t\}$
such that $0 \le |A| \le s$, $|B| = |C| = s-|A|$, all of the
transactions in $A$ contain both $X$ and $Y$, all of the transactions
in $B$ contain $X$, and all of the transactions in $C$ contain $Y$.
Therefore,
\[
\mathbf{E}[Z_X Z_Y]
\le \sum_{i=0}^s \binom{t}{i; s-i; s-i} \left(\frac{\gamma}{n}\right)^{(2k-g)i + 2k(s-i)},
\]
where the notation $\binom{m}{x; y; z}$ is a shorthand for
$\binom{m}{x}\binom{m-x}{y}\binom{m-x-y}{z}$.
It follows that
\begin{align*}
b_2
&\le \sum_{g=1}^{k-1} \binom{n}{g ; k-g; k-g} \\&\qquad {} \times \sum_{i=0}^s \binom{t}{i;s-i;s-i} \left(\frac{\gamma}{n}\right)^{(2k-g)i + 2k(s-i)}\\
&= \sum_{g=1}^{k-1} \binom{n}{g ; k-g; k-g} \left(\frac{\gamma}{n}\right)^{2ks} \\&\qquad {} \times \sum_{i=0}^s \binom{t}{i;s-i;s-i} \left(\frac{n}{\gamma}\right)^{gi}\\ &= \sum_{g=1}^{k-1} \binom{n}{g ; k-g; k-g} \left(\frac{\gamma}{n}\right)^{2ks} \\&\qquad {} \times \sum_{i=0}^s \binom{t}{i;s-i;s-i} \left(\frac{n}{\gamma}\right)^{gi}\\
&= \sum_{g=1}^{k-1} \Theta(n^{2k-g+2cs}) \left(\frac{\gamma}{n}\right)^{2ks} \sum_{i=0}^s n^{-ic} \left(\frac{n}{\gamma}\right)^{gi}\\
&= \Theta(n^{2k(1-s) + 2cs})\sum_{g=1}^{k-1} n^{-g} \sum_{i=0}^s {\gamma}^{-gi} n^{(g-c)i}\\
&= \Theta(n^{2k(1-s) + 2cs})\sum_{g=1}^{k-1} n^{-g} \begin{cases} \Theta(1) & g \le c\\ \Theta(n^{(g-c)s}) & g > c \end{cases}\\
&= \Theta(n^{2k(1-s) + 2cs}) \cdot \Theta(n^{-(k-1) + (k-1-c)s})\\
&= \Theta(n^{2k(1-s) + s(k-1+c)-k+1})
\end{align*}
Note that, in the summation where there are two cases depending on whether $g \leq c$ or $g > c$,
we have used the assumption that $c \leq (k-1)(1-1/s)$ to ensure the next equality.
Finally, it is simple to check that both $b_1$ and $b_2$ are $O(1/n^{2s-2})$ if
$c \leq (k-1)(1-1/s)$.
\end{proof}
We now provide the more general theorem.
\begin{theorem} \label{thm:poissapproxext}
Consider an asymptotic regime where as $n \to \infty$, we have $k,s =
O(1)$ with $s \ge 2$, $\mathbf{E}[R^{2s}] = O(n^{-a})$ for some constant $2 <
a \le 2s$, and $t = O(n^c)$ for some positive constant $c$. Let $U$
be a Poisson random variable such that
$\mathbf{E}[U]=\mathbf{E}[\hat{Q}_{k,s}]=\lambda<\infty$. If
\[
c \le \frac{(k-1)(a-2) + \min(2a-6,0)}{2s},
\]
then the variation distance between the distributions ${\cal
L}(\hat{Q}_{k,s})$ of $\hat{Q}_{k,s}$ and
${\cal L}(U)$ of $U$ satisfies
\[
\left\| {\cal L}(\hat{Q}_{k,s})-{\cal L}(U) \right\|
= O(1/n).
\]
\end{theorem}
\begin{proof}
Applying Theorem~\ref{arratia} gives
\[
\left\|{\cal L}(\hat{Q}_{k,s})-{\cal L}(U)\right\| \leq b_1 + b_2
\]
where
\[
b_1 = \sum_{X:|X|=k} \sum_{Y\in I(X)} p_X p_Y
\]
and
\[
b_2 = \sum_{X:|X|=k} \sum_{Y\neq X\in I(X)} \mathbf{E}[Z_X Z_Y].
\]
We now evaluate $b_1$ and $b_2$. Letting $\vec{R}$ denote the vector
of the $R_x$'s, we have that for any set $X$ of $k$ items
\[
\Pr(Z_X = 1\ |\ \vec{R}) \le \binom{t}{s} \prod_{x \in X} R_x^s.
\]
Since the $R_x$'s are independent with common distribution $R$,
\[
p_X = \mathbf{E}[\Pr(Z_X = 1\ |\ \vec{R})] \le \binom{t}{s} \mathbf{E}[R^s]^k.
\]
Using Jensen's inequality, we now have
\begin{align*}
b_1 &= \sum_{X:|X|=k} \sum_{Y\in I(X)} p_X p_Y\\
&\le \left(\binom{n}{k}^2 - \binom{n}{k}\binom{n-k}{k}\right) \binom{t}{s}^2 \mathbf{E}[R^s]^{2k}\\
&\le \binom{n}{k}^2\left(1 - \frac{\binom{n-k}{k}}{\binom{n}{k}}\right) \binom{t}{s}^2 \mathbf{E}[R^{2s}]^k\\
&=\binom{n}{k}^2\left(1 - \prod_{i=0}^{k-1} \frac{n-k-i}{n-i}\right) \binom{t}{s}^2 \mathbf{E}[R^{2s}]^k\\
&=\Theta(n^k)^2 \cdot \Theta(1/n) \cdot O(n^{2cs}) \cdot O(n^{-ka})\\
&= O(n^{k(2-a) + 2cs - 1})\\
\end{align*}
We now turn our attention to $b_2$. Consider sets $X \ne Y$ of $k$
items, and suppose $g = |X \cap Y|>0$. If $Z_X
Z_Y = 1$, there exist disjoint subsets $A,B,C \in \{1,\ldots,t\}$
such that $0 \le |A| \le s$, $|B| = |C| = s-|A|$, all of the
transactions in $A$ contain both $X$ and $Y$, all of the transactions
in $B$ contain $X$, and all of the transactions in $C$ contain $Y$.
Therefore,
\begin{align*}
\mathbf{E}[Z_X Z_Y\ |\ \vec{R}]
&\le \sum_{i=0}^s \binom{t}{i; s-i; s-i}
\left(\prod_{x \in X \cup Y} R_x^i\right) \\
& ~~~~~~~~~~~~~~~\times \left(\prod_{x \in X} R_x^{s-i}\right)
\left(\prod_{y \in Y} R_y^{s-i}\right)\\
&= \sum_{i=0}^s \binom{t}{i; s-i; s-i}
\left(\prod_{x \in X \cap Y} R_x^{2s-i}\right) \\
& ~~~~~~~~~~~~~~~\times \left(\prod_{x \in X - Y} R_x^s\right)
\left(\prod_{y \in Y - X} R_y^s\right).
\end{align*}
Applying independence of the $R_x$'s and Jensen's inequality gives
\begin{align*}
\mathbf{E}[Z_X Z_Y]
&= \mathbf{E}[\mathbf{E}[Z_X Z_Y\ |\ \vec{R}]]\\
&\le \sum_{i=0}^s \binom{t}{i; s-i; s-i} \mathbf{E}[R^{2s-i}]^g \mathbf{E}[R^s]^{2(k-g)}\\
&\le \sum_{i=0}^s t^{2s-i} \mathbf{E}[R^{2s}]^{\frac{g(2s-i)}{2s}} \mathbf{E}[R^{2s}]^{k-g}\\
&= \sum_{i=0}^s t^{2s-i} \mathbf{E}[R^{2s}]^{k - ig/2s}\\
&\le O(1) \sum_{i=0}^s n^{(2s-i)c - a\left(k - ig/2s\right)}\\
&= O(n^{2sc - ak}) \sum_{i=0}^s n^{i\left(\frac{ag}{2s} - c\right)}\\
&= O\left(n^{2sc - ak + \max\left\{0, s\left(\frac{ag}{2s} - c\right)\right\} }\right) \\
\end{align*}
It follows that
\begin{align*}
b_2 &\le \sum_{g=1}^{k-1} \binom{n}{g;k-g;k-g} O\left(n^{2sc - ak + \max\left\{0, s\left(\frac{ag}{2s} - c\right)\right\} }\right) \\
&= O(n^{2k + 2sc - ak}) \sum_{g=1}^{k-1} n^{-g} O\left(n^{\max\left\{0, s\left(\frac{ag}{2s} - c\right)\right\} }\right) \\
\end{align*}
Now, for $2sc/a < g < k$, we have (using the fact that $a \ge 2$)
\begin{align*}
n^{-g}n^{\max\left\{0, s\left(\frac{ag}{2s} - c\right)\right\}}
= n^{g(\frac{a}{2} - 1)-sc} \le n^{(k-1)(\frac{a}{2} - 1)-sc}.
\end{align*}
Thus
\[
b_2 = O(n^{2k + sc - ak + (k-1)(\frac{a}{2} - 1)}).
\]
(Here we
are using the fact that our choice of $c$ satisfies $c \le (k-1)(a -
2)/2s$ to ensure that $n^{(k-1)(\frac{a}{2} - 1) - cs} = \Omega(1)$.)
Now, we have
\[
b_1 = O(1/n)
\]
since
\[
c \le \frac{(k-1)(a-2)}{2s} \le \frac{k(a-2)}{2s},
\]
and
\[
b_2 = O(1/n)
\]
since
\[
c \le \frac{k(a-2) + (a-4)}{2s}.
\]
Thus
\[
b_1 + b_2 =
O(1/n).
\]
\end{proof}
It is easy to see that for fixed $k$, the quantities $b_1$ and $b_2$
defined in Theorem~\ref{arratia} are both decreasing in $s$. In the
following, we will use the notation $b_1(s)$ and $b_2(s)$ to indicate
explicitly that both quantities are functions of $s$. Therefore, for
a chosen $\epsilon$, with $0 < \epsilon < 1$, we can define
\begin{equation} \label{smin}
s_{\min} = \min \{s \geq 1 \; : \; b_1(s)+b_2(s) \leq \epsilon \}.
\end{equation}
It immediately follows that for every $s$ in the range $[s_{\min},\infty)$,
the variation distance between the distribution of $\hat{Q}_{k,s}$ and
the distribution of a Poisson variable with the same expectation is
less than $\epsilon$. In other words, for every $s \geq s_{\min}$ the
number of $k$-itemsets with support at least $s$ is well approximated
by a Poisson variable. Theorems~\ref{thm:poissapproxsimple} and
\ref{thm:poissapproxext} proved above establish
the existence of meaningful ranges of
$s$ for which the Poisson approximation holds, under certain
constraints on the individual item frequencies in the random dataset
and on the other parameters.
\subsection{A Monte Carlo method for determining $s_{\min}$}
While the analytical results of the previous subsection require that
the individual item frequencies in the random dataset be drawn from a
given distribution, in what follows we give experimental evidence
that the Poisson approximation for the distribution of $\hat{Q}_{k,s}$
holds also when the item frequencies are fixed arbitrarily, as is the
case of our reference random model. More specifically, we present a
method which approximates the support threshold $s_{\min}$ defined by
Equation~\ref{smin}, based on a simple Monte Carlo simulation which
returns estimates of $b_1(s)$ and $b_2(s)$. This approach is also
convenient in practice since it avoids the inevitable slack due to the
use of asymptotics in Theorem~\ref{thm:poissapproxext}.
For a given configuration of item frequencies and number of transactions,
let $\tilde{s}$ be the maximum expected support of any $k$-itemset
in a random dataset sampled according to that configuration,
that is, the product of the $k$ largest item frequencies.
Conceivably, the value $b_1(\tilde{s})$ is rather large,
hence it makes sense to search for an $s_{\min}$ larger than $\tilde{s}$.
We generate $\Delta$ random datasets and
from each such dataset we mine all of the $k$-itemsets of support at
least $\tilde{s}$. Let $W$ be the set of itemsets extracted in this
fashion from all of the generated datasets. For
each $s \geq \tilde{s}$ we can estimate $b_1(s)$ and $b_2(s)$ by
computing for each $X \in W$ the empirical probability $p_X$ of the
event $Z_X=1$, and for each pair $X,Y \in W$, with $X \cap Y \neq
\emptyset$, the empirical probability $p_{X,Y}$ of the event $(Z_X=1)
\wedge (Z_Y=1)$. Note that for itemsets not in $W$ these probabilities
are estimated as 0. If it turns out that
$b_1(\tilde{s})+b_2(\tilde{s}) > \epsilon/4$, then we let $\hat{s}_{\min}$
be the minimum $s > \tilde{s}$ such that $b_1(s)+b_2(s) \leq
\epsilon/4$. Otherwise, if $b_1(\tilde{s})+b_2(\tilde{s}) \leq
\epsilon/4$, we repeat the above procedure starting from $\tilde{s}/2$.
(Based on the above considerations this latter case will be unlikely.)
Algorithm 1 implements the above ideas.
The following theorem provides
a bound on the probability that $\hat{s}_{\min}$ be a conservative
estimate of $s_{\min}$, that is, $\hat{s}_{\min} \geq s_{\min}$.
\begin{theorem}
If $\Delta=O\left(\log(1/\delta)/\epsilon\right)$,
the output $\hat{s}_{\min}$ of the Monte-Carlo process satisfies
\[
\Pr(b_1(\hat{s}_{\min})+b_2(\hat{s}_{\min}) \leq \epsilon)\geq 1-\delta.
\]
\end{theorem}
\begin{proof}
Let assume $b_1(\hat{s}_{\min}) + b_2(\hat{s}_{\min}) > \epsilon$. Note that
$b_1(\hat{s}_{\min}) \leq b_2(\hat{s}_{\min})$, therefore we have
$b_2(\hat{s}_{\min}) > \epsilon/2$.
Let $B$ be the random variable corresponding to $\Delta$ times the estimate
of $b_2(\hat{s}_{\min})$ obtained with Algorithm
1. Thus $E[B] > \Delta \epsilon / 2$. Since Algorithm
1 returns $\hat{s}_{\min}$ as estimate of $s_{\min}$, we have that $B \le
\Delta \epsilon/4$.
Let
\[
\Delta = \frac{8 \log(1/\delta)}{\epsilon},
\]
and $c < 1$ be such that:
\[
(1-c)E[B] = \Delta \epsilon/4.
\]
Since $E[B] > \Delta \epsilon / 2$, we have $c \geq 1/2$.
Using Chernoff bound, we have that:
\begin{align*}
\Pr( B \le \Delta \epsilon/4 ) & \leq e^{- \frac{c^2 E[B]}{2}} \\
& \leq e^{- \frac{1}{4}\frac{ 8 \log(1/\delta)}{2}} \leq \delta.
\end{align*}
Thus $\Pr(b_1(\hat{s}_{\min}) + b_2(\hat{s}_{\min}) > \epsilon) \leq
\delta$.
\end{proof}
\begin{algorithm}[h]
\caption{FindPoissonThreshold}
\label{alg:montecarlo}
\begin{algorithmic}[1]
\REQUIRE Dataset ${\cal D}$ of $t$ transactions over $n$ items,
vector $\vec{f}$ of item frequencies, $k$, $\Delta$, $\varepsilon$;
\ENSURE Estimate $\hat{s}_{\min}$ of $s_{\min}$;
\STATE $\tilde{s} \gets $ highest expected support of a $k$-itemset;
\STATE $s_{\max} \gets 0$;
\STATE $W \gets \emptyset$;\label{line:alg}
\FOR {$i \gets 1$ to $\Delta$} \label{line2:alg}
\STATE $\hat{{\cal D}}_i \gets$ random dataset with parameters
$t$,$n$,$\vec{f}$;
\STATE $W \gets W \cup \left\{ \mbox{frequent }k\mbox{-itemsets in
}\hat{{\cal D}_i} \mbox{ w.r.t. }\tilde{s} \right\}$;
\ENDFOR
\IF {$W = \emptyset$}
\STATE $\tilde{s} \gets \tilde{s}/2$;
\STATE {\bf goto} \ref{line2:alg};
\ENDIF
\IF {($s_{\max} =0$)}
\STATE $s_{\max} \gets \displaystyle \max_{X \in W, \hat{{\cal D}}_i }\left\{
\mbox{support of }X \mbox{ in } \hat{{\cal D}}_i\right\} + 1$;
\ENDIF
\FOR{$s\gets \tilde{s}$ to $s_{\max}$}
\FORALL {$X \in W$}
\STATE $p_X(s) \gets $ empirical probability of \{$Z_X=1$\};
\ENDFOR
\FORALL {$X,Y \in W: X \cap Y \neq \emptyset$}
\STATE $p_{X,Y}(s) \gets$ empirical probability of \{$Z_{X,Y}=1$\};
\ENDFOR
\STATE $b_{1}(s) \gets \displaystyle \sum_{X,Y \in W; Y\in I(X)} p_X(s)
p_Y(s)$;
\STATE $b_{2}(s) \gets \displaystyle \sum_{X,Y \in W; X \neq Y\in I(X)}
p_{X,Y}(s)$;
\ENDFOR
\IF {$b_{1}(\tilde{s})+b_{2}(\tilde{s}) \leq \varepsilon/4$}
\STATE $s_{\max} \gets \tilde{s}$;
\STATE $\tilde{s} \gets \tilde{s}/2$;
\STATE {\bf goto} \ref{line:alg};
\ENDIF
\STATE $\hat{s}_{\min} \gets \min\left\{s > \tilde{s}: b_{1}(s) + b_{2}(s)
\leq
\varepsilon/4 \right\}$;
\STATE {\bf return} $\hat{s}_{\min}$;
\end{algorithmic}
\end{algorithm}
\setcounter{algorithm}{0}
For each dataset ${\cal D}$ of Table~\ref{tab:datasets} and for itemset
sizes $k=2,3,4$, we applied Algorithm 1 setting $\Delta = 1,000$ and
$\epsilon=0.01$. The values of $\hat{s}_{\min}$ we obtained are reported in
Table~\ref{fig:realpoissapprox} (we added the prefix ``Rand'' to each
dataset name, to denote the fact that the dataset is random and
features the same parameters as the corresponding real one).
\begin{table}[ht]
\begin{center}
\begin{tabular}{lccc}
\hline
& \multicolumn{3}{c}{ $\hat{s}_{\min}$ }\\
\cline{2-4}
Dataset & $k = 2$ & $ k = 3$ & $ k = 4$\\
\hline
RandRetail & 9237 & 4366 & 784 \\
RandKosarak & 273266 & 100543 & 20120 \\
RandBms1 & 268 & 23 & 5 \\
RandBms2 & 168 & 13 & 4 \\
RandBmspos & 76672 & 15714 & 2717 \\
RandPumsb$^*$ & 29303 & 21893 & 16265 \\
\hline
\end{tabular}
\end{center}
\caption{\boldmath Values of $\hat{s}_{\min}$ for
$\epsilon = 0.01$ and for $k=2,3,4$, in random datasets with the same
values of $n$, $t$, and with the same frequencies of the items as
the corresponding benchmark datasets.} \label{fig:realpoissapprox}
\end{table}
\section{Procedures for the discovery of high-support
significant itemsets} \label{sec:methodology}
For a give itemset size $k$, the value $s_{\min}$ identifies a region
of (relatively high) supports where we concentrate our quest for
statistically significant $k$-itemsets. In this section we develop
procedures to identify a family of $k$-itemsets (among those of
support greater than or equal to $s_{\min}$) which are statistically
significant with a controlled FDR. More specifically, in
Subsection~\ref{sec:chernoff} we show that a family with the desired
properties can be obtained as a subset of the frequent $k$-itemsets
with respect to $s_{\min}$, selected based on a standard
multi-comparison test. However, the returned family may turn out to be
too small (i.e., the procedures yields a large number of false
negatives). To achieve higher effectiveness, in
Subsection~\ref{sec:sstar} we devise a more sophisticated procedure
which identifies a support threshold $s^* \geq s_{\min}$ such that
\emph{all} frequent $k$-itemsets with respect to $s^*$ are
statistically significant with a controlled FDR. In the next section
we will provide experimental evidence that in many cases the latter
procedure yields much fewer false negatives.
\subsection{A procedure based on a standard multi-comparison
test}\label{sec:chernoff}
We present a first, simple procedure to discover significant itemsets with
controlled FDR, based on the following well established
result in multi-comparison testing.
\begin{theorem}[\cite{BY01}]
\label{by-test}
Assume that we are testing for $m$ null hypotheses. Let $p_{(1)}\leq
p_{(2)}\leq \dots\leq p_{(m)}$ be the ordered observed $p$-values of
the $m$ tests. For a given parameter $\beta$, with $0 < \beta < 1$, define
\begin{eqnarray}
\label{eqn:by}
\ell=\max\left\{i \geq 0: p_{(i)}\leq \frac{i}{m\sum_{j=1}^m \frac{1}{j}} \beta\right\},
\end{eqnarray}
and reject the null hypotheses corresponding to tests
$(1),\dots,(\ell)$. Then, the FDR for the set of rejected null
hypotheses is upper bounded by $\beta$.
\end{theorem}
Let ${\cal D}$ denote an input dataset consisting of $t$ transactions over
$n$ items, and let $k$ be the fixed itemset size. Recall that $s_{\min}$
is the minimum support threshold for which the distribution of
$\hat{Q}_{k,s}$ is well approximated by a Poisson distribution.
First, we mine from ${\cal D}$ the set of frequent $k$-itemsets ${\cal
F}_{(k)}(s_{\min})$. Then, for each $X \in {\cal
F}_{(k)}(s_{\min})$, we test the null hypothesis $H_0^X$ that the
observed support of $X$ in ${\cal D}$ is drawn from a Binomial distribution
with parameters $t$ and $f_X$ (the product of the individual
frequencies of the items of $X$), setting the rejection threshold as
specified by condition (\ref{eqn:by}), with parameters $\beta$ and $m
= {n \choose k}$. Based on Theorem~\ref{by-test}, the itemsets of
${\cal F}_{(k)}(s_{\min})$ whose associated null hypothesis is
rejected can be returned as significant, with FDR upper bounded by
$\beta$. The pseudocode Procedure~\ref{alg:fdrbino} implements the
strategy described above.
\floatname{algorithm}{Procedure}
\begin{algorithm}
\caption{}
\label{alg:fdrbino}
\begin{algorithmic}
\REQUIRE Dataset ${\cal D}$ of $t$ transactions over $n$ items,
vector $\vec{f}$ of item frequencies, $k$, $\beta \in (0,1)$;
\ENSURE Family of significant $k$-itemsets with FDR $\leq \beta$;
\STATE Determine $s_{\min}$ and compute ${\cal F}_{(k)}(s_{\min})$ from ${\cal D}$;
\FORALL {$X \in {\cal F}_{(k)}(s_{\min})$}
\STATE $s_X \gets $ support of $X$ in ${\cal D}$;
\STATE $f_X \gets \Pi_{i\in X}f_i$;
\STATE $p^{(X)} \gets \Pr(\mbox{Bin}(t , f_X) \geq s_X)$;
\ENDFOR
\STATE Let $p_{(1)}, p_{(2)}, \dots ,$ be the sorted sequence of the values
$p^{(X)}$, with $X\in {\cal F}_{(k)}(s_{\min})$;
\STATE $m \gets {n \choose k}$;
\STATE $\ell = \max\left\{0, i: p_{(i)}\leq \frac{i}{m\sum_{j=1}^m \frac{1}{j}}\beta\right\}$;
\RETURN $\left\{ X \in {\cal F}_{(k)}(s_{\min}): p^{(X)} = p_{(i)}, 1\leq i\leq \ell \right\}$;
\end{algorithmic}
\end{algorithm}
\subsection{Establishing a support threshold for
significant frequent itemsets} \label{sec:sstar}
Let $\alpha$ and $\beta$ be two constants in $(0,1)$. We seek a
threshold $s^*$ such that, with confidence $1-\alpha$, the
$k$-itemsets in ${\cal F}_{(k)}(s^*)$ can be flagged as statistically
significant with FDR at most $\beta$. The threshold $s^*$ is
determined through a robust statistical approach which ensures that
the number $Q_{k,s^*} = |{\cal F}_{(k)}(s^*)|$ deviates significantly
from what would be expected in a random dataset, and that the
magnitude of the deviation is sufficient to guarantee the bound on the
FDR.
Let $s_{\min}$ be the minimum support such that the Poisson
approximation for the distribution of $\hat{Q}_{k,s}$ holds for $s
\geq s_{\min}$, and let $s_{\max}$ be the maximum support of an item
(hence, of an itemset) in ${\cal D}$. Our procedure performs $h=\lfloor
\log_2 (s_{\max}-s_{\min}) \rfloor + 1$ comparisons. Let
$s_0=s_{\min}$ and $s_i =s_{\min}+2^i$, for $ 1 \leq i < h$. In the
$i$-th comparison, with $0 \leq i < h$, we test the null hypothesis
$H_0^i$ that the observed value $Q_{k,s_i}$ is drawn from the same
Poisson distribution as $\hat{Q}_{k,s_i}$. We choose as $s^*$ the
minimum of the $s_i$'s, if any, for which the null hypothesis $H_0^i$
is rejected.
For the correctness of the above procedure, it is crucial to specify a
suitable rejection condition for each $H_0^i$. Assume first that, for
$0 \leq i < h$, we reject the null hypothesis $H_0^i$ when the
$p$-value of the observed value $Q_{k,s_i}$ is smaller than
$\alpha_i$, where the $\alpha_i$'s are chosen so that
$\sum_{i=0}^{h-1} \alpha_i = \alpha$. Then, the union bound shows that
the probability of rejecting any true null hypothesis is less than
$\alpha$. However, this approach does not yield a bound on the FDR for
the set ${\cal F}_{(k)}(s^*)$. In fact, some itemsets in ${\cal
F}_{(k)}(s^*)$ are likely to occur with high support even under
$H_0^i$, hence they would represent false discoveries. The impact of
this phenomenon can be contained by ensuring that the FDR is below a
specified level $\beta$. To this purpose, we must strengthen the
rejection condition, as explained below.
Fix suitable values $\beta_0, \beta_1, \ldots, \beta_{h-1}$ such that
$\sum_{i=0}^{h-1} \beta_i^{-1} \leq \beta$. For $0 \leq i < h$, let
$\lambda_i = E[\hat{Q}_{k,s_i}]$. We now reject $H_0^i$ when the
$p$-value of $Q_{k,s_i}$ is smaller than $\alpha_i$, \emph{and}
$Q_{k,s_i} \geq \beta_i \lambda_i$. The following theorem establishes
the correctness of this approach.
\begin{theorem} \label{correctness-p2}
With confidence $1-\alpha$, ${\cal F}_{(k)}(s^*)$
is a family of statistically significant frequent
$k$-itemsets with FDR at most $\beta$.
\end{theorem}
\begin{proof}
Observe that since $\sum_{i=0}^{h-1} \alpha_i \leq \alpha$, we have
that all rejections are correct, with probability at least
$1-\alpha$. Let $E_i$ be the event \textit{``$H^i_0$ is rejected''} or
equivalently, \textit{``the $p$-value of $Q_{k,s_i}$ is smaller than
$\alpha_i$ and $Q_{k,s_i} \geq \beta_i \lambda_i$''}. Suppose that
$H^i_0$ is the first rejected null hypothesis, for some index $i$,
whence $s^*= s_i$. In this case, $Q_{k,s_i}$ itemsets are flagged as
significant. We denote by $V_i$ the number of false discoveries among
these $Q_{k,s_i}$ itemsets. It is easy to argue that the expectation
of $V_i$ is upper bounded by $E[X_i | E_i,
\bar{E}_{i-1},\dots,\bar{E}_0]$, where $X_i$ is a Poisson variable
with expectation $\lambda_i$. Since $Q_{k,s_i} \geq \beta_i \lambda_i$
when $H^i_0$ is rejected, by the law of total probability we have
\begin{eqnarray*}
FDR & \leq & \sum_{i=0}^{h-1} E\left[\frac{V_i}{Q_{k,s_i}}\right]
\Pr(E_i,\bar{E}_{i-1},\dots,\bar{E}_0) \\
& \leq & \sum_{i=0}^{h-1} \frac{E\left[V_i\right]}{\beta_i \lambda_i}
\Pr(E_i,\bar{E}_{i-1},\dots,\bar{E}_0) \\
& \leq & \sum_{i=0}^{h-1}
\frac{E[X_i~|~E_i \bar{E}_{i-1},\dots,\bar{E}_0]}{\beta_i \lambda_i}
\Pr(E_i,\bar{E}_{i-1},\dots,\bar{E}_0) \\
& = & \sum_{i=0}^{h-1}
\frac{\sum_{j\geq 0}j \Pr(X_i=j,E_i,\bar{E}_{i-1},\dots,\bar{E}_0)}
{\beta_i \lambda_i} \\
& \leq & \sum_{i=0}^{h-1} \frac{\lambda_i}{\beta_i \lambda_i} =
\sum_{i=0}^{h-1} \frac{1}{\beta_i} \leq \beta.
\end{eqnarray*}
\end{proof}
The pseudocode Procedure~\ref{alg:test1} specifies more formally our approach
to determine the support threshold $s^*$. Note that estimates for the
$\lambda_i$'s needed in the for-loop of Lines 7-9 can be obtained from
the same random datasets generated in Algorithm~\ref{alg:montecarlo},
which are used there for the estimation of $s_{\min}$.
\begin{algorithm}
\caption{}
\label{alg:test1}
\begin{algorithmic}[1]
\REQUIRE Dataset ${\cal D}$ of $t$ transactions over $n$ items,
vector $\vec{f}$ of item frequencies, $k$, $\alpha, \beta \in (0,1)$;
\ENSURE $s^*$ such that, with confidence $1-\alpha$,
${\cal F}_{(k)}(s^*)$ is a family of significant $k$-itemsets
with FDR $\leq \beta$;
\STATE Determine $s_{\min}$ and compute ${\cal F}_{(k)}(s_{\min})$ from ${\cal D}$;
\STATE $s_{\max}\gets$ maximum support of an item;
\STATE $i \gets 0$; $s_0\gets s_{\min}$;
\STATE $h \gets \lfloor \log_2 (s_{\max}-s_{\min}) \rfloor+1$;
\STATE Fix $\alpha_0,\dots,\alpha_{h-1} \in (0,1)$ s.t.
$\sum_{i=0}^{h-1}\alpha_i = \alpha$;
\STATE Fix $\beta_0,\dots,\beta_{h-1} \in (0,1)$ s.t.
$\sum_{i=0}^{h-1}\beta_i^{-1} = \beta$;
\FOR{$i \gets 0$ to $h-1$}
\STATE Compute $\lambda_i = E[\hat{Q}_{k,s_i}]$;
\ENDFOR
\WHILE{ $i < h$ }
\STATE Compute $Q_{k,s_i}$;
\IF {$(\Pr(\mbox{Poisson}(\lambda_i) \geq Q_{k,s_i}) \leq \alpha_i)$
{\bf and} ($Q_{k,s_i}\geq \beta_i \lambda_i$)}
\RETURN $s^* \gets s_i$;
\ENDIF
\STATE $s_{i+1} \gets s_{\min} + 2^{i+1} $;
\STATE $i \gets i+1$;
\ENDWHILE
\RETURN $s^* \gets \infty$;
\end{algorithmic}
\end{algorithm}
\section{Experimental Results} \label{sec:experiments}
In order to show the potential of our approach, in this section we
report on a number of experiments performed on the benchmark datasets
of Table~\ref{tab:datasets}. First, in Subsection~\ref{sec:realdata},
we validate experimentally the methodology implemented by
Procedure~\ref{alg:test1}, while in Subsection~\ref{sec:chernoffresults}, we
compare Procedure~\ref{alg:test1} against the more standard
Procedure~\ref{alg:fdrbino}, with respect to their ability to
discover significant itemsets.
\subsection{Experiments on benchmark datasets} \label{sec:realdata}
For each benchmark dataset in Table~\ref{tab:datasets} and for
$k=2,3,4$, we apply Procedure~\ref{alg:test1} with
$\alpha=\beta=0.05$, and $\alpha_i=\beta_{i}^{-1}=0.05/h$. The results
are displayed in Table~\ref{res:test3}, where, for each dataset and
for each value of $k$, we show: the support $s^*$ returned by
Procedure~\ref{alg:test1}, the number $Q_{k,s^*}$ of $k$-itemsets with
support at least $s^*$, and the expected number $\lambda(s^*)$ of
itemsets with support at least $s^*$ in a corresponding random
dataset.
\begin{table*}[ht]
\begin{center}
{\footnotesize
\begin{tabular}{lccccccccccc}
\hline
& \multicolumn{3}{c}{$k=2$} & &\multicolumn{3}{c}{$k=3$}
& & \multicolumn{3}{c}{$k=4$} \\
\cline{2-4} \cline{6-8} \cline{10-12}
Dataset & $s^*$ & $Q_{k,s^*}$ & $\lambda(s^*)$ & & $s^*$ & $Q_{k,s^*}$ &
$\lambda(s^*)$ & & $s^*$ & $Q_{k,s^*}$ & $\lambda(s^*)$ \\
\hline
Retail & $\infty$ & 0 & 0 & & $\infty$ & 0 & 0 & & 848 & 6 & 0.01 \\
Kosarak & $\infty$ & 0 & 0 & & $\infty$ & 0 & 0 & & 21144 & 12 & 0.01 \\
Bms1 & 276 & 56 & 0.19 & & 23 & 258859 & 0.06 & & 5 & ~27M & 0.05 \\
Bms2 & 168 & 429 & 0.73 & & 13 & 36112 & 0.25 & & 4 & 714045 & 0.01 \\
Bmspos & $\infty$ & 0 & 0 & & 16226 & 22 & 0.01 & & 2717 & 891 & 0.38 \\
Pumsb$^*$ & 29303 & 29 & 0.05 & & 21893 & 406 & 0.35 & & 16265 & 6293 & 1.37
\\
\hline
\end{tabular}
}
\end{center}
\caption{Results obtained by applying Procedure~\ref{alg:test1} with
$\alpha = 0.05, \beta = 0.05$ and $k= 2,3,4$ to the benchmark
datasets of Table~\ref{tab:datasets}.}\label{res:test3}
\end{table*}
We observe that for most pairs (dataset,$k$) the number of significant
frequent $k$-itemsets obtained is rather small, but, in fact, at
support $s^*$ in random instances of those datasets, less than two
(often much less than one) frequent $k$-itemsets would be expected.
These results provide evidence that our methodology not only defines
significance on statistically rigorous grounds, but also provides the
mining task with suitable support thresholds that avoid explosion of
the output size (the widely recognized ``Achilles' heel'' of
traditional frequent itemset mining). This feature crucially relies on
the identification of a region of ``rare events'' provided by the
Poisson approximation. As discussed in Section~\ref{sec:ourresults},
the discovery of significant itemsets with low support (not returned
by our method) would require the extraction of a large (possibly
exponential) number of itemsets, that would make any strategy aiming
to discover these itemsets unfeasible. Instead, we provide an efficient
method to identify, with high confidence level, the family of most
frequent itemsets that are statistically significant without
overwhelming the user with a huge number of discoveries.
There are, however, a few cases where the number of itemsets returned
is still considerably high. Their large number may serve as a sign
that the results call for further analysis, possibly using clustering
techniques \cite{XinHYC05} or limiting the search to \emph{closed
itemsets} \cite{PasquierBTL99}\footnote{An itemset is \emph{closed}
if it is not properly contained in another itemset with the same
support.}. For example, consider dataset Bms1 with $k=4$ and the
corresponding value $s^{*} = 5$ from Table~\ref{res:test3}. Extracting
the closed itemsets of support greater or equal to $s^{*}$ in that
dataset revealed the presence of a closed itemset of cardinality $154$
with support greater than $7$ in the dataset. This itemset, whose
occurrence by itself represents an extremely unlikely event in a
random dataset, accounts for more than 22M non-closed subsets with the
same support among the 27M reported as significant.
It is interesting to observe that the results obtained for dataset
Retail provide further evidence for the conclusions drawn in
\cite{GionisMMT06}, which suggested random behavior for this dataset
(although the random model in that work is slightly different from
ours, in that the family of random datasets also maintains the same
transaction lengths as the real one). Indeed, no support threshold
$s^*$ could be established for mining significant $k$-itemsets with $k
= 2,3$, while the support threshold $s^*$ identified for $k=4$ yielded
as few as 6 itemsets. However, the conclusion drawn in \cite{GionisMMT06}
was based on a qualitative assessment of the discrepancy between the
numbers of frequent itemsets in the random and real datasets, while
our methodology confirms the findings on a statistically sound
and rigorous basis.
Observe also that for some other pairs (dataset,$k$) our procedure
does not identify any support threshold useful for mining
statistically significant itemsets. This is an evidence that, for the
specific $k$ and for the high supports considered by our approach, these
datasets do not present a significant deviation from the corresponding
random datasets.
Finally, in order to assess its robustness, we applied our methodology
to random datasets. Specifically, for each benchmark dataset of
Table~\ref{tab:datasets} and for $k=2,3,4$, we generated 100 random
instances with the same parameters as those of the benchmark, and
applied Procedure~\ref{alg:test1} to each instance, searching for a
support threshold $s^*$ for mining significant itemsets. In
Table~\ref{tab:randomdatatest3} we report the number of times
Procedure~\ref{alg:test1} was successful in returning a finite value
for $s^*$. As expected, the procedure returned $s^*=\infty$, in
\emph{all cases} but for 2 of the 100 instances of the random dataset
with the same parameters as dataset Pumsb$^*$ with $k=2$. However, in
these two latter cases, mining at the identified support threshold only
yielded a very small number of significant itemsets (one and two,
respectively).
\begin{table}[h]
\begin{center}
\begin{tabular}{lccc}
\hline
& \multicolumn{3}{c}{$s^* < \infty$} \\
\cline{2-4}
Dataset & $k=2$ & $k=3$ & $k=4$ \\
\hline
RandomRetail & 0 & 0 & 0 \\
RandomKosarak & 0 & 0 & 0 \\
RandomBms1 & 0 & 0 & 0 \\
RandomBms2 & 0 & 0 & 0 \\
RandomBmspos & 0 & 0 & 0 \\
RandomPumsb$^*$ & 2 & 0 & 0 \\
\hline
\end{tabular}
\end{center}
\caption{Results for Procedure~\ref{alg:test1} with $\alpha = 0.05, \beta =
0.05$ for random versions of benchmark datasets; each entry
reports the number of times, out of 100 trials, the procedure
returned a finite value for $s^*$.}\label{tab:randomdatatest3}
\end{table}
\subsection{Relative effectiveness of Procedures~\ref{alg:fdrbino}
and \ref{alg:test1}} \label{sec:chernoffresults}
In order to assess the relative effectiveness of the two procedures
presented in the previous section, we applied them to the benchmark
datasets of Table~\ref{tab:datasets}. Specifically, we compared the number of
itemsets extracted using the threshold $s^*$ provided by
Procedure~\ref{alg:test1}, with the number of itemsets flagged as
significant using the more standard method based on Benjamini and
Yekutieli's technique (Procedure~\ref{alg:fdrbino}), imposing the same
upper bound $\beta=0.05$ on the FDR.
The results are displayed in Table~\ref{res:chernoffFDR}, where for
each pair (dataset,$k$), we report the cardinality of the family
${\cal R}$ of $k$-itemsets flagged as significant by
Procedure~\ref{alg:fdrbino}, and the ratio $r=Q_{k,s^*}/|{\cal R}|$,
where $Q_{k,s^*}$ is the number of $k$-itemsets of support at least
$s^*$, which are returned as significant with the methodology of
Subsection~\ref{sec:sstar}.
We observe that in all cases where Procedure~\ref{alg:test1} returned
a finite value of $s^*$ the ratio $r$ is greater than or equal to 1
(except for dataset Bms1 and $k=2$, and dataset Bmspos and $k=3$, where $r$
is however very close to 1). Moreover, in some cases the ratio $r$ is
rather large. Since both methodologies identify significant $k$-itemsets
among all those of support at least $s_{\min}$, these results provide
evidence that the methodology of Subsection~\ref{sec:sstar} is often more
(sometimes much more) effective. The methodology succeeds in identifying
more significant itemsets, since it evaluates the significance of the
\emph{entire} set ${\cal F}_{(k)}(s^*)$ by comparing $Q_{k,s^*}$ to
$\hat{Q}_{k,s^*}$. In contrast, Procedure~\ref{alg:fdrbino} must
implicitly test considerably more hypotheses (corresponding to the
significance all possible $k$-itemsets), thus the power of the test
(1-$Pr$(Type-II error)) is significantly smaller.
Observe that the cases where $r=0$ in Table~\ref{res:chernoffFDR}
correspond to pairs (dataset,$k$) for which Procedure~\ref{alg:test1}
returned $s^*=\infty$, that is, the procedure was not able to identify
a threshold for mining significant $k$-itemsets. Note, however, that
in all of these cases the number of significant $k$-itemsets returned
by Procedure~\ref{alg:fdrbino} is extremely small (between 1 and 3).
Hence, for these pairs, both methodologies indicate that there is very
little significant information to be mined at high supports.
\begin{table}[h]
\begin{center}
\begin{tabular}{lcccccccc}
\hline
& \multicolumn{2}{c}{$k=2$} & & \multicolumn{2}{c}{$k=3$} & &
\multicolumn{2}{c}{$k=4$} \\
\cline{2-3} \cline{5-6} \cline{8-9}
Dataset & $|{\cal R}|$ & $r$ & & $|{\cal R}|$ & $r$ & & $|{\cal R}|$ & $r$
\\
\hline
Retail & 3 & 0 & & 3 & 0 & & 6 & 1.0 \\
Kosarak & 1 & 0 & & 1 & 0 & & 12 & 1.0 \\
Bms1 & 60 & 0.933 & & 64367 & 4.441 & & 219706 & 122.9 \\
Bms2 & 429 & 1.0 & & 25906 & 1.394 & & 60927 & 11.72 \\
Bmspos & 2 & 0 & & 23 & 0.957 & & 891 & 1.0 \\
Pumsb$^*$ & 29 & 1.0 & & 406 & 1.0 & & 6288 & 1.001 \\
\hline
\end{tabular}
\end{center}
\caption{Results using Test~\ref{alg:fdrbino} to bound the FDR with $\beta =
0.05$ for itemsets of support $\geq s_{\min}$. }\label{res:chernoffFDR}
\end{table}
\section{Conclusions}\label{sec:conclusions}
The main technical contribution of this work is the proof that in a random
dataset where items are placed independently in transactions, there is
a minimum support $s_{\min}$ such that the number of $k$-itemsets with
support at least $s_{\min}$ is well approximated by a Poisson
distribution. The expectation of the Poisson distribution and the
threshold $s_{\min}$ are functions of the number of transactions,
number of items, and frequencies of individual items.
This result is at the base of a novel methodology for mining frequent
itemsets which can be flagged as statistically significant incurring a
small FDR. In particular, we use the Poisson distribution as the
distribution of the null hypothesis in a novel multi-hypothesis
statistical approach for identifying a suitable support threshold $s^*
\geq s_{\min}$ for the mining task. We control the FDR of the output
in a way which takes into account global characteristics of the
dataset, hence it turns out to be more powerful than other standard
statistical tools (e.g., \cite{BY01}). The results of a number of
experiments, reported in the paper, provide evidence of the
effectiveness of our approach.
To the best of our knowledge, our
methodology represents the first attempt at establishing a support
threshold for the classical frequent itemset mining problem with a
quantitative guarantee on the significance of the output.
\bibliographystyle{plain}
|
\section{Random Walks}
Over a century ago, the French mathematician Louis Bachelier proposed that stock prices follow a \emph{random walk}---that prices move up or down in random increments such that price changes are unpredictable\cite{Bachelier64}. When analyzing economic data, the random walk model is surprisingly accurate. It holds not only for stock prices, but also for the prices of many other items: stock indices, derivative instruments, commodities and other economic goods, and even for the prices of contracts traded on prediction markets. The regularity of this behavior across different items hints that some fundamental mechanism is behind it; perhaps some universal principle is at work.
In fact, most economists believe this is true, and they attribute the randomness of prices to the profit maximization (or loss aversion) of investors. If prices did not move randomly, but instead were in some way predictable, then this predictability would be quickly removed. After all, who would be willing to sell a stock for $\$90$ if everyone knew the price would move up to $\$100$ during the next period? Wouldn't sellers try to get something closer to $\$100$ right now, and wouldn't buyers be willing to pay something closer to $\$100$? When these individuals push the price to $\$100$, the predictability in the price movement is removed. If predictable price movements quickly disappear, then the only way for prices to move is with random increments. Paul Samuelson, an American economist, derived this result in his paper entitled ``Proof That Properly Anticipated Prices Fluctuate Randomly''\cite{Samuelson65}. It is an elegant and simple explanation for the universally observed random nature of price movements. The theory explains a large collection of phenomena and has predictive power (it predicts that any prices determined by profit maximizing agents will be random)---both are hallmarks of theories developed by seeking universal laws.
\begin{figure*}[htb]
\centering
\includegraphics[width=3.4in]{fig1a}
\includegraphics[width=3.4in]{fig1b}
\caption{(a) The probability distribution of daily returns for the S\&P 500 stock index from January 3, 1950 to November 25, 2009. Inset: Probability that the absolute daily return is above a threshold value, $x$. The five largest absolute returns are circled and their dates and sizes are shown. (b) The rescaled probability distribution of daily returns for six different traded items. Adapted from \cite{Gerig10}.}
\end{figure*}
\section{Extreme Price Movements}
There is another interesting regularity found in economic prices: very large price movements, such as stock market crashes, occur frequently\cite{Mandelbrot63, Fama63, Mandelbrot08}. This, again, happens across the board for many different economic items.
To understand just how large these price movements are, consider what it would mean if human heights behaved in a similar way. Assume for a moment that adult human heights varied between individuals in the same way that price movements vary. Within your circle of friends, there wouldn't be that much of a difference; most people would be between 5 to 6 ft tall. Outside of this circle, however, there would be dramatic changes. In your city, someone would be over 30 ft tall. In your country, the tallest person would likely reach 150 ft, and the tallest person in the world would be over 1000 ft tall.
The distinction between human heights and price movements is not just a pedagogical exercise, it is important because most financial models assume that the distribution of stock returns is the same as the distribution pattern for human heights -- the ubiquitous bell-shaped curve known as the normal (or Gaussian) distribution. If this were the case, very large returns (analogous to a 150 foot person) should never occur. But this is incorrect. For reasons we do not fully understand, stock returns are not distributed according to a normal distribution. Instead, they have a much larger peak and the `tails' or extremes of the distribution are thicker. This means that large price movements occur more often than predicted.
In Fig.~1(a), I show the distribution for the daily returns of the S\&P 500 stock index from January 3, 1950 to November 25, 2009. This plot can be replicated by downloading data from http://finance.yahoo.com. The horizontal axis measures the different sizes of returns (0.02 is a $2\%$ return, 0.04 is a $4\%$ return, etc.) and the vertical axis shows the relative likelihood of these price changes -- the higher the red bar, the more likely that event is observed. Small returns, close to zero, are the most likely occurrence. A normal distribution is fit to the data and is drawn with a black line. Notice that this does not coincide well with the S\&P 500 data.
The inset plot shows the probability that a daily return is above a certain threshold value. It enlarges the tail of the distribution---the area where large price movements are recorded. You can see that the probability of large returns is much higher than what the normal distribution predicts, i.e., the red curve is above the black curve for large values of $x$. I've circled the five highest returns and show their values and the dates they were observed. Not surprisingly, the largest return occurred on Black Monday, October 19, 1987, when stock markets crashed around the world.
If you look at the y-axis in the inset plot, the probability for a daily return to exceed $10\%$ is around $10^{-4}$, which means this has occurred approximately once out of $1/10^{-4}=10,000$ trading days, or once every 40 years. For comparison, the black curve---a normal distribution---predicts this to occur once every $7\times10^{18}$ years, which is longer than the age of the universe and for all practical purposes, means never. Obviously this is incorrect.
One way to explain the discrepancy between observed stock returns and financial models is to consider large price movements as outliers---surprising events outside of the normal model. There are several reasons to do this. First, there are good underlying reasons to assume a normal distribution for returns as a first guess, and there is no accepted theory for why it should be otherwise. Second, we usually explain large price movements in this way---stock markets crashed because computer trading malfunctioned or the global financial crises occurred because banks made large mistakes. When using these explanations, we implicitly suggest that they are one-time events---outliers---that can be accounted for and controlled in the future. The problem is, despite our efforts, they keep happening.
An alternative explanation is that something more fundamental is producing these events and that the widely reported and agreed-upon culprits are just symptoms of the same underlying cause. There are several reasons to believe this is true. First, extreme price movements are not just observed for stocks; these events occur universally across traded items. Second, the empirical evidence does not show these events as statistical outliers. You can see this for the S\&P 500 index in the inset plot where the red curve extends continuously in a smooth way down to the points where extreme price movements are recorded. These points do not exist by themselves but nicely fit where you'd expect them when extrapolating the red curve from smaller price movements. Finally, there is evidence that the probability distribution of price returns is universal, that it deviates from the normal distribution in the exact same way for different items and over different time periods\cite{Fuentes09,Gerig10}. In Fig.~1(b), I show the probability distribution for daily returns for five different traded items. By appropriately rescaling the axes for each, the distributions collapse on the same non-normal curve. Why would these unrelated price series behave in the same way unless something fundamental was the cause?
Despite the idiosyncratic behavior of individuals, regularities exist in social and economic systems that are similar to those found in natural systems. Specific examples are found in the way that economic prices behave. The reason prices are random is well understood; it occurs because individuals are profit maximizing. The reason prices deviate from a normal distribution is not understood and is currently a matter of much debate. I believe the evidence suggests some universal mechanism underlies these deviations, and that large price movements are not outliers to an otherwise correct (normal) model. If true, understanding this mechanism is extremely important. If large price movements result from human behavior or the way in which markets are structured, then there might be ways to curtail behavior or structure markets differently such that these extreme events do not occur. If they are due to some economic cause, then perhaps it is something we can only understand and better prepare for. At least then we would have correct models on which to base economic decisions.
|
\chapter*{Abstract}
\addcontentsline{toc}{chapter}{Abstract}
Given an algebraic theory which can be described by a (possibly symmetric)
operad $P$, we propose a definition of the \emph{weakening} (or
\emph{categorification}) of the theory, in which equations that hold strictly
for $P$-algebras hold only up to coherent isomorphism.
This generalizes the theories of monoidal categories and symmetric monoidal
categories, and several related notions defined in the literature.
Using this definition, we generalize the result that every monoidal category is
monoidally equivalent to a strict monoidal category, and show that the
``strictification'' functor has an interesting universal property, being
left adjoint to the forgetful functor from the category of strict
$P$-categories to the category of weak $P$-categories.
We further show that the categorification obtained is independent of our choice
of presentation for $P$, and extend some of our results to many-sorted
theories, using multicategories.
\chapter{Categorification}
\label{ch:categorification}
\section{Desiderata}
Many categorifications of individual theories have been proposed in the
literature.
We aim to replace these with a general definition, which should satisfy the
following criteria insofar as possible:
\begin{itemize}
\item \textbf{Broad:} it should cover as large a class of theories as
possible.
\item \textbf{Consistent with earlier work:} where a categorification of a
given theory is known, ours should agree with this categorification or be
demonstrably better in some way.
\item \textbf{Canonical:} it should be free of arbitrary tunable parameters
(and if possible should be given by some universal property).
\end{itemize}
We shall return to these criteria in Section \ref{sec:evaluation} and evaluate
how close we have come to achieving them.
Our strategy is as follows: we start with a na\"ive version of categorification
for strongly regular theories, which closely parallels Mac Lane and Benabou's
categorification of the theory of monoids.
This will be an \emph{unbiased} categorification, which treats all operations
equally, without regarding any as ``primitive'': for instance, if
$P$ is the terminal operad (whose strict algebras are monoids), then the weak
$P$-categories will have tensor products of all arities, not just 0 and 2.
We then re-express our definition of categorification in terms of factorization
systems, which allows us to generalize our definition in two directions
simultaneously: to symmetric operads, and to operads with presentations.
We then use this new definition to recover the classical theory of symmetric
monoidal categories (at which several other proposed general definitions of
categorification fail), and investigate what it yields in the case of some
other linear theories.
\section{Categorification of strongly regular theories}
\label{sec:naivecatn}
The idea is to consider the strict models of our theory as algebras for an
operad, then to obtain the weak models as (strict) algebras for a weakened
version of that operad (which will be a \cat{Cat}-operad).
We weaken the operad using a similar approach to that used in Penon's
definition of $n$-category, as described in \cite{penon}.
A non-rigorous summary of Penon's construction can be found in
\cite{cheng+lauda}.
Throughout this section, let $P$ be a plain (\cat{Set}-)operad.
Let $D_*: \cat{Operad} \to \cat{Cat-Operad}$ be the functor which
takes discrete categories levelwise; i.e., $(D_*P)_n$ is the discrete category
on the set $P_n$.
In terms of the ``$n$-cell'' terminology introduced in Chapter \ref{ch:factsys},
the 1-cells of $(D_* P)_n$ are $n$-ary\ arrows in $P$, and the only 2-cells are
identities.
\begin{defn}
\label{def:weakening}
\index{unbiased!weakening of a plain operad}
\index{weakening!of a plain operad}
The \defterm{unbiased weakening of $P$}, \Wk{P}, is the following \cat{Cat}-operad:
\begin{itemize}
\item \emph{1-cells:} 1-cells of $D_* {\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} P$;
\item \emph{2-cells:} if $A, B \in ({\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} P)_n$, there is a single 2-cell $A
\to B$ if $\epsilon(A) = \epsilon(B)$ (where $\epsilon$ is the counit
of the adjunction ${\ensuremath{F_{\rm pl}}} \dashv {\ensuremath{U^{\rm pl}}}$), and no 2-cells $A \to B$ otherwise;
\item \emph{Composition of 2-cells:} the composite of two arrows $A \rightarrow
B \rightarrow C$ is the unique arrow $A \rightarrow C$, and in particular, the
arrows $A \rightarrow B$ and $B \rightarrow A$ are inverses;
\item \emph{Operadic composition:} on 1-cells, as in ${\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} P$, and on 2-cells,
determined by the uniqueness property.
\end{itemize}
\end{defn}
See Fig. \ref{Wk(1)pic}, which illustrates a fragment of the unbiased weakening
of the terminal operad 1.
Since $1_n$ is a singleton set for every $n \in \natural$, then $\Wk 1_n$ is
the indiscrete category whose objects are unlabelled $n$-ary\ strongly regular
trees for all $n \in \natural$.
We may embed the discrete category on each $P_n$ in $\Wk{P}_n$, via the map $p
\mapsto p \kel (|, \dots, |)$.
We shall occasionally abuse notation and consider some $p \in P_n$ as a 1-cell
of $\Wk{P}_n$.
\begin{figure}[h]
\centerline{
\epsfxsize=3in
\epsfbox{cat-operad.eps}
}
\caption{Part of ${\Wk 1}_3$}
\label{Wk(1)pic}
\end{figure}
\begin{theorem}
\label{thm:WkP_characterization}
$\Wk{P}$ is the unique \cat{Cat}-operad with the following properties:
\begin{itemize}
\item $\Wk{P}$ has the same 1-cells as $D_* {\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} P$;
\item we may extend the counit $\epsilon_P : {\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} P \to P$ to a map of
\cat{Cat}-operads $\Wk{P} \to D_* P$, which is full and faithful levelwise.
\end{itemize}
\end{theorem}
\begin{proof}
Immediate.
\end{proof}
We may now make the following definition:
\begin{defn}
\label{def:wkpcat}
\index{weak!$P$-category}
A \defterm{weak $P$-category} is an algebra for \Wk{P}.
\end{defn}
In the case $P = 1$, this reduces exactly to Leinster's definition of unbiased
monoidal category in \cite{hohc} section 3.1.
There, two 1-cells $\phi$ and
$\psi$ have the same image under $\epsilon$ iff they have the same arity, so
the categories $\Wk{1}_i$ are indiscrete.
If $h: \Wk{P} \kel A \to A$ is a weak $P$-category, we refer to the image under
$h$
of a 2-cell $q \to q'$ in $\Wk{P}$ as $\delta_{q,q '}$.
This is clearly a natural transformation $h(q,-) \to h(q',-)$.
As a special case, we write $\delta_q$ for $\delta_{q, \epsilon(q)}$ (where we consider $\epsilon(q)$ as a 1-cell of $\Wk{P}$ as described above).
\begin{defn}
\index{strict!$P$-category}
A \defterm{strict $P$-category} is an algebra for $D_* P$.
\end{defn}
Equivalently, a strict $P$-category is a weak $P$-category in which every
component of $\delta$ is an identity arrow.
\begin{defn}
\index{weak!$P$-functor}
\index{strict!$P$-functor}
Let $(A, h)$ and $(B, h')$ be weak $P$-categories. A \defterm{weak
$P$-functor} from $(A, h)$ to $(B, h')$ is a weak map of $\Wk{P}$-algebras.
A \defterm{strict $P$-functor} from $(A, h)$ to $(B, h')$ is a strict map of
$\Wk{P}$-algebras.
\end{defn}
Equivalently, a strict $P$-functor is a weak $P$-functor for which all the
coherence maps are identities.
These definition are natural generalizations of the definition of weak and
strict unbiased monoidal functors given in \cite{hohc} section 3.1.
\begin{defn}
\index{transformation!between $P$-functors}
Let $(F, \phi)$ and $(G, \psi)$ be weak $P$-functors $(A, h) \to (B, h')$. A
\defterm{$P$-transformation} $\sigma: (F, \phi) \to (G, \psi)$ is a
$\Wk{P}$-transformation $(F, \phi) \to (G, \psi)$, in the sense of Definition
\ref{transfdef}.
\end{defn}
Note that there is only one possible level of strictness here.
There is a 2-category, \cat{Wk-$P$-Cat}, whose objects are weak $P$-categories, whose
1-cells are weak $P$-functors, and whose 2-cells are $P$-transformations.
Similarly, there is a 2-category \cat{Str-$P$-Cat}\ of strict $P$-categories, strict
$P$-functors, and $P$-transformations, which can be considered a sub-2-category
of \cat{Wk-$P$-Cat}.
\begin{defn}
A \defterm{$P$-equivalence} is an equivalence in the 2-category \cat{Wk-$P$-Cat}.
\end{defn}
\begin{lemma}
\label{inv <=> P-inv}
Let $P$ be a plain operad, $(A, h)$ and $(B, h')$ be weak $P$-categories, and
$(F, \phi), (G, \psi) : (A, h) \to (B, h')$ be weak $P$-functors.
A $P$-transformation $\sigma: (F, \phi) \to (G, \psi)$ is invertible as a
$P$-transformation if and only if it is invertible as a natural
transformation.
\end{lemma}
\begin{proof}
This is a straightforward application of Lemma \ref{inv <=> Q-inv}.
\end{proof}
\section{Examples}
Unfortunately, few well-studied theories are strongly regular.
We will consider the following examples:
\begin{enumerate}
\item the trivial theory (in other words, the theory of sets);
\item the theory of pointed sets;
\item the theory of monoids;
\item the theory of $M$-sets, for a monoid $M$.
\end{enumerate}
While we could easily invent a new strongly regular theory to categorify, this
would not help us to see how well our definition of weakening accords with our
intuitions.
Further examples will be considered later, when the machinery to categorify
theories-with-generators and linear theories has been developed.
We will first need to introduce an auxiliary definition:
\begin{defn}
\index{trivial monad}
Let \ensuremath{\mathcal{C}}\ be a category, and $(T, \mu, \eta)$ be a monad on $\ensuremath{\mathcal{C}}$.
We say that $(T, \mu, \eta)$ is \defterm{trivial} if $\eta$ is a natural
isomorphism.
\end{defn}
\begin{lemma}
\label{lem:triv<=>initial}
The identity monad on $\ensuremath{\mathcal{C}}$ is initial in the category $\cat{Mnd}(\ensuremath{\mathcal{C}})$ of monads on
\ensuremath{\mathcal{C}}, with the unique morphism of monads $(1_\ensuremath{\mathcal{C}}, 1, 1) \to (T, \mu, \eta)$ being
$\eta$.
\end{lemma}
\begin{proof}
First we show that $\eta$ is a morphism of monads in the sense of Street
(Definition \ref{def:monad_morphism}).
One axiom corresponds to the outside of the diagram
\[
\xymatrix{
1 \ar[r]^\eta \ar[d]_1 & T \ar[r]^{\eta T} \ar[dr]^1 \ar[d]_1 & T^2 \ar[d]^\mu\\
1 \ar[r]_\eta & T \ar[r]_1 & T
}
\]
commuting; all the inner segments commute (the top right triangle by the unit
axiom for monads), so the outside must commute.
The other axiom corresponds to the diagram
\[
\xymatrix{
1 \ar[r]^1 \ar[dr]_\eta & 1 \ar[d]^\eta \\
& T
}
\]
and this commutes trivially.
Hence, $\eta$ is a morphism of monads $1 \to T$.
Now suppose that $\alpha : 1 \to T$ is a morphism of monads.
From the unit axiom for monad morphisms, the diagram
\[
\xymatrix{
1 \ar[r]^1 \ar[dr]_\eta & 1 \ar[d]^\alpha \\
& T
}
\]
must commute, so $\eta = \alpha$.
\end{proof}
\begin{corollary}
\label{cor:triv<=>initial}
A monad $(T, \mu, \eta)$ on $\ensuremath{\mathcal{C}}$ is trivial if and only if it is isomorphic to
the identity monad on $\ensuremath{\mathcal{C}}$.
\end{corollary}
\begin{proof}
If $T$ is isomorphic to the identity monad, then by Lemma
\ref{lem:triv<=>initial} the isomorphism concerned must be $\eta$, so $\eta$
must be invertible.
It is readily checked that if $\eta$ is invertible, then $\eta^{-1}$ must be a
morphism of monads, so if $T$ is trivial then it is isomorphic to the identity
monad.
\end{proof}
\begin{example} \emph{The trivial theory:}
\label{ex:trivial_theory}
\index{sets}
Let $0$ be the initial operad, whose algebras are sets.
An unbiased weak $0$-category is a category equipped with a specified trivial
monad, for the following reason.
$0$ has only one operator (call it $I$), of arity one. Hence $({\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} 0)_1
\cong \natural$, and all other $({\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} 0)_n$'s are empty.
All derived operations in the theory of sets are composites of identities, and
thus equivalent to the identity.
So all objects of $\Wk{0}_1$ are isomorphic.
Hence, $I \cong \mbox{id}$.
All diagrams commute: in particular, those giving the monad and monad morphism
axioms commute, so in any weak $0$-category $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}}))$, the functor
$\hat I$ is a monad, and the isomorphism $\hat I \to 1_\ensuremath{\mathcal{C}}$ is an isomorphism of
monads.
By Corollary \ref{cor:triv<=>initial}, $\hat I$ must be trivial.
Conversely, suppose $T$ is a trivial monad on a category $\ensuremath{\mathcal{C}}$.
We wish to show that $\mu$ is also invertible, and thus that $\ensuremath{\mathcal{C}}$ is an
unbiased weak $0$-category.
From the monad axioms, we have that
\[
\xymatrix{
T \ar[r]^{\eta T} \ar[dr]_{\mbox{id}}
& T^2 \ar[d]^{\mu} \\
& T
}
\]
commutes.
But $\eta T$ is invertible, so $\mu$ must be its inverse.
So $\mbox{id} \cong T \cong T^2 \cong T^3 \cong \ldots$, and all diagrams
commute.
Hence $\ensuremath{\mathcal{C}}$ is an unbiased weak $0$-category.
If $(\ensuremath{\mathcal{C}}, S)$ and $(\ensuremath{\mathcal{D}}, T)$ are weak $0$-categories, then a weak $0$-functor
$(\ensuremath{\mathcal{C}}, S) \to (\ensuremath{\mathcal{D}}, T)$ is a functor $F: \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{D}}$ and a natural isomorphism
$\phi : TF \to FS$, such that the equations
\[
\xymatrix{
& \ensuremath{\mathcal{C}} \ar[dl]_S \ar[dd]^S \ar[rr]^F \ddrrtwocell\omit{^\phi^{-1}}
& & \ensuremath{\mathcal{D}} \ar[dd]^T \\
\ensuremath{\mathcal{C}} \ar[dr]_S \\
& \ensuremath{\mathcal{C}} \ar[rr]_F \uuuppertwocell\omit{<-3>\mu} & & D
}
\xymatrix{{} \\ = \\ {}}
\xymatrix{
& \ensuremath{\mathcal{C}} \ar[dl]_S \ar[rr]^F \drtwocell\omit{^\phi^{-1}}
& & \ensuremath{\mathcal{D}} \ar[dd]^T \ar[dl]_T \\
\ensuremath{\mathcal{C}} \ar[dr]_S \ar[rr]^F
& {} \drtwocell\omit{^\phi^{-1}} & D \ar[dr]_T \\
& \ensuremath{\mathcal{C}} \ar[rr]_F & & D \uuuppertwocell\omit{<-3>\mu}
}
\]
and
\[
\xymatrix{
\ensuremath{\mathcal{C}} \ddlowertwocell<-10>_{1_\ensuremath{\mathcal{C}}}{^\eta} \ar[dd]^S \ar[rr]^F
\ddrrtwocell\omit{^\phi^{-1}}
& & \ensuremath{\mathcal{D}} \ar[dd]^T \\ \\
\ensuremath{\mathcal{C}} \ar[rr]_F & & \ensuremath{\mathcal{D}}
}
\xymatrix{{} \\ = \\ {}}
\xymatrix{
\ensuremath{\mathcal{C}} \ddrrtwocell\omit{^\phi^{-1}} \ar[rr]^F \ar[dd]_{1_\ensuremath{\mathcal{C}}}
& & \ensuremath{\mathcal{D}} \dduppertwocell<10>^T{^\eta} \ar[dd]_{1_\ensuremath{\mathcal{D}}} \\ \\
\ensuremath{\mathcal{C}} \ar[rr]_F & & \ensuremath{\mathcal{D}}
}
\]
are satisfied.
\end{example}
\begin{example} \emph{Pointed sets:}
\label{ex:pointed_sets}
\index{pointed sets}
Let $P$ be the operad with a single element of arity 0 (call it $*$) and a
single element of arity 1 (the identity).
Strict algebras for $P$ in $\cat{Set}$ are pointed sets.
The set $({\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} P)_0$ is countable (it has elements $*, I*, I^2*, I^3*,
\ldots$, and so is $({\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} P)_1$ (it has elements $\mbox{id}, I, I^2,
\ldots$).
So an unbiased weak $P$-category is a category $\ensuremath{\mathcal{C}}$ equipped with a
distinguished object $\hat *$ and a trivial monad $\hat I$.
If $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}}))$ and $(\ensuremath{\mathcal{D}}, (\bar{\phantom{\alpha}})$ are unbiased weak $P$-categories,
then a weak $P$-functor $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}})) \to (\ensuremath{\mathcal{D}}, (\bar{\phantom{\alpha}}))$ is a triple $(F,
\phi, \psi)$, where $F$ and $\phi$ are as in Example \ref{ex:trivial_theory},
$\psi : \bar * \to F \hat *$ is an isomorphism, and there is exactly one
natural isomorphism $\bar I^n \bar * \to F \hat I^m \hat *$ composed from
$\phi$s and $\psi$s for each $m$ and each $n \in \natural$.
\end{example}
\begin{example} \emph{Monoids:}
\label{ex:monoids}
\index{monoids}
An unbiased weak $1$-category is precisely an unbiased weak monoidal
category in the sense of Definition \ref{def:umoncat}.
An unbiased weak $1$-functor is an unbiased weak monoidal functor.
For a proof, see \cite{hohc} Theorem 3.2.2.
\end{example}
\begin{example} \emph{$M$-sets:}
\label{ex:M-sets}
\index{$M$-sets}
Let $M$ be a monoid, and $N$ be the operad such that
\begin{eqnarray*}
N_1 & = & M \\
N_i & = & \emptyset \mbox{ whenever $i \neq 1$}
\end{eqnarray*}
with composition of arrows of arity 1 given by the multiplication in $M$.
An algebra for $N$ in \cat{Set}\ is an $M$-set.
An unbiased weak $N$-category is a category $\ensuremath{\mathcal{C}}$ with a functor $\hat m :
\ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{C}}$ for each $m \in \ensuremath{\mathcal{M}}$.
For every equation $m_1 m_2 \dots m_i = n_1 n_2 \dots n_j$ that is true in $M$,
there is a natural isomorphism $\delta_{m_1 \dots m_i}^{n_1 \dots n_j} : \hat
m_1 \hat m_2 \dots \hat m_i \to \hat n_1 \hat n_2 \dots \hat n_j$.
If $e$ is the identity element in $M$, then $\hat e$ is a trivial monad.
All diagrams involving these natural isomorphisms commute.
Hence, an unbiased weak $N$-category is a category $\ensuremath{\mathcal{C}}$ together with a weak
monoidal functor $M \to \mathop{\rm{End}}(\ensuremath{\mathcal{C}}$).
If $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}}))$ and $(\ensuremath{\mathcal{D}}, (\bar{\phantom{\alpha}}))$ are unbiased weak $N$-categories, an
unbiased weak $N$-functor is a functor $F : \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{D}}$ together with natural
transformations $\phi_m : \bar m F \to F \hat m$ for all $m \in M$, such that
if $m_1 m_2 \dots m_i = n_1 n_2 \dots n_j$ in $M$, there is precisely one
natural isomorphism $\bar m_1 \dots \bar m_i F \to F \hat n_1 \dots \hat n_j$
that can be formed by composing $\delta$s and $\phi$s.
\end{example}
\section{A more general approach: factorization systems}
Recall from Definition \ref{def:opd_presentation} the definition of a
presentation and a generator for an operad.
We will define a categorification of any symmetric operad equipped with a
generator, generalizing the unbiased categorification defined in Section
\ref{sec:naivecatn}.
In particular, we shall consider categorification with respect to the component
of the counit $\epsilon_P: {\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} P \to P$ at a symmetric operad $P$; this
is a generator for $P$ since both
\[
\fork{{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} {\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} P}{\epsilon {\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} }{{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} \epsilon}{{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} P} {\epsilon}{P}
\]
and
\[
\label{item:regular_pres}
\fork{{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} P\times_P {\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} P}{\pi_1}{\pi_2} {{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} P}{\epsilon}{P}\]
are coequalizer diagrams (the latter by Lemma \ref{lem:coeq-kerpair}).
We will then show that the categorification is independent of our choice of
generator, in the sense that the symmetric \cat{Cat}-operads which arise are
equivalent (and thus have equivalent categories of algebras).
\begin{defn}
\index{weakening!of a symmetric operad}
\label{def:weakening_pres}
Let $\Phi$ be a signature, $P$ be a symmetric operad, and $\phi: {\ensuremath{F_{\Sigma}}}\Phi \to
P$ be a regular epi in \cat{$\Sigma$-Operad}.
Then the \defterm{weakening} (or \defterm{categorification}) $\Wkwrt P \phi$ of
$P$ with respect to $\phi$ is the (unique-up-to-isomorphism) symmetric
\cat{Cat}-operad such that the following diagram commutes:
\[
\xymatrix{
D_*{\ensuremath{F_{\Sigma}}}\Phi \ar[rr]^{D_* \phi} \booar[dr]_b
&& D_* P \\
& \Wkwrt P \phi \lffar[ur]_f
}
\]
where $f$ is full and faithful levelwise, $b$ is levelwise bijective on
objects, and $D_*$ is the levelwise discrete category functor $\cat{$\Sigma$-Operad} \to
\cat{Cat-$\Sigma$-Operad}$.
The existence and uniqueness of $\Wkwrt P \phi$ follow from Lemma
\ref{lem:fact_standard} applied to the factorization system on \cat{Cat-$\Sigma$-Operad}\
described in \ref{ex:catopdFS} above.
\end{defn}
\begin{defn}
\label{def:wkPphicat}
\index{weak!$P$-category}
Let $\phi, \Phi$ and $P$ be as above.
A \defterm{$\phi$-weak $P$-category} is an algebra for $\Wkwrt P \phi$.
\end{defn}
Note that any strict algebra for $P$ can be considered as a $\phi$-weak
$P$-category (for any $\phi$), via the map $\xymatrix{\Wkwrt P \phi \lffar[r] &
D_* P}$.
\begin{defn}
\index{weak!$P$-functor}
Let $\phi, \Phi$ and $P$ be as above.
A \defterm{$\phi$-weak $P$-functor} is a weak map of $\Wkwrt P \phi$-algebras.
\end{defn}
\begin{defn}
\index{weakening!of a symmetric operad}
Let $P$ be a symmetric operad, and $\parallelpair {{\ensuremath{F_{\Sigma}}} E}{e_1}{e_2}{{\ensuremath{F_{\Sigma}}}
\Phi}$ be a presentation for $P$, with $\phi: {\ensuremath{F_{\Sigma}}} \Phi \to P$ being the
regular epi in Definition \ref{def:opd_presentation}.
The \defterm{weakening of $P$ with respect to $(\Phi, E)$} is the weakening of
$P$ with respect to $\phi$.
\end{defn}
\begin{defn}
\label{def:wkp}
\index{unbiased!weakening of a symmetric operad}
The \defterm{unbiased} weakening of $P$ is the weakening arising from
the counit $\epsilon: {\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} P \to P$ of the adjunction ${\ensuremath{F_{\Sigma}}} \dashv
{\ensuremath{U^{\Sigma}}}$.
Call this symmetric \cat{Cat}-operad \Wk{P}.
\end{defn}
\begin{lemma}
\label{lem:explicit_wkp}
Let $\phi, \Phi$ and $P$ be as above.
Then, for every $n \in \natural$, the category $\Wkwrt P \phi_n$ is the
equivalence relation $\sim$ on the elements of $({\ensuremath{F_{\Sigma}}} \Phi)_n$, where $t_1 \sim t_2$ if $\phi(t_1) = \phi(t_2)$.
\end{lemma}
\begin{proof}
Let $n \in \natural$, and $t_1, t_2 \in \Wkwrt P \phi_n$.
The objects of $\Wkwrt P \phi_n$ are the elements of $({\ensuremath{F_{\Sigma}}} \Phi)_n$, by
construction.
Since $\phi_n$ factors through a full functor $\xymatrix{ \Wkwrt P
\phi_n \lffar[r] & (D_*P)_n }$ and $(D_*P)_n$ is the discrete category on
$P_n$, there is an arrow $t_1 \to t_2$ in $\Wkwrt P \phi_n$ iff $\phi(t_1) =
\phi(t_2)$.
Since this functor is also faithful, such an arrow must be unique.
Hence $\Wkwrt P \phi_n$ is a poset; it is readily checked that it is also an
equivalence relation.
\end{proof}
An obvious question is how this notion of weakening is related to the version
defined for plain operads in Section \ref{sec:naivecatn}.
In light of Theorem \ref{thm:WkP_characterization}, it is clear that the
plain-operadic version can be re-phrased as in Definition \ref{def:wkp} above,
but with the factorization occurring in \cat{Cat-Operad}\ rather than \cat{Cat-$\Sigma$-Operad}.
We may generalize it to give a definition of the weakening of a plain operad
$P$ with respect to a generator $\phi$:
\begin{defn}
\label{def:phi-wkpcat_plain}
Let $P$ be a plain operad, $\Phi$ be a signature, and $\phi: {\ensuremath{F_{\rm pl}}} \Phi \to P$ be
a regular epi.
The \defterm{weakening $\Wkwrt P \phi$ of $P$ with respect to $\phi$} is the
plain $\cat{Cat}$-operad given by the bijective on objects/levelwise full and
faithful factorization
\[
\xymatrix{
D_* {\ensuremath{F_{\rm pl}}} \Phi \ar[rr]^{D_* \phi} \booar[dr] & & D_* P \\
& \Wkwrt P \phi \lffar[ur]
}
\]
in \cat{Cat-Operad}.
A \defterm{$\phi$-weak $P$-category} is an algebra for $\Wkwrt P \phi$.
\end{defn}
But do the weak algebras for a strongly regular theory $T$ change if we
consider $T$ as a linear theory instead?
We now answer that question in the negative.
\begin{theorem}
\label{thm:plain_as_symmetric}
\index{weakening!of a plain operad!considered as a symmetric operad}
Let $P$ be a plain operad, let $\Phi$ be a signature, and let $\phi : {\ensuremath{F_{\rm pl}}} \Phi
\to P$ be a regular epi.
Then $\Wkwrt{{\ensuremath{F_{\Sigma}^{\rm pl}}} P}{{\ensuremath{F_{\Sigma}^{\rm pl}}} \phi} \cong {\ensuremath{F_{\Sigma}^{\rm pl}}} (\Wkwrt{P}{\phi})$ in the
category $\cat{Cat-$\Sigma$-Operad}$.
\end{theorem}
\begin{proof}
First note that $\Wkwrt{{\ensuremath{F_{\Sigma}^{\rm pl}}} P}{{\ensuremath{F_{\Sigma}^{\rm pl}}} \phi}$ is well-defined: ${\ensuremath{F_{\Sigma}^{\rm pl}}}$ is a
left adjoint, and hence preserves colimits, so ${\ensuremath{F_{\Sigma}^{\rm pl}}} \phi$ is a regular epi
in $\cat{$\Sigma$-Operad}$.
$\Wkwrt{{\ensuremath{F_{\Sigma}^{\rm pl}}} P}{{\ensuremath{F_{\Sigma}^{\rm pl}}} \phi}$ is defined by its universal property, so it is
enough to show that the \cat{Cat}-operad ${\ensuremath{F_{\Sigma}^{\rm pl}}} (\Wkwrt P \phi)$ also has this
property.
Specifically, it is enough to show that if
\[
\xymatrix{
D_* {\ensuremath{F_{\rm pl}}} \Phi \ar[rr]^{D_* \phi} \booar[dr]_{b} && D_* P \\
& \Wkwrt P \phi \lffar[ur]_f
}
\]
is the bijective-on-objects/full-and-faithful factorization of $\phi$, then in
the diagram
\[
\xymatrix{
D_* {\ensuremath{F_{\Sigma}}} \Phi \ar[rr]^{D_* {\ensuremath{F_{\Sigma}^{\rm pl}}} \phi} \booar[dr]_{{\ensuremath{F_{\Sigma}^{\rm pl}}} b}
&& D_* {\ensuremath{F_{\Sigma}^{\rm pl}}} P \\
& {\ensuremath{F_{\Sigma}^{\rm pl}}} (\Wkwrt P \phi) \lffar[ur]_{{\ensuremath{F_{\Sigma}^{\rm pl}}} f}
}
\]
the arrow ${\ensuremath{F_{\Sigma}^{\rm pl}}} b$ is bijective on objects and the arrow ${\ensuremath{F_{\Sigma}^{\rm pl}}} f$ is
levelwise full and faithful (note that $D_* {\ensuremath{F_{\Sigma}^{\rm pl}}} = {\ensuremath{F_{\Sigma}^{\rm pl}}} D_*$).
But this follows straightforwardly from the explicit construction of ${\ensuremath{F_{\Sigma}^{\rm pl}}}$
in Section \ref{sec:explicit_Fs}.
\end{proof}
\begin{corollary}
Let $P$ be a plain operad, $\phi : {\ensuremath{F_{\rm pl}}} \Phi \to P$ generate $P$, and $A$ be
a $\phi$-weak $P$-category in the sense of Definition
\ref{def:phi-wkpcat_plain}.
Then $A$ is an ${\ensuremath{F_{\Sigma}^{\rm pl}}} \phi$-weak ${\ensuremath{F_{\Sigma}^{\rm pl}}} P$-category in the sense of
Definition \ref{def:wkPphicat}.
Conversely, every ${\ensuremath{F_{\Sigma}^{\rm pl}}} \phi$-weak ${\ensuremath{F_{\Sigma}^{\rm pl}}} P$-category is a weak
$P$-category.
\end{corollary}
\begin{proof}
A ${\ensuremath{F_{\Sigma}^{\rm pl}}} \phi$-weak ${\ensuremath{F_{\Sigma}^{\rm pl}}} P$-category is a category $A$ and a
morphism $\Wkwrt{{\ensuremath{F_{\Sigma}^{\rm pl}}} P}{{\ensuremath{F_{\Sigma}^{\rm pl}}} \phi} \to \mathop{\rm{End}}(A)$ of symmetric \cat{Cat}-operads.
By Theorem \ref{thm:plain_as_symmetric}, this is equivalent to a morphism
${\ensuremath{F_{\Sigma}^{\rm pl}}} (\Wkwrt P \phi) \to \mathop{\rm{End}}(A)$ in $\cat{Cat-$\Sigma$-Operad}$, which is equivalent
by the adjunction ${\ensuremath{F_{\Sigma}^{\rm pl}}} \dashv {\ensuremath{U^{\Sigma}_{\rm pl}}}$ to a morphism of plain
\cat{Cat}-operads $\Wkwrt P \phi \to {\ensuremath{U^{\Sigma}_{\rm pl}}} \mathop{\rm{End}}(A)$.
This is exactly a $\phi$-weak $P$-category.
\end{proof}
Note that we had to apply ${\ensuremath{F_{\Sigma}^{\rm pl}}}$ to $\phi$ to obtain a generator for ${\ensuremath{F_{\Sigma}^{\rm pl}}}
P$.
This means that the theorem does not tell us that the unbiased categorification
is unaffected by whether we consider our theory as a linear or a strongly
regular one.
In fact, it is not the case that $\Wk{{\ensuremath{F_{\Sigma}^{\rm pl}}} P} \cong {\ensuremath{F_{\Sigma}^{\rm pl}}} (\Wk{P})$ in
general.
\begin{example}
\index{monoids}
\index{\ensuremath{\mathcal{S}}!as ${\ensuremath{F_{\Sigma}^{\rm pl}}} 1$}
Consider the terminal plain operad $1$ whose algebras are monoids.
${\ensuremath{F_{\Sigma}^{\rm pl}}} 1$ is the operad $\ensuremath{\mathcal{S}}$ of Example \ref{ex:symmopd}, for which each
$\ensuremath{\mathcal{S}}_n$ is the symmetric group $S_n$.
Then the objects of $\Wk{\ensuremath{\mathcal{S}}}_n$ are $n$-leafed permuted trees with each node
labelled by a permutation, whereas the 1-cells of $({\ensuremath{F_{\Sigma}^{\rm pl}}} \Wk{1})_n$ are
unlabelled permuted trees.
These two sets are not canonically isomorphic.
Hence, there is no canonical isomorphism between $\Wk{\ensuremath{\mathcal{S}}}$ and ${\ensuremath{F_{\Sigma}^{\rm pl}}}
\Wk{1}$.
\end{example}
However, we can make a weaker statement: the two candidate unbiased weakenings
are \emph{equivalent} in the 2-category $\cat{Cat-$\Sigma$-Operad}$.
We shall return to this point in Corollary \ref{cor:unbiased_symm}.
\section{Examples}
\begin{example}
\label{ex:trivial_theory_biased}
\index{sets}
Consider the trivial theory (given by the initial operad 0), with the empty
generating set.
A weak algebra for this theory (with respect to this generating set) is simply
a category.
${\ensuremath{F_{\Sigma}}}$ is a left adjoint, and hence preserves colimits, so ${\ensuremath{F_{\Sigma}}}\emptyset$ is
the initial operad, and the coequalizer $\phi : {\ensuremath{F_{\Sigma}}}\emptyset \to 0$ is
therefore the identity.
Hence $\Wkwrt 0 \phi$ is also the initial operad, and so a $\phi$-weak
0-category is just a category.
A $\phi$-weak $0$-functor is just a functor.
\end{example}
\begin{example}
\label{ex:ptd_sets}
\index{pointed sets}
Consider the operad $P$ of Example \ref{ex:pointed_sets}, generated by one
nullary operation $*$.
Let $\phi$ be the associated regular epi.
Then $\Wkwrt P \phi$ has one nullary object and no objects of any other arity; the only arrow is the identity on the unique nullary object.
In fact, $\Wkwrt P \phi = D_* P$.
So a weak algebra for this theory and this generating set is a category \ensuremath{\mathcal{C}}\
with a distinguished object $\hat * \in \ensuremath{\mathcal{C}}$.
A $\phi$-weak $P$-functor from $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}}))$ to $(\ensuremath{\mathcal{D}}, (\bar{\phantom{\alpha}}))$ is a
functor $F : \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{D}}$ and an isomorphism $\bar * \stackrel{\sim}{\to} \hat *$.
\end{example}
\begin{example}
\label{ex:ptd_sets_ABCD}
\index{pointed sets}
Consider again the operad $P$ of Example \ref{ex:pointed_sets}, this time
generated by four nullary operations $A,B,C,D$ (which are all set equal to each
other).
Let $\phi$ be the associated regular epi.
Then $\Wkwrt P \phi_0$ is the indiscrete category on the four objects $A, B, C,
D$, and $\Wkwrt P \phi_i$ is empty for all other $i \in \natural$.
Hence a $\phi$-weak $P$-category is a category \ensuremath{\mathcal{C}}\ containing four specified
objects $\hat A, \hat B, \hat C$ and $\hat D$.
These four objects are isomorphic via specified isomorphisms $\delta_{AB},
\delta_{AC}, \delta_{AD}$ etc, and all diagrams involving these isomorphisms
commute:
\[
\xymatrixcolsep{6pc}
\xymatrixrowsep{4pc}
\xymatrix{
\hat A \ar[r]^{\delta_{AB}} \ar[dr]_(0.3){\delta_{AD}} \ar[d]_{\delta_{AC}}
& \hat B \ar[d]^{\delta_{BD}} \ar[dl]^(0.3){\delta_{BC}} \\
\hat C \ar[r]_{\delta_{CD}} & \hat D
}
\]
and $\delta_{XY}\delta_{YX} = 1_{\hat X}$ for all $X,Y \in \{A,B,C,D\}$.
Let $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}}))$ and $(\ensuremath{\mathcal{D}}, (\bar{\phantom{\alpha}}))$ be $\phi$-weak $P$-categories.
A $\phi$-weak $P$-functor $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}})) \to (\ensuremath{\mathcal{D}}, (\bar{\phantom{\alpha}}))$ consists of
\begin{itemize}
\item a functor $F : \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{D}}$,
\item an isomorphism $\phi_{XY} : \bar X \stackrel{\sim}{\to} F
\hat X$ for all $X \in \{A, B, C, D\}$,
\end{itemize}
such that, for all $X, Y \in \{A, B, C, D\}$, there is precisely one
isomorphism $\bar X \to F \hat Y$ formed by compositions of $\delta$s and
$\phi$s.
\end{example}
\section{Symmetric monoidal categories}
\index{commutative monoids}
\label{sec:symmmoncats}
\def\ensuremath{(\ensuremath{\mathcal{C}}, \otimes, I, \alpha, \lambda, \rho, \tau)}{\ensuremath{(\ensuremath{\mathcal{C}}, \otimes, I, \alpha, \lambda, \rho, \tau)}}
\def\ensuremath{(\ensuremath{\mathcal{C}}, \otimes', I', \alpha', \lambda', \rho', \tau')}{\ensuremath{(\ensuremath{\mathcal{C}}, \otimes', I', \alpha', \lambda', \rho', \tau')}}
\def\Csmcpprime{\ensuremath{(\ensuremath{\mathcal{C}}', \otimes', I', \alpha', \lambda', \rho',
\tau')}}
\def{\mbox{\upshape{can}}}{{\mbox{\upshape{can}}}}
\def{\Wkwrt P \phi}{{\Wkwrt P \phi}}
Consider the terminal symmetric operad $P$, whose algebras in \cat{Set}\ are
commutative monoids, and the following linear presentation $(\Phi, E)$ for $P$:
\begin{itemize}
\item $\Phi_0 = \{e\}, \Phi_2 = \{.\}$, all other $\Phi_i$s are empty;
\item $E$ contains the equations
\begin{enumerate}
\item $x_1.(x_2.x_3) = (x_1.x_2).x_3$
\item $e.x_1 = x_1$
\item $x_1.e = x_1$
\item $x_1.x_2 = x_2.x_1$
\end{enumerate}
\end{itemize}
This linear presentation gives rise to a symmetric-operadic presentation
$(\Phi, E)$, as described in Lemma \ref{lem:syntactic_semantic_equiv}.
Let $\phi : {\ensuremath{F_{\Sigma}}} \Phi \to P$ be the coequalizer in the diagram
\[
\NoCompileMatrices
\fork {{\ensuremath{F_{\Sigma}}} E} {} {} {{\ensuremath{F_{\Sigma}}} \Phi} \phi P
\]
We shall now prove that the algebras for ${\Wkwrt P \phi}$ are classical symmetric monoidal
categories.
More precisely, we shall show the following:
\begin{enumerate}
\item for a given category \ensuremath{\mathcal{C}}, the ${\Wkwrt P \phi}$-algebra structures on \ensuremath{\mathcal{C}}\ are in
one-to-one correspondence with the symmetric monoidal category\ structures on \ensuremath{\mathcal{C}};
\item there exists an isomorphism (which we construct) between the category
$\cat{Wk-$P$-Cat}$ and the category of symmetric monoidal categories and weak functors;
\item the isomorphism in (2) respects the correspondence in (1).
\end{enumerate}
To fix notation, we recall the classical notions of symmetric monoidal category
and symmetric monoidal functor:
\begin{defn}
\index{symmetric monoidal category}
A \defterm{symmetric monoidal category} is a 7-tuple $\ensuremath{(\ensuremath{\mathcal{C}}, \otimes, I, \alpha, \lambda, \rho, \tau)}$, where
\begin{itemize}
\item $\ensuremath{\mathcal{C}}$ is a category;
\item $\otimes : \ensuremath{\mathcal{C}} \times \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{C}}$ is a functor;
\item $I$ is an object of $\ensuremath{\mathcal{C}}$,
\item $\alpha: A \otimes (B \otimes C) \to (A \otimes B) \otimes C$ is natural
in $A, B, C \in \ensuremath{\mathcal{C}}$;
\item $\lambda : I \otimes A \to A$ and $\rho : A \otimes I \to A$ are natural in $A \in \ensuremath{\mathcal{C}}$;
\item $\tau : A \otimes B \to B \otimes A$ is natural in $A, B \in \ensuremath{\mathcal{C}}$,
\end{itemize}
$\alpha, \lambda, \rho, \tau$ are all invertible, and the following diagrams
commute:
\begin{equation}
\xymatrix{
& (A \otimes B) \otimes (C \otimes D) \ar[ddr]^\alpha \\
A \otimes (B \otimes (C \otimes D)) \ar[dd]_{1\otimes\alpha} \ar[ur]^\alpha \\
& & ((A \otimes B) \otimes C) \otimes D \\
A \otimes ((B \otimes C) \otimes D) \ar[dr]_\alpha \\
& (A \otimes (B \otimes C) ) \otimes D \ar[uur]_{\alpha \otimes 1}
}
\end{equation}
\begin{equation}
\xymatrix{
A \otimes (I \otimes C) \ar[rr]^\alpha \ar[ddr]_{1 \otimes \lambda}
& & (A \otimes I) \otimes C \ar[ddl]^{\rho \otimes 1} \\ \\
& A \otimes C
}
\end{equation}
\begin{equation}
\xymatrix{
A \otimes I \ar[rr]^\tau \ar[ddr]_\rho && I \otimes A\ar[ddl]^\lambda \\ \\
& A
}
\end{equation}
\begin{equation}
\xymatrix{
(A \otimes B) \otimes C \ar[r]^\tau \ar[d]_{\alpha^{-1}}
& C \otimes (A \otimes B) \ar[d]^\alpha \\
A \otimes (B \otimes C) \ar[d]_{1 \otimes \tau}
& (C \otimes A) \otimes B \ar[d]^{\tau \otimes 1} \\
A \otimes (C \otimes B) \ar[r]^\alpha
& (A \otimes C) \otimes B
}
\hskip 2em
\xymatrix{
A \otimes (B \otimes C) \ar[r]^\tau \ar[d]_\alpha
& (B \otimes C) \otimes A \ar[d]^{\alpha^{-1}} \\
(A \otimes B) \otimes C \ar[d]_{\tau \otimes 1}
& B \otimes (C \otimes A) \ar[d]^{1 \otimes \tau} \\
(B \otimes A) \otimes C \ar[r]^{\alpha^{-1}}
& B \otimes (A \otimes C)
}
\end{equation}
\begin{equation}
\xymatrix{
A \otimes B \ar[dr]_1 \ar[r]^{\tau_{A,B}} & B \otimes A \ar[d]^{\tau_{B,A}} \\
& A \otimes B.
}
\end{equation}
\end{defn}
\begin{defn}
Let $M = \ensuremath{(\ensuremath{\mathcal{C}}, \otimes, I, \alpha, \lambda, \rho, \tau)}$ and $N = \Csmcpprime$ be symmetric monoidal categories.
\index{symmetric monoidal functor}
A \defterm{lax symmetric monoidal functor} $F: M \to N$ consists of
\begin{itemize}
\item a functor $F : \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{C}}'$,
\item morphisms $F_2 : (F A) \otimes' (F B) \to F(A \otimes B)$, natural in
$A,B \in \ensuremath{\mathcal{C}}$,
\item a morphism $F_0 : I' \to F I$ in $\ensuremath{\mathcal{C}}'$,
\end{itemize}
such that the following diagrams commute:
\begin{equation}
\label{eqn:symm_mon_fctr_assoc}
\xymatrix{
FA \otimes' (FB \otimes' FC) \ar[r]^{\alpha'} \ar[d]_{1 \otimes' F_2}
& (FA \otimes' FB) \otimes' FC \ar[d]^{F_2 \otimes' 1} \\
FA \otimes' (F(B \otimes C)) \ar[d]_{F_2}
& F(A \otimes B) \otimes' FC \ar[d]^{F_2} \\
F(A \otimes (B \otimes C)) \ar[r]^{F\alpha}
& F((A \otimes B) \otimes C)
}
\end{equation}
\begin{equation}
\label{eqn:symm_mon_fctr_unit}
\xymatrix{
(FB) \otimes' I' \ar[r]^{\rho'} \ar[d]_{1 \otimes' F_0}FB
& F B \\
(FB) \otimes' (FI) \ar[r]^{F_2}
& F(B \otimes I) \ar[u]_{F\rho}
}
\hskip 2em
\xymatrix{
I' \otimes' (FB) \ar[r]^{\lambda'}
& FB \\
(FI) \otimes' (FB) \ar[r]^{F_2} \ar[u]^{F_0 \otimes' 1}
& F(I \otimes B) \ar[u]_{F\lambda}
}
\end{equation}
\begin{equation}
\label{eqn:symm_mon_fctr_tau}
\xymatrix{
(FA)\otimes'(FB) \ar[r]^{\tau'} \ar[d]_{F_2}
& (FB) \otimes' (FA) \ar[d]^{F_2} \\
F(A \otimes B) \ar[r]^{F\tau}
& F(B \otimes A).
}
\end{equation}
$F$ is said to be \defterm{weak} when $F_0, F_2$ are isomorphisms, and
\defterm{strict} when $F_0, F_2$ are identities.
\index{lax!symmetric monoidal functor}
\index{weak!symmetric monoidal functor}
\index{strict!symmetric monoidal functor}
\end{defn}
Recall also the coherence theorem for classical symmetric monoidal categories.
For any $n$-ary\ permuted $\Phi$-tree $(\sigma \act t)$, let $(\sigma \act t)_M$
be the functor $M^n \to M$ obtained by replacing every $.$ in $t$ by $\otimes$
and every $e$ by $I$, and permuting the arguments according to $\sigma$, so
$(\sigma \act t)_M(A_1, \dots, A_n) = t_M(A_{\sigma 1}, \dots, A_{\sigma n})$
for all $A_1, \dots, A_n \in M$.
In particular, we do not make use of the symmetry maps on $M$ in constructing
these functors.
Then:
\begin{theorem}
\label{thm:smc_coherence}
(Mac Lane) \index{Mac Lane, Saunders}
\index{coherence!for symmetric monoidal categories}
In each weak symmetric monoidal category $M$ there is a function which assigns
to each pair $(\sigma \act t_1, \rho \act t_2)$ of permuted $\Phi$-trees of
the same arity $n$ a unique natural isomorphism
\[
{\mbox{\upshape{can}}}_M (\sigma \act t_1, \rho \act t_2) : (\sigma \act t_1)_M
\to (\rho \act t_2)_M : M^n \to M
\]
called the \defterm{canonical map} from $\sigma \act t_1$ to $\rho \act t_2$,
in such a way that the identity of $M$ and all instances of $\alpha, \lambda,
\rho$ and $\tau$ are canonical, and the composite as well as the
$\otimes$-product of two canonical maps is canonical.
\end{theorem}
\begin{proof}
See \cite{catwork} XI.1.
\end{proof}
Finally, recall the coherence theorem for weak monoidal functors:
\begin{lemma}
\label{lem:mon_fctr_coherence}
\index{coherence!for monoidal functors}
Let $M, N$ be monoidal categories, and $F : M \to N$ be a weak monoidal
functor.
For every $n \in \natural$ and every strongly regular $\Phi$-tree $v$ of arity
$n$, there is a unique map $F_v : v_N(F A_1, \dots, F A_n) \to Fv_M(A_1, \dots,
A_n)$ natural in $A_1, \dots, A_n \in M$ and formed by taking composites and
tensors of $F_0$ and $F_2$, such that the diagram
\[
\commsquare{v_N(F A_1, \dots F A_n)}{F_{v}}
{Fv_M(A_1, \dots, A_n)} {{\mbox{\upshape{can}}}_N}{F{\mbox{\upshape{can}}}_M}
{w_N(F A_1, \dots F A_n)}
{F_{w}}{F(w)_M(A_1, \dots, A_n)}
\]
commutes for all $n \in \natural$, all $v, w \in ({\ensuremath{F_{\rm pl}}} \Phi)_n$, and all $A_1,
\dots, A_n \in \ensuremath{\mathcal{M}}$.
\end{lemma}
\begin{proof}
See \cite{catwork}, p. 257.
\end{proof}
We may use this result to sketch a proof of a coherence theorem for weak
symmetric monoidal functors:
\begin{theorem}
\label{thm:smcf_coherence}
\index{coherence!for symmetric monoidal functors}
Let $M,N$ be symmetric monoidal categories, and $F:M \to N$ be a weak symmetric monoidal functor.
Let $\sigma\act v$ be an $n$-ary\ permuted $\Phi$-tree.
Then there is a unique natural transformation
\[
\xymatrix{
M^n \ar[r]^{F^n} \ar[d]_{(\sigma\act v)_M}
\drtwocell\omit{^*{!(1,-1)\object{F_{\sigma\act v}}}}
& N^n \ar[d]^{(\sigma\act v)_N} \\
M \ar[r]_F & N
}
\]
formed by composing tensor products of $F_2$ and $F_0$, possibly with their
arguments permuted.
Furthermore, if $\rho\act w$ is another permuted $\Phi$-tree, then the diagram
\[
\commsquare{(\sigma\act v)_N(F A_1, \dots F A_n)}{F_{\sigma\act v}}
{F(\sigma \act v)_M(A_1, \dots, A_n)} {{\mbox{\upshape{can}}}_N}{F{\mbox{\upshape{can}}}_M}
{(\rho \act w)_N(F A_1, \dots F A_n)}
{F_{\rho \act w}}{F(\rho\act w)_M(A_1, \dots, A_n)}
\]
commutes.
\end{theorem}
\begin{proof}
\def\mbox{\upshape{perm}}{\mbox{\upshape{perm}}}
Let $F_{\sigma\act v}(A_1, \dots, A_n) = F_v(A_{\sigma(1)}, \dots,
A_{\sigma(n)})$, and similarly on morphisms.
Then $F_{\sigma\act v}$ has the required type.
We may decompose ${\mbox{\upshape{can}}}_M(\sigma\act v, \rho\act w)$ as
$\mbox{\upshape{perm}}_M(\sigma,\rho)\ {\mbox{\upshape{can}}}_M(v, w)$, where $\mbox{\upshape{perm}}_M(\sigma,\rho) :
F_{\sigma\act v} \to F_{\rho\act v}$ is a composite of $\tau$s.
Equation \ref{eqn:symm_mon_fctr_tau} and Lemma \ref{lem:mon_fctr_coherence}
together imply that the diagram
\[
\xymatrix{
(\sigma\act v)_N(F A_1, \dots F A_n)\ar[r]^{F_{\sigma\act v}} \ar[d]_{\mbox{\upshape{perm}}_N}
& {F((\sigma \act v)_M(A_1, \dots, A_n))} \ar[d]^{F\mbox{\upshape{perm}}_M} \\
(\rho\act v)_N(F A_1, \dots F A_n)\ar[r]^{F_{\rho\act v}} \ar[d]_{{\mbox{\upshape{can}}}_N}
& {F((\rho \act v)_M(A_1, \dots, A_n))} \ar[d]^{F{\mbox{\upshape{can}}}_M} \\
{(\rho \act w)_N(F A_1, \dots F A_n)} \ar[r]^{F_{\rho \act w}}
& {F((\rho\act w)_M(A_1, \dots, A_n))}
}
\]
commutes.
It remains to show that $F_{\sigma \act v}$ is unique with this property.
Suppose that $F_{\sigma\act v}$ is not unique for some $\sigma \act v$, and
that there exists some natural transformation $G: (\sigma \act v)_N(FA_1,
\dots, FA_n) \to F((\sigma\act v)_M (A_1, \dots, A_n))$, composed of tensor
products of components of $F_0$ and $F_2$, such that $G \neq F_{\sigma\act
v}$.
Suppose further that $\sigma\act v$ and $G$ have been chosen to be a minimal
counterexample, in the sense that of all such counterexamples, $\sigma$ may be
written as a product of the smallest number of transpositions.
If no transpositions are used, then we have a contradiction, because then
$\sigma = 1_n$, and Lemma \ref{lem:mon_fctr_coherence} tells us that $G = F_v$.
But suppose $\sigma = t_1t_2\dots t_m$ where each $t_i$ is a transposition:
then $t_1 \act G$ is a natural transformation $(\sigma\act v)_N (FA_{t_1
1}, \dots, FA_{t_1 n}) \to F((\sigma\act v)_M(A_{t_1 1}, \dots, A_{t_1 n}))$,
and thus a transformation $(t_1\sigma\act v)_N (FA_\bullet) \to F((t_1\sigma\act v)_M(A_\bullet))$.
But $t_1\sigma = t_2t_3\dots t_m$, and thus (by minimality of $\sigma$), it
must be the case that $t_1 \act G = F_{t_1\sigma\act v} = t_1\act F_{\sigma\act
v}$.
Hence $G = F_{\sigma \act v}$.
\end{proof}
We now proceed to relate the classical theory of symmetric monoidal categories\ to the more general
notion of categorification we developed in previous sections.
By Lemma \ref{lem:explicit_wkp}, if $\tau_1, \tau_2$ are $n$-ary\ 1-cells in ${\Wkwrt P \phi}$
(in other words $n$-ary\ permuted $\Phi$-trees), there is a (unique) 2-cell
$\tau_1 \to \tau_2$ in ${\Wkwrt P \phi}$ iff $\tau_1 \sim \tau_2$ under the congruence
generated by $E$.
By standard properties of commutative monoids, this relation holds iff $\tau_1$
and $\tau_2$ take the same number of arguments, so there is exactly one 2-cell
$\tau_1 \to \tau_2$ for every $n \in \natural$ and every pair $(\tau_1,
\tau_2)$ of $n$-ary\ 1-cells in ${\Wkwrt P \phi}$.
Let $\cat{SMC}$ denote the category of symmetric monoidal categories\ and weak maps between them.
We shall define functors $S : \cat{SMC} \to \cat{Wk-$P$-Cat}$ and $R : \cat{Wk-$P$-Cat} \to
\cat{SMC}$, and show that they are inverses of each other.
Let $M = \ensuremath{(\ensuremath{\mathcal{C}}, \otimes, I, \alpha, \lambda, \rho, \tau)}$ be a symmetric monoidal category.
Let $SM$ be the weak $P$-category $(\hat{\phantom{\alpha}}) :{\Wkwrt P \phi} \to \mathop{\rm{End}}(\ensuremath{\mathcal{C}})$ defined as
follows:
\begin{itemize}
\item On 1-cells of ${\Wkwrt P \phi}$, $(\hat{\phantom{\alpha}})$ is determined by $\hat . = \otimes$ and
$\hat e = I$.
\item If $\delta : \tau_1 \to \tau_2$ is an $n$-ary\ 2-cell in ${\Wkwrt P \phi}$ (i.e. a
morphism in the category ${\Wkwrt P \phi}_n$), let $\hat \delta$ be the canonical map $\hat
\tau_1 \to \hat \tau_2$.
\end{itemize}
\begin{lemma}
$SM$ is a well-defined ${\Wkwrt P \phi}$-algebra for all $M \in \cat{SMC}$.
\end{lemma}
\begin{proof}
The 1-cells of ${\Wkwrt P \phi}$ are the same as those of ${\ensuremath{F_{\Sigma}}} \Phi$; hence, $(\hat{\phantom{\alpha}})$ is
entirely determined on 1-cells by a map of signatures $\Phi \to {\ensuremath{U^{\Sigma}}} \mathop{\rm{End}}(\ensuremath{\mathcal{C}})$,
which we have given.
On 2-cells, Theorem \ref{thm:smc_coherence} and the uniqueness property of
2-cells in ${\Wkwrt P \phi}$ tell us that if $\delta_1, \delta_2$ are 2-cells in ${\Wkwrt P \phi}$, then
$\widehat{\delta_1 . \delta_2} = \hat \delta_1 \otimes \hat \delta_2 = \hat
\delta_1 \hat . \hat \delta_2$, and $\widehat{\delta_1 \delta_2} =
\hat{\delta_1} \hat{\delta_2}$ wherever $\delta_1, \delta_2$ are composable.
Hence, $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}}))$ is a well-defined ${\Wkwrt P \phi}$-algebra.
\end{proof}
Given symmetric monoidal categories $M$ and $N$, and a weak symmetric monoidal
functor $F : M \to N$, we would like to define a weak $P$-functor $SF = (F,
\psi): SM \to SN$.
Let $\psi_{\sigma\act v, A_\bullet} = F_{\sigma\act v}$ for all $n \in \natural$, all $\sigma\act v \in ({\ensuremath{F_{\Sigma}}} \Phi)_n$, and all $A_1, \dots, A_n \in M$.
By Theorem \ref{thm:smcf_coherence}, this is natural in $\sigma\act v$ and in
$A_1, \dots, A_n$.
The other axioms for a weak $P$-functor are all implied by the coherence
theorem (Theorem \ref{thm:smcf_coherence}).
This can be generalized: a lax symmetric monoidal functor $F$ determines a
lax $P$-functor $SF$, and a strict symmetric monoidal functor $F$
determines a strict $P$-functor $SF$.
Now, let \ensuremath{\mathcal{C}}\ be a ${\Wkwrt P \phi}$-algebra, with map $(\hat{\phantom{\alpha}}): {\Wkwrt P \phi} \to \mathop{\rm{End}}(\ensuremath{\mathcal{C}})$.
We shall construct a symmetric monoidal category\ $R(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}})) = \ensuremath{(\ensuremath{\mathcal{C}}, \otimes, I, \alpha, \lambda, \rho, \tau)}$.
Take
\begin{itemize}
\item $\otimes = \hat .$
\item $ I = \hat e$
\item $ \alpha = \hat \delta_1$, where $\delta_1 : -.(-.-) \to (-.-).-$ in
${\Wkwrt P \phi}_3$,
\item $ \lambda = \hat \delta_2$, where $\delta_2 : e.- \to -$ in ${\Wkwrt P \phi}_1$,
\item $ \rho = \hat \delta_3$, where $\delta_3 : -.e \to e$ in ${\Wkwrt P \phi}_1$,
\item $ \tau = \hat \delta_4 $, where $\delta_4 : (-.-) \to (12)\cdot(-.-)$ in ${\Wkwrt P \phi}_2$.
\end{itemize}
\begin{lemma}
$R(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}}))$ is a symmetric monoidal category.
\end{lemma}
\begin{proof}
Because there is at most one 2-cell $\tau_1 \to \tau_2$ for any pair of 1-cells
$\tau_1, \tau_2$ in ${\Wkwrt P \phi}$, all diagrams involving these commute.
In particular, the axioms for a symmetric monoidal category\ are satisfied.
The 2-cells in $\mathop{\rm{End}}(\ensuremath{\mathcal{C}})$ are natural transformations, so $\alpha, \lambda,
\rho$ and $\tau$ (as images of 2-cells in ${\Wkwrt P \phi}$ under the map $(\hat{\phantom{\alpha}}) : {\Wkwrt P \phi} \to
\mathop{\rm{End}}(\ensuremath{\mathcal{C}})$) are natural transformations.
All 2-cells in ${\Wkwrt P \phi}$ are invertible, so $\alpha, \lambda, \rho$ and $\tau$ are
all natural isomorphisms.
Hence $(\ensuremath{\mathcal{C}}, \otimes, I, \alpha, \lambda, \rho, \tau)$ is a symmetric monoidal category.
\end{proof}
Let $(F, \psi) : (\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}})) \to (\ensuremath{\mathcal{C}}', (\check{\phantom{\alpha}}))$ be a weak morphism of
${\Wkwrt P \phi}$-algebras.
Then let $R(F,\psi) : R(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}})) \to R(\ensuremath{\mathcal{C}}', (\check{\phantom{\alpha}}))$ be the
following symmetric monoidal functor:
\begin{itemize}
\item the underlying functor is $F$,
\item $F_0$ is $\psi_{e \ocomp 1}: \check e \to F \hat e$,
\item $F_2$ is $\psi_{. \ocomp 1}: (\check . )F^2 \to F (\hat .)$.
\end{itemize}
The coherence diagrams (\ref{eqn:symm_mon_fctr_assoc}),
(\ref{eqn:symm_mon_fctr_unit}) and (\ref{eqn:symm_mon_fctr_tau}) all commute
by virtue of the coherence axioms for a weak morphism of ${\Wkwrt P \phi}$-algebras and the
naturality of $\psi$.
Hence $(F, F_0, F_2)$ is a symmetric monoidal functor.
\begin{lemma}
\label{lem:RS = 1}
Let \ensuremath{(\ensuremath{\mathcal{C}}, \otimes, I, \alpha, \lambda, \rho, \tau)}\ be a symmetric monoidal category. Then \[ RS\ensuremath{(\ensuremath{\mathcal{C}}, \otimes, I, \alpha, \lambda, \rho, \tau)} = \ensuremath{(\ensuremath{\mathcal{C}}, \otimes, I, \alpha, \lambda, \rho, \tau)}. \]
\end{lemma}
\begin{proof}
Let $RS\ensuremath{(\ensuremath{\mathcal{C}}, \otimes, I, \alpha, \lambda, \rho, \tau)} = \ensuremath{(\ensuremath{\mathcal{C}}, \otimes', I', \alpha', \lambda', \rho', \tau')}$.
Their underlying categories are equal, both being $\ensuremath{\mathcal{C}}$.
\[
\begin{array}{rcccl}
\otimes' &= &\hat . &= &\otimes \\
I' &= & \hat e &= & I \\
\alpha' & = &\hat \delta_1 & = & \alpha \mbox{, the unique canonical map of
the correct type}\\
\lambda' & = &\hat \delta_2 & = & \lambda \\
\rho' & = &\hat \delta_3 & = & \rho \\
\tau' & = &\hat \delta_4 & = & \tau \\
\end{array}
\]
\end{proof}
\begin{lemma}
\label{lem:SR = 1}
Let $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}}))$ be a ${\Wkwrt P \phi}$-algebra, and let
$(\ensuremath{\mathcal{C}}', (\check{\phantom{\alpha}})) = SR(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}}))$.
Then $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}})) = (\ensuremath{\mathcal{C}}', (\check{\phantom{\alpha}}))$.
\end{lemma}
\begin{proof}
Their underlying categories are the same.
As above, $(\check{\phantom{\alpha}})$ is determined on objects by the values it takes on $.$
and $e$: these are $\otimes = \hat .$ and $I = \hat e$ respectively.
So $(\check{\phantom{\alpha}}) = (\hat{\phantom{\alpha}})$ on objects.
If $\delta : \tau_1 \to \tau_2$, then $\check \delta$ is the unique canonical
map from $\check \tau_1 \to \check \tau_2$, which, by an easy induction, must be
$\hat \delta$.
So $(\check{\phantom{\alpha}}) = (\hat{\phantom{\alpha}})$, and hence $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}})) = SR(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}}))$.
\end{proof}
\begin{lemma}
\label{lem:RS = 1 on morphisms}
Let $M = \ensuremath{(\ensuremath{\mathcal{C}}, \otimes, I, \alpha, \lambda, \rho, \tau)}$ and $N = \Csmcpprime$ be symmetric monoidal categories, and let $(F, F_0, F_2)$ be a
weak symmetric monoidal functor $M \to N$.
Then $RS(F, F_0, F_2) = (F, F_0, F_2)$.
\end{lemma}
\begin{proof}
Let $(G, G_0, G_2) = RS(F, F_0, F_2)$.
Then $G$ is the underlying functor of $S(F, F_0, F_2)$ which is $F$, and
$G_0, G_2$ are both the canonical maps with the correct types given by
Theorem \ref{thm:smcf_coherence}: that is to say, they are $F_0$ and $F_2$
respectively.
\end{proof}
\begin{lemma}
\label{lem:SR = 1 on morphisms}
Let $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}}))$ and $(\ensuremath{\mathcal{C}}', (\check{\phantom{\alpha}}))$ be ${\Wkwrt P \phi}$-algebras, and let $(F,
\phi) : (\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}})) \to (\ensuremath{\mathcal{C}}', (\check{\phantom{\alpha}}))$ be a weak morphism of
${\Wkwrt P \phi}$-algebras.
Then $SR(F, \phi) = (F, \phi)$.
\end{lemma}
\begin{proof}
Let $(G, \gamma) = SR(F, \phi)$.
Then $G$ is the underlying functor of $R(F, \phi)$, which is $F$.
Let $(F, F_0, F_2) = R(F, \phi)$.
Each component of $\gamma$ is then by definition the correct component of the
canonical map arising from $F_0, F_2$ in the process described in Theorem
\ref{thm:smcf_coherence}.
By the ``uniqueness'' part of the Theorem, this must be the corresponding
component of $\phi$.
Hence $\gamma = \phi$.
\end{proof}
\begin{theorem}
$S$ and $R$ form an isomorphism of categories $\cat{SMC} \cong \cat{Wk-$P$-Cat}$.
\end{theorem}
\begin{proof}
Lemmas \ref{lem:RS = 1} and \ref{lem:SR = 1} show that $R$ and $S$ are
bijective on objects; Lemmas \ref{lem:RS = 1 on morphisms} and \ref{lem:SR = 1
on morphisms} show that $R$ and $S$ are locally bijective on morphisms.
Hence, $R$ and $S$ are a pair of mutually inverse isomorphisms of categories.
\end{proof}
\section{Multicategories}
We can tell this whole story for (symmetric) multicategories as well as just
operads.
We sketch this development briefly here, although the remainder of the thesis
will continue to focus on the special case of operads.
\begin{defn} A (directed) \defterm{multigraph} consists of
\index{multigraph}
\begin{enumerate}
\item a set of \defterm{vertices} $V$,
\item for each $n \in \natural$ and each sequence $v_1, v_2, \dots,
v_n, w$ of vertices, a set $E(v_1, \dots, v_n; w)$ of \defterm{funnels from
$v_1, \dots, v_n$ to $w$}.
\index{funnel}
\end{enumerate}
\end{defn}
\begin{defn} Let $M_1 = (V_1, E_1)$ and $M_2 = (V_2, E_2)$ be multigraphs.
\index{morphism!of multigraphs}
A \defterm{morphism of multigraphs} $f: M_1 \to M_2$ is
\begin{enumerate}
\item a function $f_V : V_1 \to V_2$,
\item for each finite sequence $v_1, v_2 \dots v_n, w$ of vertices in $M_1$, a
function
\[
f^{v_1,\dots,v_n}_w : E_1(v_1, \dots, v_n; w)
\to E_2(f_V(v_1), \dots, f_V(v_n); f_V(w)).
\]
\end{enumerate}
\end{defn}
We say that a funnel $f \in E(v_1, \dots, v_n; w)$ has \defterm{source} $v_1,
\dots, v_n$ and \defterm{target} $w$; we say that two funnels are
\defterm{parallel} if they have the same source and target.
The reason for the ``funnel'' terminology should be clear from Figure
\ref{fig:mgraph}.
We shall say that a multigraph has some property $P$ \defterm{locally} if every
$E(v_1, \dots, v_n; w)$ is $P$, and similarly a morphism $f$ of multicategories
is locally $P$ if every $f^{v_1,\dots,v_n}_w$ is $P$.
Multigraphs and their morphisms form a category which we shall call \cat{Multigraph}.
\index{\cat{Multigraph}}
\begin{figure}[h]
\centerline{
\epsfxsize=290pt
\epsfbox{multigraph.eps}
}
\caption{A multigraph}
\label{fig:mgraph}
\end{figure}
In order to proceed with the rest of the construction, we will need to consider
subcategories of \cat{Multicat}, \cat{Multigraph}\ etc.
\begin{defn}
\index{$\cat{Multigraph}_X$}
Let $X$ be a set.
Then $\cat{Multigraph}_X$ is the subcategory of $\cat{Multigraph}$ whose objects are
multigraphs with vertex set $X$, and whose morphisms are identity-on-vertices
maps of multigraphs.
We define $\cat{Multicat}_X$ and $\cat{$\Sigma$-Multicat}_X$ similarly.
\end{defn}
For each $X \in \cat{Set}$, there is a chain of adjunctions similar to that given in
Section \ref{sec:adj}:
\index{adjunctions!for multicategories}
\[
\xymatrix{
{\cat{FP-Multicat}}_X \ar@<1.2ex>[d]^{\ensuremath{U^{\rm fp}_{\Sigma}}}
\ar@/r1.4cm/[dd]^{\ensuremath{U^{\rm fp}_{\rm pl}}}
\ar@/r3cm/[ddd]^{\ensuremath{U^{\rm fp}}} \\
\cat{$\Sigma$-Multicat}_X \ar@<1.2ex>[d]^{\ensuremath{U^{\Sigma}_{\rm pl}}} \ar@<1.2ex>[u]^{\ensuremath{F_{\rm fp}^{\Sigma}}}_{\dashv}
\ar@/r1.4cm/[dd]^{\ensuremath{U^{\Sigma}}} \\
\cat{Multicat}_X \ar@<1.2ex>[u]^{\ensuremath{F_{\Sigma}^{\rm pl}}}_{\dashv} \ar@<1.2ex>[d]^{\ensuremath{U^{\rm pl}}}
\ar@/l1.4cm/[uu]^{\ensuremath{F_{\rm fp}^{\rm pl}}} \\
\cat{Multigraph}_X \ar@<1.2ex>[u]^{\ensuremath{F_{\rm pl}}}_{\dashv} \ar@/l1.4cm/[uu]^{\ensuremath{F_{\Sigma}}}
\ar@/l3cm/[uuu]^{\ensuremath{F_{\rm fp}}}
}
\]
These adjunctions are monadic, by Lemma \ref{lem:frees_exist} and Lemma
\ref{lem:monadj}.
Note that $\cat{Set}^\natural$ can be regarded as $\cat{Multigraph}_1$: thus, the
adjunctions of Section \ref{sec:adj} are just the restrictions of the
adjunctions above to the one-vertex case.
We can consider multigraphs enriched in some category \ensuremath{\mathcal{V}}:
\begin{defn}
\index{multigraph!enriched}
Let \ensuremath{\mathcal{V}}\ be a category.
A \defterm{\ensuremath{\mathcal{V}}-multigraph} $M = (V,E)$ consists of
\begin{enumerate}
\item a set $V$ of \defterm{vertices},
\item for each $n \in \natural$ and each finite sequence $v_1, v_2 \dots v_n, w$
of vertices, an object of \ensuremath{\mathcal{V}}\ called $E(v_1, \dots, v_n; w)$ of
\defterm{funnels from $v_1, \dots, v_n$ to $w$}.
\end{enumerate}
\end{defn}
\begin{defn} Let $M_1 = (V_1, E_1)$ and $M_2 = (V_2, E_2)$ be \ensuremath{\mathcal{V}}-multigraphs.
A \defterm{morphism of \ensuremath{\mathcal{V}}-multigraphs} $f: M_1 \to M_2$ is
\index{morphism!of enriched multigraphs}
\begin{enumerate}
\item a function $f_V : V_1 \to V_2$,
\item for each $n \in \natural$ and each finite sequence $v_1, v_2 \dots v_n,
w$ of vertices in $M_1$, an arrow $E_1(v_1, \dots, v_n; w)
\to E_2(f_V(v_1), \dots, f_V(v_n); f_V(w))$ in \ensuremath{\mathcal{V}}.
\end{enumerate}
\end{defn}
The category of \ensuremath{\mathcal{V}}-multigraphs and their morphisms is called \cat{\V-Multigraph}.
The category whose objects are \ensuremath{\mathcal{V}}-multigraphs with vertex-set $X$ and whose
morphisms are identity-on-vertices maps is called $\cat{\V-Multigraph}_X$.
In particular, we shall consider multigraphs enriched in the category \cat{Digraph}\
of directed graphs.
An object of the category \cat{Digraph-Multigraph}\ consists of
\begin{enumerate}
\item \defterm{vertices} (or \defterm{objects});
\item \defterm{funnels}, each of which has one object as its target, and a
sequence of objects as its source;
\item \defterm{edges}, which each have one funnel as a source and one as a
target: the source and target of a given edge must be parallel.
\end{enumerate}
\begin{figure}[h]
\centerline{
\epsfxsize=290pt
\epsfbox{gphmgph.eps}
}
\caption{A multigraph enriched in directed graphs}
\label{fig:gphmgph}
\end{figure}
The factorization system construction of Example \ref{ex:catopdFS} works in
this broader setting too.
Let $X$ be a set.
\index{factorization system!bijective on objects/full and faithful!on
$\cat{Digraph-Multigraph}_X$}
The factorization system $(\ensuremath{\mathcal{E}}, \ensuremath{\mathcal{M}})$ on \cat{Digraph}\ of Example \ref{ex:digraphFS}
gives rise to a factorization system $(\ensuremath{\mathcal{E}}', \ensuremath{\mathcal{M}}')$ on $\cat{Digraph-Multigraph}_X$,
where \ensuremath{\mathcal{E}}\ consists of maps which are bijective on objects and funnels, and \ensuremath{\mathcal{M}}\
consists of maps which are locally full-and-faithful.
This lifts to a factorization system $(\ensuremath{\mathcal{E}}'', \ensuremath{\mathcal{M}}'')$ on $\cat{Cat-Multigraph}_X$ via
Lemma \ref{lem:monad}.
By the usual argument, there is a chain of monadic adjunctions:
\index{adjunctions!for \cat{Cat}-multicategories}
\[
\xymatrixrowsep{3pc}
\xymatrix{
{\cat{\Cat-FP-Multicat}}_X \ar@<1.2ex>[d]^{\ensuremath{U^{\rm fp}_{\Sigma}}}
\ar@/r1.7cm/[dd]^{\ensuremath{U^{\rm fp}_{\rm pl}}}
\ar@/r3cm/[ddd]^{\ensuremath{U^{\rm fp}}} \\
\cat{Cat-$\Sigma$-Multicat}_X \ar@<1.2ex>[d]^{\ensuremath{U^{\Sigma}_{\rm pl}}} \ar@<1.2ex>[u]^{\ensuremath{F_{\rm fp}^{\Sigma}}}_{\dashv}
\ar@/r1.7cm/[dd]^{\ensuremath{U^{\Sigma}}} \\
\cat{Cat-Multicat}_X \ar@<1.2ex>[u]^{\ensuremath{F_{\Sigma}^{\rm pl}}}_{\dashv} \ar@<1.2ex>[d]^{\ensuremath{U^{\rm pl}}}
\ar@/l1.7cm/[uu]^{\ensuremath{F_{\rm fp}^{\rm pl}}} \\
\cat{Cat-Multigraph}_X \ar@<1.2ex>[u]^{\ensuremath{F_{\rm pl}}}_{\dashv} \ar@/l1.7cm/[uu]^{\ensuremath{F_{\Sigma}}}
\ar@/l3cm/[uuu]^{\ensuremath{F_{\rm fp}}}
}
\]
Since $\cat{Cat-Multicat}_X$ is monadic over $\cat{Cat-Multigraph}_X$, this in turn
lifts to a factorization system on $\cat{Cat-Multicat}_X$.
Similarly, we obtain a factorization system on $\cat{Cat-$\Sigma$-Multicat}_X$.
A \defterm{generator} for a plain multicategory $M$ with object-set $X$ is a
multigraph $\Phi = (X,E)$ together with a regular epi ${\ensuremath{F_{\rm pl}}} \Phi \to M$ in
$\cat{Multicat}_X$.
Similarly, a \defterm{generator} for a symmetric multicategory $M$ with
object-set $X$ is a multigraph $\Phi = (X,E)$ together with a regular epi
${\ensuremath{F_{\Sigma}}} \Phi \to M$ in $\cat{$\Sigma$-Multicat}_X$.
We can therefore extend Definition \ref{def:phi-wkpcat_plain} above, in the
obvious way.
Let $D_*$ be the embedding of $\cat{Multicat}_X$\ into $\cat{Cat-Multicat}_X$ via the
(full and faithful) discrete category functor applied locally.
\begin{defn}
\index{weakening!of a multicategory}
Let $M$ be a plain multicategory with object-set $X$, and let $\phi: {\ensuremath{F_{\rm pl}}}\Phi
\to M$ be a regular epi in $\cat{Multicat}_X$.
Then the \defterm{weakening of $M$ with respect to $\phi$} is the
unique-up-to-isomorphism \cat{Cat}-multicategory $\Wkwrt M \phi$ such that the
following diagram commutes:
\[
\xymatrix{
D_* F\Phi \ar[rr]^{D_*\phi} \booar[dr]_b
&& D_* M \\
& \Wkwrt M \phi \lffar[ur]_f
}
\]
where $f$ is locally full and faithful, and $b$ is locally bijective on objects
(i.e., each map of sets of funnels in $b$ is a bijection).
The uniqueness of $\Wkwrt M \phi$ follows from properties of the factorization
system on $\cat{Cat-Multicat}_X$\ given above.
\end{defn}
\begin{defn}
\index{unbiased!weakening of a multicategory}
Let $M$ be a (symmetric) multicategory.
The \defterm{unbiased weakening} of $M$ is the weakening of $M$ with respect to
the counit $\epsilon$ of the adjunction ${\ensuremath{F_{\rm pl}}} \dashv {\ensuremath{U^{\rm pl}}}$ (respectively, the
adjunction ${\ensuremath{F_{\Sigma}}} \dashv {\ensuremath{U^{\Sigma}}}$).
\end{defn}
\begin{defn}
Let $M$ be a multicategory, and let $\phi: {\ensuremath{F_{\rm pl}}}\Phi \to M$ be a regular epi in
$\cat{Multicat}_X$.
A \defterm{$\phi$-weak $M$-algebra} is an algebra for $\Wkwrt M \phi$.
An \defterm{unbiased weak $M$-algebra} is an algebra for $\Wk M$.
\end{defn}
We define weakenings of symmetric multicategories analogously.
\section{Examples}
\begin{example}
Let $M$ be the multicategory generated by
\[
\xymatrix{0 \ar[r]^f & 1 \ar[r]^g & 2}
\]
Then a weak algebra for $M$ in \cat{Cat}\ with respect to this generating set
consists of a diagram
\[
\xymatrix{\hat 0 \ar[r]^{\hat f} & \hat 1 \ar[r]^{\hat g} & \hat 2}
\]
in \cat{Cat}, whereas an unbiased weak $M$-algebra is a diagram
\[
\xymatrix{
\hat 0 \ar[dr]_{\hat f} \ar[rr]^{\widehat{gf}} \rrlowertwocell\omit{<3>\sim}
\ar@(ul,dl)[]_{\widehat{I_0}}
& & \hat 2 \ar@(ur,dr)[]^{\widehat{I_2}} \\
& \hat 1 \ar[ur]_{\hat g} \ar@(dl,dr)[]_{\widehat{I_1}}
}
\]
where $\widehat{I_0}, \widehat{I_1}$ and $\widehat{I_2}$ are trivial monads.
\end{example}
\begin{example}
\index{$M$-sets}
Consider the theory $T$ whose algebras are a monoid $M$ together with an
$M$-set $A$.
Then a weak $T$-algebra with respect to the standard presentation (a binary and
nullary operation on $M$, and an operation $M \times A \to A$) is a classical
monoidal category $\hat M$, a category $\hat A$, and a weak monoidal
functor $\hat M \to \mathop{\rm{End}}(\hat A)$.
An unbiased weak $T$-algebra is an unbiased monoidal category $\hat M$, a
category $\hat A$ equipped with a trivial monad $\hat I_A$, and an unbiased
monoidal functor $\hat M \to \mathop{\rm{End}}(\hat A)$ which commutes up to coherent
isomorphism with $\hat I_A$.
\end{example}
\begin{example}
\index{weak!$P$-functor}
Let $P$ be an operad, and let $\bar P$ be the multicategory from Section
\ref{sec:Pbar} whose algebras are pairs of $P$-algebras with a morphism between
them.
It seems clear that an unbiased weak $\bar P$-category is a pair of unbiased
weak $P$-categories and an unbiased weak $P$-functor between them; a rigorous
proof would first require a coherence theorem to be proven for weak maps of
$\cat{Cat}$-operad algebras, and currently no such theorem is known.
\end{example}
\section{Evaluation}
\label{sec:evaluation}
At the beginning of this chapter, we proposed three criteria that a successful
definition of categorification should satisfy: namely, it should be broad,
consistent with earlier work, and canonical.
The examples considered throughout the chapter show that our theory agrees with
the standard categorifications that are within its scope.
It is determined by the universal property given by the factorization system on
$\cat{Cat-$\Sigma$-Operad}$: the only tunable parameter is the choice of generator of a
given theory, and in Chapter \ref{ch:coherence} we shall see that the weakening
of a given theory is independent (up to equivalence) of the generator used.
The main problem is the breadth of our theory: as presented, it is restricted
to linear theories, preventing us from categorifying the theories of groups,
rings, Lie algebras, and many other interesting nonlinear theories.
We shall now show what happens when we try to extend our theory to general
algebraic theories.
\begin{lemma}
There is a factorization system $(\ensuremath{\mathcal{E}}, \ensuremath{\mathcal{M}})$ on {\cat{\Cat-FP-Operad}}\, where \ensuremath{\mathcal{E}}\ is the
collection of maps which are bijective on objects, and \ensuremath{\mathcal{M}}\ is the collection of
maps which are levelwise full and faithful.
\end{lemma}
\begin{proof}
The proof is exactly as for the proof of existence of the factorization systems
on \cat{Cat-Operad}\ and \cat{Cat-$\Sigma$-Operad}\ given in Example \ref{ex:catopdFS}.
\end{proof}
\begin{theorem}
Let $P$ be the finite product operad\ whose algebras are commutative monoids, and $D_* : {\cat{FP-Operad}}
\to {\cat{\Cat-FP-Operad}}$ be the levelwise ``discrete category'' functor.
Let $Q$ be the finite product \Cat-operad\ given by the factorization
\[
\xymatrix{
D_* {\ensuremath{F_{\rm fp}}}{\ensuremath{U^{\rm fp}}} P \ar[rr]^{D_* \epsilon_P} \booar[dr] & & D_* P \\
& Q \lffar[ur]
}
\]
Then an algebra for $Q$ is an unbiased symmetric monoidal category $\ensuremath{\mathcal{C}}$ such
that, for all $A \in \ensuremath{\mathcal{C}}$, the component $\tau_{AA}$ of the symmetry map $\tau$
is the identity on $A \otimes A$.
\end{theorem}
\begin{proof}
\def\left[1 \atop 1\right]{\left[1 \atop 1\right]}
We adopt the notation for elements of $P$ introduced in Example
\ref{ex:comm_monoid_fp}.
Let $f$ be the unique function $\fs 2 \to \fs 1$, and let $t : \fs 2 \to \fs 2$
be the permutation transposing 1 and 2.
Then $\epsilon(f \cdot \left[1 \atop 1\right]) = [2] \in P_1$.
Let $(\ensuremath{\mathcal{C}}, (\hat{\phantom{\alpha}}))$ be a $Q$-algebra.
We shall write $\hat \left[1 \atop 1\right](A, B)$ as $A \otimes B$.
We may impose a symmetric monoidal category structure on $\ensuremath{\mathcal{C}}$, where the
symmetry map is the image under $(\hat{\phantom{\alpha}})$ of the unique map $\delta : \left[1 \atop 1\right]
\to t \cdot \left[1 \atop 1\right]$ in $Q_1$.
All diagrams in $Q_1$ commute, so in particular, the following diagram
commutes:
\[
\xymatrix{
[2] \ar[r] & f\cdot\left[1 \atop 1\right] \\
f\cdot\left[1 \atop 1\right] \ar[ur]_{f\cdot\delta} \ar[u]
}
\]
The two unlabelled arrows are mutually inverse.
Applying $(\hat{\phantom{\alpha}})$ to the entire diagram, and evaluating the resulting
functors at $A \in \ensuremath{\mathcal{C}}$, we see that the following diagram commutes:
\[
\xymatrix{
\hat{[2]}(A) \ar[r] & A \otimes A \\
A \otimes A \ar[ur]_{\tau_{A A}} \ar[u]
}
\]
and hence $\tau_{A A} = 1_{A \otimes A}$.
\end{proof}
This is not the case for most interesting symmetric monoidal categories.
Hence this definition of categorification would fail to be consistent with
earlier work.
\chapter{Coherence}
\label{ch:coherence}
\index{Kelly, Max}
\index{coherence!general case}
There are many ``coherence theorems'' in category theory, but in practice they usually fall into one of two classes:
\begin{enumerate}
\item
\label{all_commute}
``All diagrams commute'', or more precisely, that diagrams in a given class
commute if and only if some quantity is invariant.
\item
\label{weak=strict}
Every ``weak'' object is equivalent to an appropriate ``strict'' object.
\end{enumerate}
Since the diagrams of interest in theorems of type \ref{all_commute} will
usually commute trivially in a strict object, a coherence theorem of type
\ref{weak=strict} usually implies one of type \ref{all_commute}.
However, establishing the converse is usually harder.
In the previous chapter, our ``weak $P$-categories'' were defined explicitly in
terms of an infinite class of commuting diagrams (namely, those diagrams which
become identities under the application of the counit of the adjunctions
${\ensuremath{F_{\Sigma}}} \dashv {\ensuremath{U^{\Sigma}}}$ or ${\ensuremath{F_{\rm pl}}} \dashv {\ensuremath{U^{\rm pl}}}$): it is therefore interesting to see
if we can prove a theorem of type \ref{weak=strict} about them.
We do this in Section \ref{sec:strictification}, and investigate an interesting property of the strictification functor in Section \ref{stsect}.
In Section \ref{sec:presindep}, we investigate how the operad defining weak
$P$-categories is affected by our choice of presentation for $P$.
While the independence result obtained is not a coherence theorem of the usual
form, it can be seen as a coherence theorem in a higher-dimensional sense: that
the process of categorification is itself coherent.
For other related work, see Power's paper \cite{power}.
\section{Strictification}
\label{sec:strictification}
Let $P$ be a plain operad, and $Q = \mbox{Wk}(P)$, with $\pi:Q \to D_* P$ the
levelwise full-and-faithful map in Theorem \ref{thm:WkP_characterization}.
We again adopt the ${}_\bullet$ notation from chain complexes and write, for
instance, $p_\bullet$ for a finite sequence of objects in $P$, and $p_\bullet^\bullet$ for
a double sequence.
Let $Q \kel A \toletter{h} A$ be a weak $P$-category.
We shall construct a strict $P$-category \st A and a weak $P$-functor
$(F,\phi):\st A \to A$, and show that it is an equivalence of weak
$P$-categories.
This ``strictification'' construction is closely related to that given in
\cite{j+s} for monoidal categories; however, it is more general, and since we
work for the moment with unbiased weak $P$-categories, our construction has
some additional pleasant properties.
In fact, {\rm\textbf{st} } is functorial, and is left adjoint to the forgetful
functor $\cat{Str-$P$-Cat}\linebreak[0] \to \cat{Wk-$P$-Cat}$ (see Section \ref{stsect}).
The theorem then says that the
unit of this adjunction is pseudo-invertible, and that the strict
$P$-categories and strict $P$-functors form a weakly coreflective
sub-2-category of \cat{Wk-$P$-Cat}.
If $P$ is a plain operad, let $\iota$ be the embedding
\begin{eqnarray*}
\iota : {\ensuremath{U^{\rm pl}}} D_* P & \to & {\ensuremath{U^{\rm pl}}} \Wk{P} \\
\iota(p) & = & p \ocomp (|, \dots, |)
\end{eqnarray*}
Note that this is a morphism in $\cat{Cat}^\natural$, not in $\cat{Cat-Operad}$.
\begin{defn}
\index{strictification}
Let $P$, $Q$, $h$, $A$, $\iota$ be as above. The \defterm{strictification of
$A$}, written \st A, is given by the bijective-on-objects/full and
faithful factorization of $h(\iota \kel 1)$ in $\cat{Cat}$:
\[
\xymatrix{
P\kel A \ar[r]^{\iota \kel 1} \booar[dr] & Q \kel A \ar[r]^h & A. \\
& \st A \lffar[ur]
}
\]
\end{defn}
We shall show that \st A is a strict $P$-category.
We may describe \st A explicitly as follows:
\begin{itemize}
\item An object of \st A is an object of $P \kel A$.
\item If $p \in P_n$ and $a_1, \dots, a_n \in A$, an arrow $(p, a\seq) \to (p',
a\seq')$ in \st A is an arrow $h(p, a\seq) \to h(p', a\seq')$ in $A$.
We say that such an arrow is a \defterm{lifting} of $h(p, a\seq)$.
Composition and identities are as in $A$.
\end{itemize}
We define an action $h'$ of $P$ on \st A as follows:
\begin{itemize}
\item On objects, $h'$ acts by $h'(q, (p, a\seq)^\bullet)
= (\pi(p\cmp {p^\bullet}), a\seq^\bullet)$ where $p \in P_n$ and $(p^i, a\seq^i) \in
\st A$ for $n \in \natural$ and $i = 1, \dots n$.
\item Let $f_i :(p_i, a_i) \to (p_i', a_i')$ for $i = 1, \dots,
n$. Then $h'(p, f_\bullet)$ is the composite
\[
\begin{array}{rcl}
h(p\ocomp(p_\bullet), a\seq)& \toletter{\delta_{p\cmp {p_\bullet}}^{-1}}&
h(p\cmp {p_\bullet}, a\seq) = h(p, h(p_1, a_1), \dots , h(p_n, a_n))\\
&\toletter{h(p, f_\bullet)}
&h(p, h(p_1', a_1'), \dots, h(p_n', a_n'))
= h(p\cmp {p_\bullet'}, a\seq')\\
&\toletter{\delta_{p\cmp {p_\bullet'}}}
&h(p\ocomp(p_\bullet'), a\seq').
\end{array}
\]
\end{itemize}
\begin{lemma}\st A is a strict $P$-category.\end{lemma}
\begin{proof}
It is clear that the action we have defined is strict and associative on
objects and that $1_P$ acts as a unit: we must show that the action on arrows
is associative.
Let $f_i^j:(p_i^j, a_{i\bullet}^j) \to (q_i^j, b_{i\bullet}^j)$, $\sigma \in Q_n$, and
$\tau_i \in Q_{k_i}$ for $j = 1, \dots, k_i$ and $i = 1,\dots,n$.
We wish to show that $h'(\sigma\ocomp(\tau_\bullet), f\seq\udot) = h'(\sigma,
h'(\tau_1, f_1^\bullet), \dots, h'(\tau_n, f_n^\bullet))$.
The LHS is
\[
\begin{array}{rcl}
h(\sigma\ocomp(\tau_\bullet)\ocomp(p\seq\udot), a\seq^\bullet)&
\toletter{\delta_{\sigma\ocomp(\tau_i)\cmp{p\seq\udot}}^{-1}}&
h(\sigma\ocomp(\tau_\bullet), h(p_1^1, a_{1\bullet}^1), \dots, h(p_n^{k_n},
a_{n\bullet}^{k_n}))\\
&\toletter{h(\sigma\ocomp(\tau_\bullet), f\seq\udot)}
&h(\sigma\ocomp(\tau_\bullet), h(q_1^1, b_{1\bullet}^1), \dots, h(q_n^{k'_n},
b_{n\bullet}^{k'_n}))\\
&\toletter{\delta_{\sigma\ocomp(\tau_\bullet)\cmp{q\seq\udot}}}
&h(\sigma\ocomp(\tau_\bullet)\ocomp(q\seq\udot), b\seq\udot).
\end{array}
\]
The RHS is
\[
\begin{array}{rcl}
h(\sigma\ocomp(\tau_\bullet)\ocomp(p\seq\udot), a\seq^\bullet)&
\toletter{\delta_{\sigma\cmp{\tau_i\ocomp(p\seq\udot)}}^{-1}}&
h(\sigma, h(\tau_1\ocomp(p_1^\bullet), a_{1\bullet}^\bullet), \dots,
h(\tau_n\ocomp(p_n^\bullet), a_{n\bullet}^\bullet)) \\
&\toletter{h(\sigma, h'(\tau_\bullet, f\seq\udot))}
&h(\sigma, h(\tau_1\ocomp(q_1^\bullet), b_{1\bullet}^\bullet), \dots,
h(\tau_n\ocomp(q_n^\bullet), b_{n\bullet}^\bullet))\\
&\toletter{\delta_{\sigma\cmp{\tau_i\ocomp(p\seq\udot)}}}
&h(\sigma\ocomp(\tau_\bullet)\ocomp(q\seq\udot), b\seq\udot),
\end{array}
\]
where each $h'(\tau_i, f_i^\bullet)$ is
\[
\begin{array}{rcl}
h(\tau_i\ocomp(p_i^\bullet),a_i^\bullet))
&\toletter{\delta_{\tau_i\cmp{p_i^\bullet}}^{-1}}
& h(\tau_i, h(p_i^1, a_{i\bullet}^1), \dots, h(p_i^{k_i}, a_{i\bullet}^{k_i}))\\
&\toletter{h(\tau_i, f_i^\bullet)}
& h(\tau_i, h(q_i^1, b_{i\bullet}^1), \dots, h(q_i^{k_i}, b_{i\bullet}^{k_i}))\\
&\toletter{\delta_{\tau_i\cmp{p_i^\bullet}}}
&h(\tau_i\ocomp(q_i^\bullet),b_i^\bullet).
\end{array}
\]
So the equation holds if the following diagram commutes:
\[
\xymatrix @+10pt {
& h(\sigma\ocomp(\tau_\bullet)\ocomp(p\seq\udot), a\seq^\bullet)
\ar[dl]_{\delta_{\sigma\ocomp(\tau_i)\cmp{p\seq\udot}}^{-1}}
\ar[dr]^{\delta_{\sigma\cmp{\tau_i\ocomp(p\seq\udot)}}^{-1}}
\ar[d]|{\delta_{\sigma\cmp{\tau_\bullet}\cmp{p\seq\udot}}^{-1}} \\
h(\sigma\ocomp(\tau_\bullet), h(p\seq\udot, a\seq\udot))
\ar[d]|{h(\sigma\ocomp(\tau_\bullet), f\seq\udot)}
& h(\sigma, h(\tau_\bullet\ocomp(p\seq\udot),a\seq\udot))
\ar[l]^{\delta_{\sigma\cmp{\tau_\bullet}}}
\ar[d]|{h(\sigma, h(\tau_i, f_i^\bullet))}
\ar @{} [dl]|*+[o][F-]{1}
\ar @{} [dr]|*+[o][F-]{2}
& h(\sigma\ocomp(\tau_\bullet), h(p\seq\udot, a\seq\udot))
\ar[l]^{h(\sigma, \delta_{\tau_\bullet\cmp{p\seq\udot}}^{-1})}
\ar[d]|{h(\sigma, h'(\tau_\bullet, f\seq\udot))}\\
h(\sigma\ocomp(\tau_\bullet), h(q\seq\udot, b\seq\udot))
\ar[dr]_{\delta_{\sigma\ocomp(\tau_\bullet)\cmp{q\seq\udot}}}
& h(\sigma, h(\tau_\bullet\ocomp(q\seq\udot), b\seq\udot))
\ar[d]|{\delta_{\sigma\cmp{\tau_\bullet\ocomp(q\seq\udot)}}}
\ar[l]^{\delta_{\sigma\cmp{\tau_\bullet}}}
\ar[r]_{h(\sigma, \delta_{\tau_i\cmp{p_i^\bullet}})}
& h(\sigma, h(\tau_\bullet\ocomp(q\seq\udot), b\seq\udot))
\ar[dl]^{\delta_{\sigma\cmp{\tau_\bullet\ocomp(q\seq\udot)}}}\\
&h(\sigma\ocomp(\tau_\bullet)\ocomp(q\seq\udot), b\seq\udot)\\
}
\]
The triangles all commute because all $\delta$s are images of arrows in $Q$,
and there is at most one 2-cell between any two 1-cells in $Q$. \xylabel 2
commutes by the definition of $h'(\tau_i, f_i^\bullet)$, and \xylabel 1
commutes by naturality of $\delta$.
\end{proof}
\begin{lemma}
\label{lem:equiv<=>Q-equiv}
Let $Q\kel A \toletter{h} A$ and $Q \kel B \toletter{h'} B$
be weak $P$-categories, $(F, \pi):A \to B$ be a weak $P$-functor, and $(F, G,
\eta, \epsilon)$ be an adjoint equivalence in \cat{Cat}.
Then $G$ naturally carries the structure of a weak $P$-functor, and $(F, G,
\eta, \epsilon)$ is an adjoint equivalence in \cat{Wk-$P$-Cat}.
\end{lemma}
\begin{proof}
We want a sequence $(\psi_\bullet)$ of natural transformations:
\[
\xymatrixrowsep{3pc}
\xymatrixcolsep{3pc}
\xymatrix{
Q_i\times B^i
\ar[d]_{1\times G^i}
\ar[r]^{h'_i}
\drtwocell\omit{^*{!(-1,-1.5)\object{\psi_i}}}
& B \ar[d]^G \\
Q_i\times A^i \ar[r]_{h_i}
& A \\
}
\]
Let $\psi_i$ be given by
\def\epsiloncell#1{
\ultwocell\omit{<-0.7>*{!(-2,-1.2){1\times \epsilon^{#1}}}}
}
\def\epsiloncell{\sum k_i}{\epsiloncell{\sum k_i}}
\def\epsiloncell{n}{\epsiloncell{n}}
\[
\xymatrixrowsep{3.5pc}
\xymatrixcolsep{3pc}
\xymatrix{
Q_i \times B^i
\ar[d]_{1\times G^i}
\ar[r]^{h'_i}
\drtwocell\omit{^*{!(-1,-1.5)\object{\psi_i}}}
& B
\ar[d]^G \\
Q_i \times A^i
\ar[r]_{h_i}
& A \\
}
\midlabel{=}
\xymatrixrowsep{1.5pc}
\xymatrixcolsep{1.2pc}
\xymatrix{
Q_i \times B^i
\ar[rr]^1
\ar[dd]_{1\times G^i}
&& Q_i \times B^i
\ar[rr]^{h'_i}
&& B
\ar[dd]^G
\\
& {}
\ultwocell\omit{<-0.7>*{!(-2,-1.8){1\times \epsilon^{i}}}}
&& {}
\lltwocell\omit{*{!(-5,-4)\object{{\pi^{-1}_i}}}}
\\
Q_i \times A^i
\ar[urur]|{1 \times F^i}
\ar[rr]_{h_i}
&& A
\ar[urur]^F
\ar[rr]_1
&& A
\ultwocell\omit{{\eta}} \\
}
\]
We must check that $\psi$ satisfies (\ref{weakfunctordef}) and
(\ref{weakfunctordef2}) from Lemma \ref{lem:wkfunc_explicit}.
For (\ref{weakfunctordef}):
\begin{eqnarray*}
\midlabel{\mbox{LHS}} & \midlabel{=}
& \xymatrixrowsep{5pc}
\xymatrixcolsep{5pc}
\xymatrix{
{}
\ar[d]_{1\times G^{\sum k_i}}
\ar[r]^{\prodkn{h'}}
\drtwocell\omit{^*{!(-1,-1.5)\object{\prodkn{\psi}}}}
& {}
\ar[d]|{1\times G^n}
\ar[r]^{h'_n}
\drtwocell\omit{^*{!(-1,-1.5)\object{\psi_n}}}
& {}
\ar[d]^G
\\
{}
\ar[r]_{\prodkn{h}}
& {}
\ar[r]_h
& {} \\
} \\
& \midlabel{=}
& \xymatrixrowsep{2.3pc}
\xymatrixcolsep{2.3pc}
\xymatrix{
{}
\ar[rr]^1
\ar[dd]_{1\times G^{\sum k_i}}
&& {}
\ar[rr]^{\prodkn{h'}}
& {}
& {}
\ar[rr]^1
\ar[dd]|{1\times G^n}
&& {}
\ar[rr]^{h'_n}
& {}
& {}
\ar[dd]^G
\\
& {}
\epsiloncell{\sum k_i}
&& {}
\lltwocell\omit{*{!(-5,-4)\object{\prodkn{\pi^{-1}}}}}
&& {}
\epsiloncell{n}
&& {}
\lltwocell\omit{*{!(-5,-4)\object{\pi_n^{-1}}}}
&& \\
{}
\ar[uurr]|{1 \times F^{\sum k_i}}
\ar[rr]_{\prodkn{h}}
&& {}
\ar[uurr]|{1 \times F^n}
\ar[rr]_1
&& {}
\ar[uurr]|{1 \times F^n}
\ar[rr]_h
\ultwocell\omit{*{!(1,0)\object{1 \times \eta^n}}}
&& {}
\ar[uurr]|F
\ar[rr]_1
&& {}
\ultwocell\omit{{\eta}} \\
} \\
& \midlabel{=}
& \xymatrixrowsep{2.3pc}
\xymatrixcolsep{2.3pc}
\xymatrix{
{}
\ar[rr]^1
\ar[dd]_{1\times G^{\sum k_i}}
&& {}
\ar[rr]^{\prodkn{h'}}
& {}
& {}
\ar[rr]^1
&& {}
\ar[rr]^{h'_n}
& {}
& {}
\ar[dd]^G
\\
& {}
\epsiloncell{\sum k_i}
&& {}
\lltwocell\omit{*{!(-5,-4)\object{\prodkn{\pi^{-1}}}}}
&&&& {}
\lltwocell\omit{*{!(-5,-4)\object{\pi_n^{-1}}}}
& \\
{}
\ar[urur]|{1 \times F^{\sum k_i}}
\ar[rr]_{\prodkn{h}}
&& {}
\ar[urur]|{1 \times F^n}
\ar[rr]_1
&& {}
\ar[urur]|{1 \times F^n}
\ar[rr]_h
\uutwocell\omit{=}
&& {}
\ar[urur]|F
\ar[rr]_1
&& {}
\ultwocell\omit{{\eta}} \\
} \\
& \midlabel{=}
& \xymatrixrowsep{2.3pc}
\xymatrixcolsep{2.3pc}
\xymatrix{
{}
\ar[rr]^1
\ar[dd]_{1\times G^{\sum k_i}}
&& {}
\ar[rr]^{\prodkn{h'}}
& {}
& {}
\ar[rr]^{h'_n}
& {}
& {}
\ar[dd]^G
\\
& {}
\epsiloncell{\sum k_i}
&& {}
\lltwocell\omit{*{!(-5,-4)\object{\prodkn{\pi^{-1}}}}}
&& {}
\lltwocell\omit{*{!(-5,-4)\object{\pi_n^{-1}}}}
&& \\
{}
\ar[urur]|{1 \times F^{\sum k_i}}
\ar[rr]_{\prodkn{h}}
&& {}
\ar[urur]|{1 \times F^n}
\ar[rr]_h
&& {}
\ar[urur]|F
\ar[rr]_1
&& {}
\ultwocell\omit{{\eta}} \\
} \\
& \midlabel{=}
& \xymatrixrowsep{2.3pc}
\xymatrixcolsep{2.3pc}
\xymatrix{
{}
\ar[rr]^1
\ar[dd]_{1\times G^{\sum k_i}}
&& {}
\ar[rr]^{h'_{\sum k_i}}
& {}
& {}
\ar[dd]^G
\\
& {}
\epsiloncell{\sum k_i}
&& {}
\lltwocell\omit{*{!(-5,-4)\object{\pi_{\sum k_i}^{-1}}}}
& \\
{}
\ar[urur]|{1 \times F^{\sum k_i}}
\ar[rr]_{h_{\sum k_i}}
&& {}
\ar[urur]|F
\ar[rr]_1
&& {}
\ultwocell\omit{{\eta}} \\
} \\
& \midlabel{=}
& \xymatrixrowsep{5pc}
\xymatrixcolsep{5pc}
\xymatrix{
{}
\ar[d]_{1\times G^{\sum k_i}}
\ar[r]^{h'_{\sum k_i}}
\drtwocell\omit{^*{!(-1,-1.5)\object{\psi_{\sum k_i}}}}
& {}
\ar[d]^G
\\
{}
\ar[r]_{h_{\sum k_i}}
& {} \\
} \\
& = & \mbox{RHS.}
\end{eqnarray*}
For (\ref{weakfunctordef2}), consider the following diagram:
\[
\xymatrix{
&&G b \ar[r]^{\delta_{1_Q}} \ar@/l3cm/[ddd]_{1} \ar[d]_{\eta G}
\ar @{} [dr]|*+[o][F-]{2}
& h(1_P, G b) \ar@/r3cm/[ddd]^{\psi_1} \ar[d]^{\eta} \\
\ar @{} [drr]|*+[o][F-]{1}
&&GFG b \ar[r]^{GF\delta_{1_Q}} \ar[d]_{1}
\ar @{} [dr]|*+[o][F-]{3}
& GFh(1_P, Gb) \ar[d]^{\pi^{-1}_1}
\ar @{} [drr]|*+[o][F-]{5}
\\
&&GFG b \ar[r]_{G\delta'_{1_Q}} \ar[d]_{G\epsilon}
\ar @{} [dr]|*+[o][F-]{4}
& G h'(1_P, FGb) \ar[d]^{Gh'(1_P, \epsilon)} &&\\
&&Gb \ar[r]_{G\delta'_{1_Q}}
& Gh'(1_P, b)
}
\]
(\ref{weakfunctordef2}) is the outside of the diagram.
\xylabel{1} commutes by the triangle identities.
\xylabel{2} commutes by naturality of $\eta$.
\xylabel{3} commutes since $(F, \pi)$ is a $P$-functor.
\xylabel{4} commutes by naturality of $\delta$.
\xylabel{5} is the definition of $\psi$.
Hence the whole diagram commutes, and $(G, \psi)$ is a $P$-functor.
To see that $(F, G, \eta, \epsilon)$ is a $P$-equivalence, it is now enough to
show that $\eta$ and $\epsilon$ are $P$-transformations, since they satisfy the
triangle identities by hypothesis.
Write $(GF, \chi) = (G, \psi) \fcomp (F, \pi)$.
We wish to show that $\eta$ is a $P$-transformation $(1,1) \to (GF, \chi)$.
Each $\chi_{q, a\seq}$ is the composite
\[
\xymatrix{
h(q, GFa\seq) \ar[r]^{\psi_{q, Fa\seq}}
& G h(q, Fa\seq) \ar[r]^{G\pi_{q,a\seq}}
& GFh(q, a\seq)
}
\]
Applying the definition of $\psi$, this is
\[
\centerline{
\xymatrix{
h(q, GFa\seq) \ar[r]^\eta
& GF h(q, GFa\seq) \ar[r]^{G\pi^{-1}}
& G h(q, FGFa\seq) \ar[r]^{Gh_q \epsilon F}
& Gh(q, F a\seq) \ar[r]^{G\pi}
& GFh(q, a\seq)
}
}
\]
The axiom on $\eta$ is the outside of the diagram
\[
\centerline{
\xymatrix{
h(q, a\seq) \ar[rrrr]^1 \ar[dd]_{h(q, \eta)} \ar[dr]^\eta
&&&& h(q, a\seq) \ar[dd]^\eta \\
& GFh(q, a\seq) \ar[r]^{G\pi^{-1}} \ar[d]|{GFh(q, \eta)}
\ar @{} [l]|*+[o][F-]{1}
\ar @{} [dr]|*+[o][F-]{2}
& Gh(q, Fa\seq) \ar[d]|{Gh(q, F\eta)} \ar[dr]^1
\ar @{} [rr]|*+[o][F-]{3} & & \\
h(q, GFa\seq) \ar[r]^\eta
& GFh(q, GFa\seq) \ar[r]^{G\pi^{-1}}
& Gh(q, FGFa\seq) \ar[r]^{Gh(q, \epsilon F)}
& Gh(q, Fa\seq) \ar[r]^{G\pi}
& GFh(q, a\seq)
}
}
\]
\xylabel{1} commutes by naturality of $\eta$, \xylabel{2} commutes by
naturality of $\pi^{-1}$, and \xylabel{3} commutes since $G\pi \fcomp G\pi^{-1}
= G(\pi \fcomp \pi^{-1}) = G1 = 1G$. The triangle commutes by the triangle
identities. So the whole diagram commutes, and $\eta$ is a $P$-transformation.
By Lemma \ref{inv <=> P-inv}, $\eta^{-1}$ is also a $P$-transformation.
Similarly, $\epsilon$ and $\epsilon^{-1}$ are $P$-transformations.
\end{proof}
The statement of the lemma is a fragment of the statement that \cat{Wk-$P$-Cat}\ is
2-monadic over \cat{Cat}.
Compare the fact that monadic functors reflect isos.
\begin{theorem}
\label{mainthm}
\index{coherence!for weak $P$-categories}
Let $Q\kel A \toletter{h} A$ be a weak $P$-category. Then $A$
is equivalent to $\st A$ in the 2-category \cat{Wk-$P$-Cat}.
\end{theorem}
\begin{proof}
Let $F: \st A \to A$ be given by $F(p, a\seq) = h(p, a\seq)$ and
identification of maps. This is certainly full and faithful, and it is
essentially surjective on objects because $\delta^{-1}_{1_Q}: h(1_P, a)
\to a$ is an isomorphism. By Lemma \ref{lem:equiv<=>Q-equiv}, it remains only
to show that $F$ is a weak $P$-functor.
We must find a sequence $(\phi_i : h_i (1 \times F^i) \to Fh')$ of natural
transformations satisfying equations (\ref{weakfunctordef}) and
(\ref{weakfunctordef2}) from Lemma \ref{lem:wkfunc_explicit}.
We can take $(\phi_i)_{q, (p_\bullet, a\seq\udot)} = (\delta_{q \cmp{p_\bullet}})_{a\seq\udot}$
for $q \in Q_n$ and $(p_1, a^1_\bullet), \dots, (p_n, a^n_\bullet) \in \st A$.
For (\ref{weakfunctordef}), we must show that
\[
\xymatrixrowsep{4pc}
\xymatrixcolsep{4pc}
\xymatrix{
{}
\ar[d]_{1\times F^{\sum k_i}}
\ar[r]^{\prodkn{h'}}
\drtwocell\omit{^*{!(-1,-1.5)\object{\prodkn{\phi}}}}
& {}
\ar[d]|{1\times F^n}
\ar[r]^{h'_n}
\drtwocell\omit{^*{!(-1,-1.5)\object{\phi_n}}}
& {}
\ar[d]^F
\\
{}
\ar[r]_{\prodkn{h}}
& {}
\ar[r]_h
& {} \\
}
\midlabel{=}
\xymatrixrowsep{4pc}
\xymatrixcolsep{4pc}
\xymatrix{
{}
\ar[d]_{1\times F^{\sum k_i}}
\ar[r]^{h'_{\sum k_i}}
\drtwocell\omit{^*{!(-1,-1.5)\object{\phi_{\sum k_i}}}}
& {}
\ar[d]^F
\\
{}
\ar[r]_{h_{\sum k_i}}
& {} \\
}
\]
All 2-cells in this equation are instances of $\delta$. Since there is at most
one 2-cell between two 1-cells in $Q$, the equation holds.
For (\ref{weakfunctordef2}) to hold, we must have
\begin{equation}
\label{Fwkfctr}
\xymatrix{
F(p,a\seq) \ar[d]_1 \ar[r]^{\delta_{1_Q}}
& h(1_P, F(p,a\seq)) \ar[d]^{\phi_{1_P}} \\
F(p,a\seq) \ar[r]_{F\delta'_{1_Q}}
& Fh'(1_P,(p, a\seq))
}
\end{equation}
Since \st A is a strict monoidal category, $\delta' = 1$.
Apply this observation, and the definitions of $F$, $\phi$ and $h'$; then
(\ref{Fwkfctr}) becomes
\[
\xymatrix{
h(p,a\seq) \ar[d]_1 \ar[r]^{\delta_{1_Q}}
& h(1_P, h(p,a\seq)) \ar[d]^{\delta_{1_P\cmp{p}}} \\
h(p,a\seq) \ar[r]_1
& h(p,a\seq)
}
\]
Since there is at most one arrow between two 1-cells in $Q$, this diagram
commutes.
So $(F,\phi)$ is a weak $P$-functor, and hence (by Lemma
\ref{lem:equiv<=>Q-equiv}) an equivalence in \cat{Wk-$P$-Cat}.
\end{proof}
\begin{example}
\index{sets}
Consider the initial operad $0$, whose algebras are sets.
We saw in Example \ref{ex:trivial_theory} that unbiased weak $0$-categories
are categories equipped with a trivial monad.
By Theorem \ref{mainthm}, every unbiased weak $0$-category is equivalent via
weak $0$-functors to a category equipped with a monad which \emph{is} the
identity: in other words, a category.
\end{example}
\begin{example}
\index{monoids}
Consider the terminal operad $1$, whose algebras are monoids.
Theorem \ref{mainthm} tells us that every unbiased weak monoidal category is
monoidally equivalent to a strict monoidal category.
\end{example}
\section{Universal property of {\rm\textbf{st} }}
\label{stsect}
\index{strictification!universal property}
Let $P$ be a plain operad.
\begin{theorem}
Let $U'$ be the forgetful functor $\cat{Str-$P$-Cat} \to \cat{Wk-$P$-Cat}$ (considering both of
these as 1-categories). Then {\rm\textbf{st} } is left adjoint to $U'$.
\end{theorem}
\begin{proof}
For each $(A,h) \in \cat{Wk-$P$-Cat}$, we construct an
initial object $A \toletter{(F', \psi)} \st A$ of the comma category
$(A\downarrow U')$, thus showing that {\rm\textbf{st} } is functorial and that ${\rm\textbf{st} }
\dashv U'$ (and that $(F', \psi)$ is the component of the unit at $A$). Let
$(B, h'')$ be a strict $P$-category, and
$(G, \gamma): A \to U'B$ be a weak $P$-functor. We must show that there is a
unique strict $P$-functor $H$ making the following diagram commute:
\begin{equation}
\label{A->U}
\xymatrix{
& A \ar[ddl]_{(F', \psi)} \ar[ddr]^{(G, \gamma)} \\ \\
U'\,\st A \ar@{-->}[rr]^{(H, {\rm id})} & & U'B
}
\end{equation}
$(F', \psi)$ is given as follows:
\begin{itemize}
\item If $a \in A$, then ${F'}(a) = (1, a)$.
\item If $f: a \to a'$ in $A$ then ${F'}f$ is the lifting of $h(1,f)$ with
source $(1, a)$ and target $(1, a')$.
\item Each $\psi_{(p, a\seq)}$ is the lifting of $(\delta_{1_Q})_{h(p, a\seq)}:
h(p, a\seq) \to h(1, h(p, a\seq))$ to a morphism
$h'(p, F'(a)_\bullet) = (p, a\seq) \to (1, h(p, a\seq)) = F'(h(p, a\seq))$.
\end{itemize}
For commutativity of (\ref{A->U}), we must have $H(1, a) = G(a)$, and
for strictness of $H$, we must have $H(p, a\seq) = h''(p, H(1, a)_\bullet)$. These
two conditions completely determine $H$ on objects.
Now, take a morphism $f:(p, a\seq) \to (p', a\seq')$, which is a lifting of a
morphism $g: h(p,a\seq) \to h(p',a\seq')$ in $A$.
Then $Hf$ is a morphism $h''(p, Ga\seq) \to h''(p', Ga\seq')$: the obvious
thing for it to be is the composite
\[
\xymatrix{
h''(p, Ga\seq) \ar[r]^\gamma
& G h''(p, a\seq) \ar[r]^{Gg}
& G h''(p', a\seq') \ar[r]^{\gamma^{-1}}
& h''(p', Ga\seq')
}
\]
and we shall show that this is in fact the only possibility.
Consider the composite
\[
\xymatrix{
(1, h(p, a\seq)) \ar[r]^-{\psi^{-1}}
&(p, a\seq) \ar[r]^{f}
& (p', a\seq) \ar[r]^-{\psi}
& (1, h(p', a\seq'))
}
\]
in \st A. Composition in \st A is given by composition in $A$, so this is
equal to the lifting of $\delta_{1_Q} \fcomp g \fcomp \delta_{1_Q}^{-1} = h(1,
g)$ to a morphism $(1, h(p, a\seq)) \to (1, h(p',
a\seq'))$, namely $F'g$. So $f = \psi^{-1} \fcomp F'g \fcomp \psi$, and
$Hf = H\psi^{-1} \fcomp HF'g \fcomp H\psi$. By commutativity of
(\ref{A->U}), $HF' = G$ and $H\psi = \gamma$, so $Hf = \gamma^{-1} \fcomp
Gg \fcomp \gamma$ as required.
This completely defines $H$. So we have constructed a unique
$H$ which makes (\ref{A->U}) commute and which is strict. Hence $(F',
\psi): A \to U'\,\st A$ is initial in $(A\downarrow U')$, and so ${\rm\textbf{st} }
\dashv U'$.
\end{proof}
The $P$-functor $(F, \phi):\st A \to A$ constructed in Theorem \ref{mainthm} is
pseudo-inverse to $(F', \psi)$, which we have just shown to be the
$A$-component of the unit of the adjunction {\rm\textbf{st} } $\dashv U'$. We can
therefore say that \cat{Str-$P$-Cat}\ is a weakly coreflective sub-2-category of
\cat{Wk-$P$-Cat}. Note that the counit is \emph{not} pseudo-invertible, so this is not
a 2-equivalence.
\begin{example}
\index{sets}
Consider again the initial operad $0$, whose algebras are sets.
We saw in Example \ref{ex:trivial_theory} that unbiased weak $0$-categories
are categories equipped with a specified trivial monad.
Let $\cat{Triv}$ denote the category of such categories, with morphisms being
functors that preserve the trivial monad up to coherent isomorphism.
A strict unbiased $0$-category is a category equipped with a monad equal to
the identity monad, which is simply a category.
So $\cat{Cat}$ is a weakly coreflective sub-2-category of $\cat{Triv}$.
\end{example}
\section{Presentation-independence}
\label{sec:presindep}
We will now show that the weakening of a symmetric operad $P$ is essentially
independent of the generators chosen.
This generalizes Leinster's result (in \cite{hohc} section 3.2) that the
theory of weak monoidal categories is essentially unaffected by the choice of a
different presentation for the theory of monoids.
We will need the following lemma:
\begin{lemma}
\label{lem:fork}
In \cat{Cat-$\Sigma$-Operad}, if $\fork P \alpha \beta Q \gamma R$ is a fork, and $\gamma$
is levelwise full and faithful, then $\alpha \cong \beta$.
\end{lemma}
\begin{proof}
We shall construct an invertible \cat{Cat}-$\Sigma$-operad transformation $\eta:\alpha \to
\beta$.
We form the $\eta_n$s as follows: for all $p \in P_n$, let $\gamma\alpha(p) =
\gamma\beta(p)$.
Since $\gamma$ is levelwise full, there exists an arrow $(\eta_n)_p: \alpha(p)
\to \beta(p)$ such that $\gamma_n((\eta_n)_p) = 1_{\gamma\alpha(p)}$.
Since $\gamma$ is levelwise full and faithful, this arrow is an isomorphism.
Each $\eta_n$ is natural because, for all $n \in \natural$ and $f: p \to q$ in
$P_n$, the image under $\gamma$ of the naturality square
\[
\xymatrix{
\alpha(p) \ar[r]^{(\eta_n)_p} \ar[d]_{\alpha(f)}
& \beta(p) \ar[d]^{\beta(f)} \\
\alpha(q) \ar[r]^{(\eta_n)_q}
& \beta(q)
}
\]
is
\[
\xymatrix{
\gamma\alpha(p) \ar[r]^1 \ar[d]_{\gamma\alpha(f)}
& \gamma\beta(p) \ar[d]^{\gamma\beta(f)} \\
\gamma\alpha(q) \ar[r]^1
& \gamma\beta(q)
}
\]
which commutes since $\gamma\alpha = \gamma\beta$.
Since $\gamma$ is faithful, the naturality square commutes, and $\eta_n$ is
natural.
It remains to show that the collection $(\eta_n)_{n \in \natural}$ forms a
\cat{Cat}-$\Sigma$-operad transformation, in other words that the equations
{\def\nxseq#1{{#1}_n \times {#1}_\bullet}
\def_{\sum{k_i}}{_{\sum{k_i}}}
\begin{eqnarray}
\xymatrixrowsep{5pc}
\xymatrixcolsep{5pc}
\xymatrix{
\nxseq P \rtwocell^{\nxseq{\alpha}}_{\nxseq{\beta}}
{*{!(-3.5,0)\object{\nxseq{\eta}}}}
\ar[d]_\ocomp
\drtwocell\omit{=}
& \nxseq Q \ar[d]^\ocomp \\
P_{\sum{k_i}} \ar[r]_{\beta_{\sum{k_i}}} & Q_{\sum{k_i}}
}
&\midlabel{=}
&\xymatrixrowsep{5pc}
\xymatrixcolsep{5pc}
\xymatrix{
\nxseq P \ar[r]^{\nxseq{\alpha}}
\ar[d]_\ocomp
\drtwocell\omit{=}
& \nxseq Q \ar[d]^\ocomp \\
P_{\sum{k_i}} \rtwocell^{\alpha_{\sum{k_i}}}_{\beta_{\sum{k_i}}}
{*{!(-3.5,0)\object{\eta_{\sum{k_i}}}}}
& Q_{\sum{k_i}}
} \\
(\eta_1)_{1} &= &1 \\
\xymatrixrowsep{4pc}
\xymatrixcolsep{4pc}
\xymatrix{
P_n \rtwocell^{\alpha_n}_{\beta_n}{*{!(-1,0)\object{\eta_n}}}
\ar[d]_{\sigma\cdot-}
\drtwocell\omit{=}
& Q_n \ar[d]^{\sigma\cdot-} \\
P_n \ar[r]_{\beta_n} & Q_n
}
&\midlabel{=}
&\xymatrixrowsep{4pc}
\xymatrixcolsep{4pc}
\xymatrix{
P_n \ar[r]^{\alpha_n}
\ar[d]_{\sigma\cdot-}
\drtwocell\omit{=}
& Q_n \ar[d]^{\sigma\cdot-} \\
P_n \rtwocell^{\alpha_n}_{\beta_n}{*{!(-1,0)\object{\eta_n}}} & Q_n
}
\end{eqnarray}}
hold, for all $n, k_1 \dots k_n \in \natural$ and every $\sigma \in S_n$.
As above, it is enough to show that the images of both sides under $\gamma$
are equal, and this is trivially true by definition of $\eta$.
\end{proof}
Let $P$ be a symmetric operad.
\begin{theorem}
\label{thm:pres_indep}
\index{presentation!independence of weakenings}
Let $\Phi \in \cat{Set}^\natural$ and let $\phi : {\ensuremath{F_{\Sigma}}} \Phi \to P$ be a regular epi.
Then $\Wkwrt P \phi$ is equivalent as a symmetric \cat{Cat}-operad to \Wk{P}.
\end{theorem}
\begin{proof}
Let $Q$ be the weakening of $P$ with respect to $\phi: {\ensuremath{F_{\Sigma}}} \Phi \to P$.
By the triangle identities, we have a commutative square
\[
\xymatrix{
{\ensuremath{F_{\Sigma}}} \Phi \ar[r]^\phi \ar[d]_{{\ensuremath{F_{\Sigma}}} \overline\phi}
& P \ar[d]^1 \\
{\ensuremath{F_{\Sigma}}} {\ensuremath{U^{\Sigma}}} P \ar[r]^{\epsilon_P}
& P
}
\]
By functoriality of the factorization system, this gives rise to a unique map
$\chi: Q \to \Wk{P}$ such that
\[
\xymatrix{
{\ensuremath{F_{\Sigma}}} \Phi \booar[r] \ar@/u0.7cm/[rr]^\phi \ar[d]_{{\ensuremath{F_{\Sigma}}} \overline\phi}
& Q \lffar[r] \unar[d]^\chi
& P \ar[d]^1 \\
{\ensuremath{F_{\Sigma}}} {\ensuremath{U^{\Sigma}}} P \booar[r] \ar@/d0.7cm/[rr]_{\epsilon_P}
& \Wk{P} \lffar[r]
& P
}
\]
commutes.
We wish to find a pseudo-inverse to $\chi$.
Since \cat{$\Sigma$-Operad}\ is monadic over $\cat{Set}^\natural$, a regular epi in
\cat{$\Sigma$-Operad}\ is a levelwise surjection by Theorem
\ref{thm:reg epis split in Set^X^T}.
So we may choose a section $\psi_n$ of $\phi_n : ({\ensuremath{F_{\Sigma}}} \Phi)_n \to P_n$ for
all $n \in \natural$.
So we have a morphism $\psi : {\ensuremath{U^{\Sigma}}} P \to {\ensuremath{U^{\Sigma}}} {\ensuremath{F_{\Sigma}}} \Phi$ in $\cat{Set}^\natural$.
We wish to show that
\[
\xymatrix{
{\ensuremath{F_{\Sigma}}} {\ensuremath{U^{\Sigma}}} P \ar[r]^{\epsilon_P} \ar[d]_{\overline\psi}
& P \ar[d]^1 \\
{\ensuremath{F_{\Sigma}}} \Phi \ar[r]^\phi
& P
}
\]
commutes.
This follows from a simple transpose argument:
\[
\newdir{((}{{}*!/-5pt/{}}
\xymatrixrowsep{0.15pc}
\xymatrix{
& {\ensuremath{F_{\Sigma}}} {\ensuremath{U^{\Sigma}}} P \ar[r]^{\bar \psi} & {\ensuremath{F_{\Sigma}}} \Phi \ar[r]^\phi & P \\
\ar@{((-((}[rrrr] & & & & \\
& {\ensuremath{U^{\Sigma}}} P \ar[r]^\psi & {\ensuremath{U^{\Sigma}}} {\ensuremath{F_{\Sigma}}} \Phi \ar[r]^{{\ensuremath{U^{\Sigma}}} \phi} & {\ensuremath{U^{\Sigma}}} P & =
& {\ensuremath{U^{\Sigma}}} P \ar[r]^1 & {\ensuremath{U^{\Sigma}}} P \\
& & & & \ar@{-}[rrr] & & & \\
& & & & & {\ensuremath{F_{\Sigma}}} {\ensuremath{U^{\Sigma}}} P \ar[r]^\epsilon & P.
}
\]
This induces a map
\[
\xymatrix{
{\ensuremath{F_{\Sigma}}} {\ensuremath{U^{\Sigma}}} P \booar[r] \ar@/u0.7cm/[rr]^{\epsilon_P}
\ar[d]_{\overline\psi}
& \Wk{P} \lffar[r] \unar[d]^\omega
& P \ar[d]^1 \\
{\ensuremath{F_{\Sigma}}} \Phi \booar[r] \ar@/d0.7cm/[rr]_\phi
& Q \lffar[r]
& P
}
\]
We will show that $\omega$ is pseudo-inverse to $\chi$.
Now,
\[
\xymatrix{
Q \lffar[r] \ar[d]_\omega
& P \ar[d]^1 \\
\Wk{P} \lffar[r] \ar[d]_\chi
& P \ar[d]^1 \\
Q \lffar[r]
& P
}
\]
commutes.
So $\xymatrix{Q\parallelars{1_Q}{\chi\omega} & Q \lffar[r] & P}$ is a fork.
By Lemma \ref{lem:fork}, $\chi\omega \cong 1_Q$, and similarly $\omega\chi
\cong 1_{\Wk{P}}$.
So $Q \simeq \Wk{P}$ as a symmetric \cat{Cat}-operad, as required.
\end{proof}
\begin{corollary}
\label{cor:unbiased_symm}
\index{unbiased!weakening of a plain operad!comparison to symmetric weakening}
Let $P$ be a plain operad.
Then ${\ensuremath{F_{\Sigma}^{\rm pl}}} (\Wk{P}) \simeq \Wk{{\ensuremath{F_{\Sigma}^{\rm pl}}} P}$.
\end{corollary}
\begin{proof}
Let $\phi: {\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} P \to P$ be the component at $P$ of the counit of the
adjunction ${\ensuremath{F_{\rm pl}}} \dashv {\ensuremath{U^{\rm pl}}}$.
Let $\epsilon$ be the counit of the adjunction ${\ensuremath{F_{\Sigma}^{\rm pl}}} \dashv {\ensuremath{U^{\Sigma}_{\rm pl}}}$.
By Theorem \ref{thm:plain_as_symmetric}, there is an isomorphism ${\ensuremath{F_{\Sigma}^{\rm pl}}}
(\Wkwrt P \phi) \cong \Wkwrt{{\ensuremath{F_{\Sigma}^{\rm pl}}} P}{{\ensuremath{F_{\Sigma}^{\rm pl}}} \phi}$, and by Theorem
\ref{thm:pres_indep}, there is an equivalence $\Wkwrt{{\ensuremath{F_{\Sigma}^{\rm pl}}} P}{{\ensuremath{F_{\Sigma}^{\rm pl}}} \phi}
\simeq \Wk{{\ensuremath{F_{\Sigma}^{\rm pl}}} P}$.
Hence ${\ensuremath{F_{\Sigma}^{\rm pl}}} (\Wk P) \simeq \Wk{{\ensuremath{F_{\Sigma}^{\rm pl}}} P}$.
\end{proof}
\begin{corollary}
Let $P$ be a plain operad.
Then the category \cat{Wk-$P$-Cat}\ is equivalent to the category
\cat{Wk-${\ensuremath{F_{\Sigma}^{\rm pl}}} P$-Cat}.
\end{corollary}
This tells us that the unbiased categorification of a strongly regular theory
is essentially unaffected by our treating it as a linear theory instead.
\begin{example}
\index{sets}
Considering again the trivial theory $0$, we see that $\cat{Triv} \simeq \cat{Cat}$.
\end{example}
This can be generalised to the multi-sorted situation:
\begin{lemma}
Let $X$ be a set, and $f$ be a regular epi in the category $\cat{Cat-Multicat}_X$ or
in the category $\cat{Cat-$\Sigma$-Multicat}_X$.
Then $f$ is locally surjective on objects.
\end{lemma}
\begin{proof}
$\cat{Multicat}_X$ is monadic over $\cat{Multigraph}_X$ by Lemma
\ref{lem:frees_exist} and Theorem \ref{lem:monadj}, and an object of
$\cat{Multigraph}_X$ can be considered as an object of $\cat{Set}^Y$, where $Y
= X \times X^*$, and $X^*$ is the free monoid on $X$: for each $x \in X$, and
each sequence $x_1, \dots, x_n \in X^*$, there is a set of funnels $x_1, \dots,
x_n \to x$.
Hence, by \ref{thm:reg epis split in Set^X^T}, every regular epi in
$\cat{Multicat}_X$ is locally surjective.
The objects functor $O: \cat{Cat} \to \cat{Set}$ has both a left adjoint $D$
and a right adjoint $I$.
Hence $O$ and $I$ preserve products, and hence by Lemma \ref{lem:frees_exist}
they induce an adjunction
\[
\adjunction{\cat{Cat-Multicat}_X}{\cat{Multicat}_X}{O_*}{I_*}.
\]
Since $O_*$ is a left adjoint, it preserves colimits, and in particular regular
epis: hence, every regular epi in $\cat{Cat-Multicat}_X$ must be locally surjective
on objects.
The symmetric case is proved analogously.
\end{proof}
\begin{theorem}
Let $M$ be a (symmetric) multicategory, and $\phi : {\ensuremath{F_{\rm pl}}}\Phi \to M$
(or in the symmetric case, $\phi : {\ensuremath{F_{\Sigma}}}\Phi \to M$) be a regular epi.
Then the weakening of $M$ with respect to $\phi$ is equivalent as a
\cat{Cat}-multicategory to $\Wk M$.
\end{theorem}
\begin{proof}
The proof is exactly as for Theorem \ref{thm:pres_indep}.
\end{proof}
\chapter{Conclusions}
\label{ch:conclusions}
\begin{quotation}
Epic fail.
-- John Walker
\end{quotation}
\chapter{Factorization Systems}
\label{ch:factsys}
The theory of factorization systems was introduced by Freyd and Kelly in
\cite{freyd+kelly} (though it was implicit in work of Isbell in the 1950s).
We shall use it in subsequent chapters to define the weakening of an algebraic
theory.
Here, we recall the basic definitions and some relevant theorems.
The material in this chapter is standard, and may be found in (for instance)
\cite{borceux} or \cite{acc}; for an alternative perspective and some more
historical background (as well as the interesting generalization to
\emph{weak} factorization systems), see \cite{k+t}.
\begin{defn}
\index{orthogonal}
Let $e: a \to b$ and $m: c \to d$ be arrows in a category \ensuremath{\mathcal{C}}.
We say that $e$ is \defterm{left orthogonal} to $m$, written $e \mathop{\bot} m$, if,
for all arrows $f: a \to c$ and $g: b \to d$ such that $mf = ge$, there exists
a unique map $t: b \to c$ such that the following diagram commutes:
\[
\xymatrix{
a \ar[r]^{\forall f} \ar[d]_e
& c \ar[d]^m \\
b \unar[ur]^{\exists! t} \ar[r]_{\forall g}
& d
}
\]
\end{defn}
\begin{defn} \label{def:FS}
\index{factorization system}
Let \ensuremath{\mathcal{C}}\ be a category.
A \defterm{factorization system} on \ensuremath{\mathcal{C}}\ is a pair $(\ensuremath{\mathcal{E}}, \ensuremath{\mathcal{M}})$ of classes of maps
in \ensuremath{\mathcal{C}}\ such that
\begin{enumerate}
\item \label{axm:factor} for all maps $f$ in \ensuremath{\mathcal{C}}, there exist $e \in \ensuremath{\mathcal{E}}$ and $m
\in \ensuremath{\mathcal{M}}$ such that $f = m e$;
\item \label{axm:iso_closure} \ensuremath{\mathcal{E}}\ and \ensuremath{\mathcal{M}}\ contain all identities, and are
closed under composition with isomorphisms on both sides;
\item \label{axm:orthogonal} $\ensuremath{\mathcal{E}} \mathop{\bot} \ensuremath{\mathcal{M}}$, i.e. $e \mathop{\bot} m$ for all $e \in
\ensuremath{\mathcal{E}}$ and $m \in \ensuremath{\mathcal{M}}$.
\end{enumerate}
\end{defn}
\begin{example}
\label{ex:epi-mono}
\index{factorization system!epi-mono}
Let $\ensuremath{\mathcal{C}} = \cat{Set}$, $\ensuremath{\mathcal{E}}$ be the epimorphisms, and $\ensuremath{\mathcal{M}}$ be the monomorphisms.
Then $(\ensuremath{\mathcal{E}},\ensuremath{\mathcal{M}})$ is a factorization system.
\end{example}
\begin{example} More generally, let $\ensuremath{\mathcal{C}}$ be some variety of algebras,
$\ensuremath{\mathcal{E}}$ be the regular epimorphisms (i.e., the surjections), and $\ensuremath{\mathcal{M}}$ be the
monomorphisms.
Then $(\ensuremath{\mathcal{E}},\ensuremath{\mathcal{M}})$ is a factorization system.
\end{example}
\begin{example}
\index{factorization system!bijective on objects/full and faithful!on \cat{Digraph}}
\label{ex:digraphFS}
Let $\ensuremath{\mathcal{C}} = \cat{Digraph}$, the category of directed graphs and graph morphisms.
Let $\ensuremath{\mathcal{E}}$ be the maps bijective on objects, and \ensuremath{\mathcal{M}}\ be the full and faithful
maps.
Then $(\ensuremath{\mathcal{E}},\ensuremath{\mathcal{M}})$ is a factorization system.
\end{example}
In deference to Example \ref{ex:epi-mono}, we shall use arrows like
$\xymatrix{{} \booar[r] & {}}$ to denote members of \ensuremath{\mathcal{E}}\ in commutative
diagrams, and arrows like $\xymatrix{{} \lffar[r] & {}}$ to denote members of
$\ensuremath{\mathcal{M}}$, for whatever values of \ensuremath{\mathcal{E}}\ and \ensuremath{\mathcal{M}}\ happen to be in force at the time.
We will use without proof the following standard properties of factorization
systems:
\begin{lemma}
\label{lem:fact_standard}
Let \ensuremath{\mathcal{C}}\ be a category, and $(\ensuremath{\mathcal{E}}, \ensuremath{\mathcal{M}})$ be a factorization system on \ensuremath{\mathcal{C}}.
\begin{enumerate}
\item $\ensuremath{\mathcal{E}} \cap \ensuremath{\mathcal{M}}$ is the class of isomorphisms in \ensuremath{\mathcal{C}}.
\item \label{fact_unique} The factorization in \ref{def:FS} (\ref{axm:factor})
is unique up to unique isomorphism.
\item The factorization in \ref{def:FS} (\ref{axm:factor}) is functorial, in
the following sense: if the square
\[
\commsquare A f B g h C {f'} D
\]
commutes, and $f = me, f' = m'e'$, then there is a unique morphism $i$ making
\[
\xymatrix{
A \ear[r]^{e} \ar[d]_g
& {} \mar[r]^{m} \unar[d]_i
& B \ar[d]^h \\
C \ear[r]^{e'}
& {} \mar[r]^{m'}
& D
}
\]
commute.
Thus, given a choice of $e \in \ensuremath{\mathcal{E}}$ and $m \in \ensuremath{\mathcal{M}}$ for each $f$ in \ensuremath{\mathcal{C}} (such that
$f = me$), we may construct functors $\ensuremath{\mathcal{E}}_*, \ensuremath{\mathcal{M}}_* : [\cat{2}, \ensuremath{\mathcal{C}}] \to [\cat{2},
\ensuremath{\mathcal{C}}]$:
\begin{eqnarray*}
& \ensuremath{\mathcal{E}}_* : f \mapsto e \\
& \ensuremath{\mathcal{E}}_* : (g,h) \mapsto (g,i) \\
& \ensuremath{\mathcal{M}}_* : f \mapsto m \\
& \ensuremath{\mathcal{M}}_* : (g,h) \mapsto (i,h).
\end{eqnarray*}
These functors are determined by \ensuremath{\mathcal{E}}\ and \ensuremath{\mathcal{M}}\ uniquely up to unique isomorphism.
\item \ensuremath{\mathcal{E}}\ and \ensuremath{\mathcal{M}}\ are closed under composition.
\item $\ensuremath{\mathcal{E}}^{\bot} = \ensuremath{\mathcal{M}}$ and ${}^{\bot} \ensuremath{\mathcal{M}} = \ensuremath{\mathcal{E}}$, where $\ensuremath{\mathcal{E}}^{\bot} = \{f
\mbox{ in } \ensuremath{\mathcal{C}} : e \mathop{\bot} f \mbox{ for all } e \in \ensuremath{\mathcal{E}}\}$ and $^{\bot}\ensuremath{\mathcal{M}} = \{f
\mbox{ in } \ensuremath{\mathcal{C}} : f \mathop{\bot} m \mbox{ for all } m \in \ensuremath{\mathcal{M}}\}$.
\end{enumerate}
\end{lemma}
Proofs of these statements may be found in \cite{acc} section 14.
We will also use the following fact:
\begin{lemma}
\label{lem:monad}
Let \ensuremath{\mathcal{C}}\ be a category with a factorization system $(\ensuremath{\mathcal{E}}, \ensuremath{\mathcal{M}})$.
Let $T$ be a monad on \ensuremath{\mathcal{C}}\ and let $\overline \E = \{ f \mathop{\mbox{in}}\ \ensuremath{\mathcal{C}} : U f \in \ensuremath{\mathcal{E}}\}$
and $\overline \M = \{f \mathop{\mbox{in}}\ \ensuremath{\mathcal{C}} : U f \in \ensuremath{\mathcal{M}}\}$, where $U$ is the forgetful functor
$\ensuremath{\mathcal{C}}^T \to \ensuremath{\mathcal{C}}$.
Then $(\overline \E, \overline \M)$ is a factorization system on $\ensuremath{\mathcal{C}}^T$ if $T$ preserves
\ensuremath{\mathcal{E}}-arrows.
\end{lemma}
\begin{proof}
This is established in \cite{acc}, Proposition 20.24: however, we shall provide
a proof for the reader's convenience.
We shall establish the axioms listed in Definition \ref{def:FS}.
\ref{axm:factor}.
Take an algebra map
\[
\algmap f T A a B b
\]
Applying axiom \ref{axm:factor} to the factorization system $(\ensuremath{\mathcal{E}}, \ensuremath{\mathcal{M}})$, we
obtain a decomposition $f = m e$, where $e: A \to I$ and $m: I \to B$.
We wish to lift this to a decomposition of $f$ as an algebra map.
In other words, we need a map $i: TI \to I$ making the diagram
\[
\xymatrix{
TA \ear[r]^{Te} \ar[d]_a
& TI \ar[r]^{Tm} \unar[d]_i
& TB \ar[d]^b \\
A \ear[r]^e
& I \mar[r]^m
& B
}
\]
commute, such that $(I,i)$ is a $T$-algebra.
Since $T$ preserves \ensuremath{\mathcal{E}}-arrows, $T e \mathop{\bot} m$, and we may obtain $i$ by
applying this orthogonality to the diagram \[
\xymatrixrowsep{3pc}
\xymatrix{
TA \ar[r]^{a} \ar[d]_{T e}
& A \ar[r]^e
& I \ar[d]^m \\
TI \ar[r]_{Tm} \unar[urr]^{\exists!i}
& TB \ar[r]_b
& B.
}
\]
It remains to show that $(I,i)$ is a $T$-algebra.
For the unit axiom, consider the diagram
\[
\xymatrix{
A \ear[r]^e \ar[d]^{\eta_A} \ar@/l1cm/[dd]_1
& I \mar[r]^m \ar[d]^{\eta_I} & B \ar[d]^{\eta_B} \ar@/r1cm/[dd]^1\\
TA \ear[r]^{Te} \ar[d]_a & TI \ar[r]^{Tm} \ar[d]^i & TB \ar[d]^b \\
A \ear[r]_e & E \mar[r]_m & B
}
\]
The top squares commute by naturality, and the outside triangles commute since $(A,a)$ and $(B,b)$ are $T$-algebras.
Hence the diagram
\[
\xymatrix{
& I \mar[dr]^m \unar[dd] \\
A \ear[ur]^e \ear[dr]_e & & B \\
& I \mar[ur]_m
}
\]
commutes if the dotted arrow is either $1_I$ or $i\eta_I$.
By orthogonality, $i\eta_I = 1_I$.
For the multiplication axiom, observe that the diagrams
\[
\xymatrixcolsep{3pc}
\xymatrixrowsep{3pc}
\xymatrix{
T^2 A \ear[r]^{T^2 e} \ar[d]_{\mu_A}
& T^2 I \ar[r]^{T^2 m} \ar[d]^{\mu_I}
& T^2 B \ar[d]^{\mu_B} \\
TA \ear[r]^{Te} \ar[d]_a & TI \ar[r]^{Tm} \ar[d]^i & TB \ar[d]^b \\
A \ear[r]_e & E \mar[r]_m & B
}
\hskip 1cm
\xymatrix{
T^2 A \ear[r]^{T^2 e} \ar[d]_{T a} \ar@/l1cm/[dd]_{a\mu_A}
& T^2 I \ar[r]^{T^2 m} \ar[d]^{T i}
& T^2 B \ar[d]^{T b} \ar@/r1cm/[dd]^{b\mu_B} \\
TA \ear[r]^{Te} \ar[d]_a & TI \ar[r]^{Tm} \ar[d]^i & TB \ar[d]^b \\
A \ear[r]_e & E \mar[r]_m & B
}
\]
both commute.
So the diagram
\[
\xymatrixcolsep{3pc}
\xymatrixrowsep{3pc}
\xymatrix{
T^2 A \ar[r]^{\mu_A} \ear[d]_{T^2 e}
& TA \ar[r]^a & A \ar[r]^e & I \mar[d]^m \\
T^2 I \ar[r]_{T^2 m} \unar[urrr]
& T^2 B \ar[r]_{\mu_B} & TB \ar[r]_b & B
}
\]
commutes if we take the dotted arrow to be either $i\mu_I$ or $i(Ti)$.
By orthogonality, $i\mu_I = i(Ti)$.
\ref{axm:iso_closure}.
The image under $U$ of an isomorphism in $\ensuremath{\mathcal{C}}^T$ is an isomorphism in \ensuremath{\mathcal{C}}.
The class \ensuremath{\mathcal{E}}\ contains all isomorphisms in \ensuremath{\mathcal{C}}, so $\overline \E = U^{-1}(\ensuremath{\mathcal{E}})$ contains
all isomorphisms in $\ensuremath{\mathcal{C}}^T$.
By similar reasoning, $\overline \E$ is closed under composition with isomorphisms,
and $\overline \M$ also satisfies these conditions.
\ref{axm:orthogonal}.
We wish to show that $\overline \E \mathop{\bot} \overline \M$.
Take $T$-algebras
\[
\valg T A a,
\valg T B b,
\valg T I i,
\valg T J j
\]
and algebra maps
\[
\algmap e T A a I i,
\algmap m T J j B b,
\algmap f T A a J j,
\algmap g T I i B b
\]
where the first two maps are in $\overline \E$ and $\overline \M$ respectively.
Suppose that $ge = mf$.
Now, $e \in \ensuremath{\mathcal{E}}$ and $m \in \ensuremath{\mathcal{M}}$, so $e \mathop{\bot} m$, and there is a unique map $t$
in \ensuremath{\mathcal{C}}\ such that
\[
\commsquare A f J e m {I \unar[ur]^{\exists! t}} g B
\]
commutes.
We wish to show that $t$ is a map of $T$-algebras.
Consider the diagram
\[
\xymatrix{
TA \ar[r]^{Te} \ar[d]_a \ar@/u0.7cm/[rr]^{Tf}
&TI \ar[r]^{Tt} \ar[d]_i \ar@/u0.7cm/[rr]^{Tg}
& TJ \ar[r]^{Tm} \ar[d]^j
& TB \ar[d]^b \\
A \ar[r]^e \ar@/d0.7cm/[rr]_f
& I \ar[r]^t \ar@/d0.7cm/[rr]_g
& J \ar[r]^m
& B
}
\]
We wish to show that the middle square commutes: the assumptions tell us that
all other squares commute.
Recall that $T e \mathop{\bot} m$, and apply orthogonality to the square
\[
\commsquare{TA}{j Tf} J {Te} m {TI \unar[ur]^{\exists!u}} {gi} B
\]
Now,
\[
\begin{array}{rcll}
j(Tf) &= & fa & \mbox{($f$ is a map of $T$-algebras)} \\
&= & tea & \mbox{(Definition of $t$)} \\
&= & ti(Te) & \mbox{($e$ is a map of $T$-algebras)} \\
\end{array}
\]
and $mti = gi$ by definition of $t$, so $ti = u$ by uniqueness.
Similarly,
\[
\begin{array}{rcll}
gi &= & b(Tg) & \mbox{($g$ is a map of $T$-algebras)} \\
&= & b(Tm)(Tt) & \mbox{(Definition of $t$)} \\
&= & mj(Tt) & \mbox{($m$ is a map of $T$-algebras)} \\
\end{array}
\]
and $j(Tf) = j(Tt)(Te)$ by definition of $t$, so $j(Tt) = u$ by uniqueness.
Hence $j(Tt) = ti$, and $t$ is a map of $T$-algebras.
By construction, $t$ is unique.
So $e \mathop{\bot} m$ in $\ensuremath{\mathcal{C}}^T$, so $\overline \E \mathop{\bot} \overline \M$.
All the axioms are satisfied, and so $(\overline \E, \overline \M)$ is a factorization
system on $\ensuremath{\mathcal{C}}^T$.
\end{proof}
\begin{example}
\index{factorization system!bijective on objects/full and faithful!on \cat{Cat}}
Let $(\ensuremath{\mathcal{E}}, \ensuremath{\mathcal{M}})$ be the factorization system on \cat{Digraph}\ described in Example
\ref{ex:digraphFS} above, and let $T$ be the free category monad.
\cat{Cat}\ is monadic over \cat{Digraph}, and $T$ preserves the property of being
bijective on objects.
Hence, this gives a factorization system $(\overline \E, \overline \M)$ on
\cat{Cat}\ where $\overline \E$ is the collection of bijective-on-objects functors, and
$\overline \M$ is the collection of full and faithful functors.
\end{example}
\begin{example}
Similarly, there is a factorization system on $\cat{Digraph}^\natural$, where \ensuremath{\mathcal{E}}\ is
the class of maps that are pointwise bijective on objects, and \ensuremath{\mathcal{M}}\ is the
class of maps that are pointwise full and faithful.
This lifts to a factorization system $(\overline \E, \overline \M)$ on $\cat{Cat}^\natural$, in
which $\overline \E$ is the class of pointwise bijective-on-objects arrows, and
$\overline \M$ is the class of pointwise full-and-faithful arrows.
\end{example}
\begin{example}
\label{ex:catopdFS}
\index{factorization system!bijective on objects/full and faithful!on \cat{Cat-Operad}}
Let $\ensuremath{\mathcal{C}} = \cat{Cat}^\natural$, \ensuremath{\mathcal{E}}\ be the pointwise bijective-on-objects maps, and
\ensuremath{\mathcal{M}}\ be those that are pointwise full and faithful.
Since \cat{Cat-Operad}\ is monadic over $\cat{Cat}^\natural$ and the monad preserves
bijective-on-objects maps, this gives a factorization system $(\overline \E, \overline \M)$
on \cat{Cat-Operad}\ where $\overline \E$ is the class of levelwise bijective-on-objects
maps, and $\overline \M$ is the class of levelwise full and faithful ones.
Similarly, there is a factorization system $(\overline \E', \overline \M')$ on
\cat{Cat-$\Sigma$-Operad}\, where $\overline \E'$ is the class of bijective-on-objects maps, and
$\overline \M'$ is the class of levelwise full and faithful ones.
\end{example}
We shall need one final piece of background:
{
\def\ensuremath{\Set^X}{\ensuremath{\cat{Set}^X}}
\def\ensuremath{(\SetX)^T}{\ensuremath{(\ensuremath{\Set^X})^T}}
\begin{theorem}
\label{thm:reg epis split in Set^X^T}
If $X$ is a set and $T$ is a monad on \ensuremath{\Set^X}\ then the regular epis in
\ensuremath{(\SetX)^T}\ are the pointwise surjections.
In other words, the forgetful functor $U:\ensuremath{(\SetX)^T} \to \ensuremath{\Set^X}$ preserves and
reflects regular epis.
\end{theorem}
\begin{proof}
See again \cite{acc} section 20, in particular Definition 20.21 and Proposition
20.30.
\end{proof}
}
\chapter*{Acknowledgements}
\addcontentsline{toc}{chapter}{Acknowledgements}
\input{acknowledgements.tex}
\chapter*{Declaration}
\addcontentsline{toc}{chapter}{Declaration}
I declare that this thesis is my own original work, except where credited to
others.
This thesis does not include work forming part of a thesis presented
for another degree.
\setcounter{chapter}{-1}
\chapter{Introduction}
Many definitions exist of categories with some kind of ``weakened'' algebraic
structure, in which the defining equations hold only up to coherent
isomorphism.
The paradigmatic example is the theory of weak monoidal
categories, as presented in \cite{catwork}, but there are also definitions of
categories with weakened versions of the structure of groups \cite{baez+lauda},
Lie algebras \cite{baez+crans}, crossed monoids \cite{crossedmonoid},
sets acted on by a monoid \cite{modulecat}, rigs \cite{laplaza}, vector spaces
\cite{k+v} and others.
A general definition of such categories-with-weakened-structure is obviously
desirable, but hard in the general case.
In this thesis, we restrict our attention to the case of theories that can be
described by (possibly symmetric) operads, and present possible definitions of
weak $P$-category and weak $P$-functor for any symmetric operad $P$.
We show that this definition is independent (up to equivalence) of our choice
of presentation for $P$; this generalizes the equivalence of classical and
unbiased monoidal categories.
In support of our definition, we present a generalization of Joyal and Street's
result from \cite{j+s} that every weak monoidal category is monoidally
equivalent to a strict monoidal category: this holds straightforwardly when $P$
is a plain operad.
This generalization includes the classical theorem that every symmetric
monoidal category is equivalent via symmetric monoidal functors and
transformations to a symmetric monoidal category whose associators and unit
maps are identities.
The idea is to consider the strict models of our theory as algebras for an
operad, then to obtain the weak models as (strict) algebras for a weakened
version of that operad (which will be a \cat{Cat}-operad).
In particular, we do not make use of the pseudo-algebras of Blackwell, Kelly
and Power, for which see \cite{bkp}.
Their definition is related to ours in the non-symmetric case, however: we
explore the connections in Chapter \ref{ch:others}.
We weaken the operad using a similar approach to that used in Penon's
definition of $n$-category: see \cite{penon}, or \cite{cheng+lauda} for a
non-rigorous summary.
In Chapters \ref{ch:theories}, \ref{ch:operads} and \ref{ch:factsys}, we review
some essential background material on theories, operads and factorization
systems.
Most of this is well-known, and only one result (in Section \ref{sec:synclass})
is new.
In Chapter \ref{ch:categorification}, we present our definitions of weak
$P$-category, weak $P$-functor and $P$-transformation.
We start with a na\"ive, syntactic definition that is only effective for
strongly regular (plain-operadic) theories.
We then re-state this definition using the theory of factorization systems,
which allows us to apply it to the more general symmetric operads.
Section \ref{sec:symmmoncats} uses this definition to explicitly calculate the
categorification of the theory of commutative monoids with their standard
signature, and shows that this is exactly the classical theory of symmetric
monoidal categories.
In Chapter \ref{ch:coherence}, we treat the problem of different presentations
of a given operad: we use this to prove that the weakening of a given theory is
independent of the choice of presentation.
We also prove some theorems about strictification of weak $P$-categories.
In Chapter \ref{ch:others}, we compare our approach to other approaches to
categorification which have been proposed in the literature.
Material in this thesis has appeared in two previous papers: the
material on strictification for strongly regular theories was in my preprint
\cite{WkPeqStP}, and the material on signature-independence was in my paper
\cite{presindep}, which was presented at the 85th Peripatetic Seminar
on Sheaves and Logic in Nice in March 2007, and at CT 2007 in Carvoeiro,
Portugal.
\section{Remarks on notation}
Throughout this thesis, the set $\natural$ of natural numbers is taken to
include 0.
We shall occasionally adopt the ${}_\bullet$ notation from chain complexes and
write, for instance, $p_\bullet$ for a finite sequence $p_1, \dots, p_n$ and
$p_\bullet^\bullet$ for a double sequence.
We shall use the notation $\fs n$ to refer to the set $\{1, \dots, n\}$ for all $n \in \natural$: the set $\fs 0$ is the empty set.
We shall use the symbol 1 to refer to terminal objects of categories and
identity arrows, as well as to the first nonzero natural number; it is my hope
that no confusion results.
\chapter{Operads}
\label{ch:operads}
Operads arose in the study of homotopy theory with the work of Boardman and
Vogt \cite{b+v}, and May \cite{may}.
In that field they are an invaluable tool: \cite{mss} describes a diverse range
of applications.
Independently, multicategories (which are to operads as categories are to
monoids) had arisen in categorical logic with the work of Lambek
\cite{lambek}.
Multicategories are sometimes called ``coloured operads''.
We will use multicategories and operads as tools to approach universal algebra:
while operads are not as expressive as Lawvere theories, they can be easily
extended to be so, and the theories that \emph{can} be represented by operads
provide a useful ``toy problem'' to help us get started.
Informally, categories have objects and arrows, where an arrow has one source
and one target; multicategories have objects and arrows with one target but
multiple sources (see Fig. \ref{fig:multicat}); and operads are one-object
multicategories.
Multicategories (and thus operads) have a composition operation that is
associative and unital.
\begin{figure}
\centerline{
\epsfxsize=3in
\epsfbox{operad.eps}
}
\caption{Composition in a multicategory}
\label{fig:multicat}
\end{figure}
\section{Plain operads}
\label{sec:plain_operads}
\begin{defn}
\index{multicategory!plain}
A \defterm{plain multicategory} (or simply ``multicategory'') $\ensuremath{\mathcal{C}}$
consists of the following:
\begin{itemize}
\item a collection $\ensuremath{\mathcal{C}}_0$ of \emph{objects},
\item for all $n \in \natural$ and all $c_1, \dots, c_n,
d \in \ensuremath{\mathcal{C}}_0$, a set of \emph{arrows} $\ensuremath{\mathcal{C}}(c_1, \dots, c_n; d)$,
\item for all $n, k_1, \dots k_n \in \natural$ and $c_1^1 \dots, c_{k_n}^n,
d_1, \dots, d_n, e \in \ensuremath{\mathcal{C}}_0$, a function called \emph{composition}
\[\ocomp
: \ensuremath{\mathcal{C}}(d_1, \dots, d_n;e) \times
\ensuremath{\mathcal{C}}(c^1_1, \dots, c^1_{k_1};d_1) \times \dots \times
\ensuremath{\mathcal{C}}(c^n_1, \dots, c^n_{k_n};d_n)
\to
\ensuremath{\mathcal{C}}(c^1_1, \dots, c^n_{k_n};e)\]
\item for all $c \in \ensuremath{\mathcal{C}}$, an \emph{identity arrow} $1_c \in \ensuremath{\mathcal{C}}(c;c)$
\end{itemize}
satisfying the following axioms:
\begin{itemize}
\item \emph{Associativity:} $f \ocomp (g_\bullet \ocomp h\seq\udot) = (f \ocomp g_\bullet)
\ocomp h\seq\udot$ wherever this makes sense (we borrow the ${}_\bullet$ notation for
sequences from chain complexes)
\item \emph{Units:} $1 \ocomp f = f = f \ocomp (1, \dots, 1)$ for all $f$.
\end{itemize}
A plain multicategory \ensuremath{\mathcal{C}}\ is \defterm{small} if $\ensuremath{\mathcal{C}}_0$ forms a set.
In line with the definition above, we shall take all our multicategories to be
locally small: this restriction is not essential.
\end{defn}
We say that an arrow in $\ensuremath{\mathcal{C}}(c_1, \dots, c_n; d)$ is \defterm{$n$-ary}, or has
\defterm{arity $n$}.
We remark that taking $n = 0$ gives us \emph{nullary} arrows.
This is in contrast to the definition used by some authors, who do not allow
nullary arrows.
\begin{defn}
\index{morphism!of plain multicategories}
A \defterm{morphism} of multicategories $F: \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{D}}$ is a map $F : \ensuremath{\mathcal{C}}_0
\to \ensuremath{\mathcal{D}}_0$ together with maps $F : \ensuremath{\mathcal{C}}(c_1, \dots, c_n ; c) \to \ensuremath{\mathcal{D}}(F c_1,
\dots, F c_n ; F c)$ which commute with $\ocomp$ and identities.
\index{transformation!between morphisms of plain multicategories}
A \defterm{transformation} of multicategory maps $\alpha : F \to G$ is a family
of arrows $\alpha_c \in \ensuremath{\mathcal{D}}(F c ; G c)$, one for each $c \in \ensuremath{\mathcal{C}}$, satisfying the
analogue of the usual naturality squares: for all maps $f: c_1, \dots, c_k \to
c$ in \ensuremath{\mathcal{C}}, we must have
\[
\alpha_c \ocomp Ff = Gf \ocomp (\alpha_{c_1}, \dots, \alpha_{c_k})
\]
\end{defn}
One is tempted to write this last condition as
\[
\xymatrix{
Fc_1, \dots, Fc_k \ar[rr]^{Ff} \ar[d]_{\alpha_{c_1}, \dots, \alpha_{c_k}}
&& Fc \ar[d]^{\alpha_c} \\
Gc_1, \dots, Gc_k \ar[rr]_{Gf}
&& Gc
}
\]
but care must be taken: in a general multicategory, $\alpha_{c_1}, \dots,
\alpha_{c_k}$ does not correspond to any single map, as it would in a monoidal
category.
Small plain multicategories, their morphisms and their transformations form a
2-category: we shall use the notation \cat{Multicat}\ for both this 2-category and
its underlying 1-category. \index{\cat{Multicat}}
To simplify the presentation of our first example, we recall the notion of
unbiased monoidal category from \cite{hohc} section 3.1:
\begin{defn}
\label{def:umoncat}
\index{unbiased!monoidal category}
An \defterm{unbiased weak monoidal category} $(\ensuremath{\mathcal{C}}, \otimes, \gamma, \iota)$
consists of
\begin{itemize}
\item a category $\ensuremath{\mathcal{C}}$,
\item for each $n \in \natural$, a functor $\otimes_n : \ensuremath{\mathcal{C}}^n \to \ensuremath{\mathcal{C}}$ called \defterm{$n$-fold tensor} and written
\[
(a_1, \dots, a_n) \mapsto (a_1 \otimes \dots \otimes a_n)
\]
\item for each $n, k_1, \dots, k_n \in \natural$, a natural isomorphism
\[
\gamma : \otimes_n \ocomp (\otimes_{k_1} \times \dots \times \otimes_{k_n})
\longrightarrow \otimes_{\sum k_i}
\]
\item a natural isomorphism
\[
\iota: 1_A \to \otimes_1
\]
\end{itemize}
satisfying
\begin{itemize}
\item associativity: for any triple sequence $a\udot\ldseq$ of objects in $\ensuremath{\mathcal{C}}$, the
diagram
\[
\xymatrixrowsep{4pc}
\xymatrixcolsep{-10pc}
\xymatrix{
&(
((\otimes a_{1 1}^\bullet) \otimes \dots \otimes (\otimes a_{1 k_1}^\bullet))
\otimes \dots \otimes
((\otimes a_{n 1}^\bullet) \otimes \dots \otimes (\otimes a_{n k_n}^\bullet)))
\ar[dl] \ar[dr] \\
((\otimes a_{1 \bullet}^\bullet)
\otimes \dots \otimes
(\otimes a_{n \bullet}^\bullet))
\ar[dr]
& & ((\otimes a_{1 1}^\bullet) \otimes \dots \otimes
(\otimes a_{n k_n}^\bullet))
\ar[dl] \\
& (a_{1 1}^1 \otimes \dots \otimes a_{n k_n}^{m_{k_n}})
}
\]
commutes.
\item identity: for any $n \in \natural$ and any sequence $a_1, \dots, a_n$ of
objects in $\ensuremath{\mathcal{C}}$, the diagrams
\[
\xymatrixcolsep{4pc}
\xymatrixrowsep{4pc}
\xymatrix{
(a_1 \otimes \dots \otimes a_n) \ar[r]^{(\iota\otimes\dots\otimes\iota)}
\ar[dr]_1
& ((a_1) \otimes \dots \otimes (a_n)) \ar[d]^\gamma \\
& (a_1 \otimes \dots \otimes a_n)
}
\]
\[
\xymatrixcolsep{4pc}
\xymatrixrowsep{4pc}
\xymatrix{
(a_1 \otimes \dots \otimes a_n) \ar[r]^{\iota}
\ar[dr]_1
& ((a_1 \otimes \dots \otimes a_n)) \ar[d]^\gamma \\
& (a_1 \otimes \dots \otimes a_n)
}
\]
commute.
\end{itemize}
\end{defn}
\begin{example}
\index{multicategory!underlying multicategory of a monoidal category}
\index{unbiased!monoidal category!underlying multicategory}
Let $\ensuremath{\mathcal{C}}$ be a locally small unbiased weak monoidal category.
The \defterm{underlying multicategory} $\ensuremath{\mathcal{C}}'$ of $\ensuremath{\mathcal{C}}$ has
\begin{itemize}
\item objects: objects of $\ensuremath{\mathcal{C}}$;
\item arrows: $\ensuremath{\mathcal{C}}'(a_1, \dots, a_n; b) = \ensuremath{\mathcal{C}}(a_1\otimes\dots\otimes a_n, b)$;
\item composition given as follows: if $f_i \in \ensuremath{\mathcal{C}}'(a_1^i, \dots, a_{k_i}^i ;
b_i)$ for $i = 1, \dots, n$ and $g \in \ensuremath{\mathcal{C}}'(b_1, \dots, b_n; c)$, then we define
$g \ocomp (f_1, \dots, f_n)$ as
\[
\xymatrixnocompile{
\bigotimes_{i,j} a^i_j \ar[dd]_{f \ocomp (g_1, \dots, g_n)}
\ar[rr]^{\gamma \otimes \dots \otimes \gamma}
&& \bigotimes_i ( \bigotimes_j a^i_j ) \ar[dd]^{\bigotimes_i {f_i}} \\
\\
c
&& \bigotimes_i b_i \ar[ll]^g
}
\]
\end{itemize}
\end{example}
\begin{defn}
\index{algebra!for a plain multicategory}
Let $M$ and $\ensuremath{\mathcal{C}}$ be plain multicategories.
An \defterm{algebra} for $M$ in \ensuremath{\mathcal{C}}\ is a morphism of multicategories $M \to \ensuremath{\mathcal{C}}$.
\end{defn}
\begin{defn}
Let $M$ be a plain multicategory, and \ensuremath{\mathcal{C}}\ be an unbiased monoidal category.
An \defterm{algebra} for $M$ in \ensuremath{\mathcal{C}}\ is a morphism of multicategories from $M$
to the underlying multicategory of $\ensuremath{\mathcal{C}}$.
\end{defn}
A \defterm{plain operad} (or simply ``operad'') is now a one-object
multicategory. \index{operad!plain}
Morphisms and transformations of operads are defined as for general
multicategories.
\index{morphism!of plain operads}
\index{transformation!between morphisms of plain operads}
As before, we use the notation \cat{Operad}\ \index{\cat{Operad}} for both the 2-category
of operads, morphisms and transformations, and its underlying 1-category.
Operads are to multicategories as monoids are to categories: just as
with monoids, this allows us to present the theory of operads in a
simplified way.
\begin{lemma}
\index{operad!plain!concrete description}
\label{lem:operad_description}
An operad $P$ can be given by the following data:
\begin{itemize}
\item A sequence $P_0, P_1, \dots$ of sets
\item For all $n, k_1, \dots, k_n \in \natural$, a function $\ocomp: P_n
\times \prodkn{P} \to P_{\sum k_i}$
\item An \defterm{identity element} $1 \in P_1$
\end{itemize}
satisfying the following axioms:
\begin{itemize}
\item \emph{Associativity:} $f \ocomp (g_\bullet \ocomp h\seq\udot) = (f \ocomp g_\bullet)
\ocomp h\seq\udot$ wherever this makes sense
\item \emph{Units:} $1 \ocomp f = f = f \ocomp (1, \dots, 1)$ for all f.
\end{itemize}
\end{lemma}
\begin{proof}
Using the symbol $*$ for the unique object, let $P_n = P(*, \dots, * ; *)$,
where the input is repeated $n$ times.
The rest of the conditions follow trivially from the definition of
a multicategory.
\end{proof}
\begin{lemma}
Let $P$ and $Q$ be operads.
A morphism $f: P \to Q$ consists of a function $f_n : P_n \to Q_n$ for each $n
\in \natural$ such that, for all $n, k_1, \dots, k_n$, the diagram
\[
\xymatrix{
P_n \times \prodkn P \ar[d]_{f_n \times \prodkn f} \ar[rr]^\ocomp
&& P_{\sum k_i} \ar[d]^{f_{\sum k_i}} \\
Q_n \times \prodkn Q \ar[rr]_\ocomp
&& Q_{\sum k_i}
}
\]
commutes, and that $f_1$ preserves the identity object.
If $f$ and $g$ are morphisms of operads from $P$ to $Q$, then a transformation
from $f$ to $g$ is an element $\alpha \in Q_1$ such that $\alpha \ocomp F p = G p \ocomp (\alpha, \dots, \alpha)$ for all $n \in \natural$ and all $p \in P_n$.
\end{lemma}
\begin{proof}
Trivial.
\end{proof}
\begin{defn}
If a morphism of operads $f: P \to Q$ is such that $f_n$ has some property $X$
for all $n \in \natural$, we say that $f$ is \defterm{levelwise} $X$.
\end{defn}
\begin{example}
\label{ex:plEnd}
\index{endomorphism operad!plain}
Let $A$ be an object of a multicategory $\ensuremath{\mathcal{C}}$.
The \defterm{endomorphism operad} of $A$ is the full sub-multicategory
$\mathop{\rm{End}}(A)$ of $\ensuremath{\mathcal{C}}$ whose only object is $A$.
In terms of the description in Lemma \ref{lem:operad_description}, $\mathop{\rm{End}}(A)_n$
is the set of $n$-ary\ arrows from $A, \dots, A$ to $A$.
Composition is as in $\ensuremath{\mathcal{C}}$.
In particular, if $\ensuremath{\mathcal{C}}$ is the underlying multicategory of some monoidal
category $\ensuremath{\mathcal{C}}'$, then $\mathop{\rm{End}}(A)_n = \ensuremath{\mathcal{C}}'(A \otimes \dots \otimes A, A)$.
This is the case we shall use most frequently.
\end{example}
\begin{example}
\label{ex:symmopd}
\index{\ensuremath{\mathcal{S}}!as a plain operad}
\index{operad of symmetries|see{\ensuremath{\mathcal{S}}}}
There is an operad $\ensuremath{\mathcal{S}}$ for which each $\ensuremath{\mathcal{S}}_n$ is the symmetric group
$S_n$.
Operadic composition is given as follows: if $\sigma \in S_n$, and $\tau_i \in
S_{k_i}$ for $i = 1, \dots, n$, then
\[
\sigma\ocomp(\tau_1, \dots, \tau_n) : \sum_{i = 1}^j k_i + m \mapsto
\sum_{i: \sigma(i) < \sigma(j+1)} k_i + \tau_{j+1}(m)
\]
for all $j \in \{1, \dots, n\}$ and $m \in \{0, \dots, k_{j+1}-1\}$.
Informally, the inputs are divided into ``blocks'' of length $k_1, k_2, \dots,
k_n$, which are then permuted by $\sigma$: the elements of each block are then
permuted by the appropriate $\tau_i$.
For an example, see Figure \ref{fig:symm_comp}.
\begin{figure}[h]
\centerline{
\epsfxsize=5in
\epsfbox{symmcomp.eps}
}
\caption{Composition in the operad $\ensuremath{\mathcal{S}}$ of symmetries}
\label{fig:symm_comp}
\end{figure}
\end{example}
\begin{example}
\index{\ensuremath{\mathcal{B}}}
\index{operad of braids|see{\ensuremath{\mathcal{B}}}}
There is an operad $\ensuremath{\mathcal{B}}$ for which each $\ensuremath{\mathcal{B}}_n$ is the Artin braid group
$B_n$.
Composition is analogous to that for $\ensuremath{\mathcal{S}}$: the inputs are divided into
blocks, which are braided, and then the elements of the blocks are braided.
\end{example}
\begin{example}
\index{little $m$-discs operad}
\index{operad!little $m$-discs|see{little $m$-discs operad}}
Fix an $m \in \natural$.
There is an operad $LD$ for which each $LD_n$ is an embedding of $n$ copies of
the closed unit disc $D_m$ into $D_m$.
Composition is by gluing -- see Figure \ref{fig:little_discs}.
\begin{figure}[h]
\centerline{
\epsfxsize=5in
\epsfbox{little_discs.eps}
}
\caption{Composition in the little 2-discs operad}
\label{fig:little_discs}
\end{figure}
$LD$ is known as the \defterm{little $m$-discs operad}.
\end{example}
Since we wish to use operads to represent theories, we need to have some way
of describing the models of those theories.
\begin{defn}
\index{algebra!for a plain operad}
Let $P$ be an operad.
An \defterm{algebra} for $P$ in a multicategory $\ensuremath{\mathcal{C}}$ is an object $A \in \ensuremath{\mathcal{C}}$
and a morphism of operads $(\hat{\phantom{\alpha}}) : P \to \mathop{\rm{End}}(A)$.
\end{defn}
Where $\ensuremath{\mathcal{C}}$ is a monoidal category, this is equivalent to requiring an object
$A \in \ensuremath{\mathcal{C}}$, and for each $p \in P_n$ a morphism $\hat p : A^{\otimes n} \to A$
such that $\hat 1 = 1_A$ and $\hat p \ocomp (\hat q_1 \otimes \dots \otimes
\hat q_n) = \widehat{p \ocomp (q_1 \otimes \dots \otimes q_n)}$ for all $p, q_1,
\dots, q_n \in P$.
A third equivalent definition is, for each $n \in \natural$, a map $h_n: P_n
\otimes A^{\otimes n}
\to A$, such that $h_n(p, h_n(q_\bullet, -)) = h_{\sum k_i}(p \ocomp q_\bullet,
-)$ for all $p \in P_n$, $q_i \in P_{k_i}$, and $h_1(1, -) = 1_A$.
We leave the proofs of these equivalences as an easy exercise for the reader,
and will make use of whichever formulation is most convenient at the time.
\begin{defn}
\index{morphism!of algebras for a plain operad}
Let $P$ be a plain operad, and $(A, (\hat{\phantom{\alpha}}))$ and $(B, (\check{\phantom{\alpha}}))$ be
algebras for $P$ in a multicategory $\ensuremath{\mathcal{C}}$.
A \defterm{morphism} of algebras is an arrow $F : A \to B$ in $\ensuremath{\mathcal{C}}$ such that,
for all $n \in \natural$, the diagram
\[
\xymatrix{
P_n \ar[d]_{(\hat{\phantom{\alpha}})} \ar[r]^{(\check{\phantom{\alpha}})}
& \mathop{\rm{End}}(B)_n \ar[d]^{-\ocomp(F,\dots,F)} \\
\mathop{\rm{End}}(A)_n \ar[r]^-{F\ocomp -} & \ensuremath{\mathcal{C}}(A, \dots, A ; B)
}
\]
commutes.
\end{defn}
The definition of morphism may be stated equivalently in terms of any of the
three characterizations of algebras given above.
\section{Symmetric operads}
\label{sec:symm_operads}
\begin{defn}
\label{def:symm_mcat}
\index{multicategory!symmetric}
A \defterm{symmetric multicategory} is a multicategory \ensuremath{\mathcal{C}}\ and, for every $n
\in \natural$, every $\sigma \in S_n$, and every $A_1, \dots A_n, B \in \ensuremath{\mathcal{C}}$, a
map
\[
\begin{array}{rccl}
\sigma \act - : & \ensuremath{\mathcal{C}}(A_1, \dots, A_n; B) & \longrightarrow
& \ensuremath{\mathcal{C}}(A_{\sigma 1}, \dots, A_{\sigma n}; B) \\
& f & \longmapsto & \sigma \act f
\end{array}
\]
such that
\begin{itemize}
\item For each $f \in \ensuremath{\mathcal{C}}(A_1, \dots, A_n; B)$, $1\act f = f$.
\item For each $\sigma, \rho \in S_n$, and each $f \in \ensuremath{\mathcal{C}}(A_1, \dots, A_n; B)$,
\[
\rho \act (\sigma \act f) = (\rho\sigma) \act f
\]
\item For each permutation $\sigma \in S_n$, all objects $A^1_1, \dots,
A^n_{k_n}, B_1, \dots, B_n, C \in \ensuremath{\mathcal{C}}$ and all arrows $f_i \in \ensuremath{\mathcal{C}}(A^i_1, \dots,
A^i_{k_i}; B_i)$ and $g \in \ensuremath{\mathcal{C}}(B_1, \dots, B_n; C)$,
\[
(\sigma\act g) \ocomp (f_{\sigma 1}, \dots, f_{\sigma n})
= (\sigma \ocomp (1,\dots, 1)) \act (g \ocomp (f_1, \dots f_n)).
\]
\item For each $A^1_1, \dots, A^n_{k_n}, B_1, \dots, B_n, C \in \ensuremath{\mathcal{C}}$,
$\sigma_i \in S_{k_i}$ for $i=1,\dots, n$, and each $f_i \in \ensuremath{\mathcal{C}}(A^i_1, \dots,
A^i_{k_i}; B_i), g \in \ensuremath{\mathcal{C}}(B_1, \dots, B_n; C)$,
\[
g \ocomp (\sigma_1 \act f_1, \dots, \sigma_n \act f_n)
= (1 \ocomp (\sigma_1, \dots, \sigma_n)) \act (g\ocomp (f_1, \dots, f_n)).
\]
\end{itemize}
where $\sigma\ocomp(1, \dots, 1)$ and $1\ocomp(\sigma_1, \dots, \sigma_n)$ are
as defined in Example \ref{ex:symmopd}.
\end{defn}
This definition is unusual in that the symmetric groups act on the left rather
than on the right as is more common: however, this change is essential for our
later generalization to finite product multicategories\ in Section \ref{sec:dops}.
\begin{defn}
\index{morphism!of symmetric multicategories}
Let $\ensuremath{\mathcal{C}}_1$ and $\ensuremath{\mathcal{C}}_2$ be symmetric multicategories.
A \defterm{morphism} (or \defterm{map}) $F$ of symmetric multicategories is a
map $F: \ensuremath{\mathcal{C}}_1 \to \ensuremath{\mathcal{C}}_2$ of multicategories such that $F(\sigma \act f) = \sigma
\act F(f)$ for all $n \in \natural$, all $n$-ary\ $f$ in $\ensuremath{\mathcal{C}}_1$, and all $\sigma
\in S_n$.
\end{defn}
\begin{defn}
\index{algebra!for a symmetric multicategory}
Let $M$ and $\ensuremath{\mathcal{C}}$ be symmetric multicategories.
An \defterm{algebra} for $M$ in \ensuremath{\mathcal{C}}\ is a morphism of symmetric multicategories
$M \to \ensuremath{\mathcal{C}}$.
\end{defn}
\begin{defn}
\index{operad!symmetric}
A \defterm{symmetric operad} is a symmetric multicategory
with only one object.
\end{defn}
In this case, the definition is equivalent to the following:
\begin{defn}
\label{def:symmopd}
\xymatrixrowsep{5pc}
\xymatrixcolsep{6pc}
A \defterm{symmetric operad} is an operad $P$ together with an action of the
symmetric group $S_n$ on each $P_n$, which is compatible with the operadic
composition:
\[
\xymatrixrowsep{2pc}
\xymatrix{
P_n \times \prod P_{k_i} \ar[r]^{(\sigma \act -) \times 1 \times \dots
\times 1}
\ar[d]_{1 \times \sigma_*}
& P_n \times \prod P_{k_i} \ar[dd]^{\ocomp} \\
P_n \times \prod P_{\sigma k_i} \ar[d]_\ocomp
\\
P_{\sum k_i} \ar[r]^{(\sigma \ocomp (1, \dots, 1)) \act -}
& P_{\sum k_i}
}
\xymatrix{
P_n \times \prod P_{k_i}
\ar[r]^{1 \times (\rho_1 \act -) \times \dots
\times (\rho_n \act -)}
\ar[dd]_{\ocomp}
& P_n \times \prod P_{k_i} \ar[dd]^{\ocomp} \\
\\
P_{\sum k_i} \ar[r]^{(1 \ocomp (\rho_1, \dots, \rho_n)) \act -}
& P_{\sum k_i}
}
\]
\[
\xymatrix{
P_n \ar@/u0.5cm/[r]^{1 \act -} \ar@/d0.5cm/[r]_1 & P_n
}
\]
\end{defn}
Maps of symmetric operads are just maps of symmetric multicategories.
\begin{example}
\index{\ensuremath{\mathcal{S}}!as a symmetric operad}
The operad $\ensuremath{\mathcal{S}}$ of symmetric groups, as given in Example \ref{ex:symmopd}.
The action of $S_n$ on $\ensuremath{\mathcal{S}}_n$ is given by $\sigma \cdot \tau = \tau\sigma^{-1}$.
\end{example}
\begin{example}
\label{ex:symmEnd}
\index{endomorphism operad!symmetric}
Let $\ensuremath{\mathcal{C}}$ be a symmetric multicategory, and $A \in \ensuremath{\mathcal{C}}$.
The \defterm{symmetric endomorphism operad} $\mathop{\rm{End}}(A)$ of $A$ is the full sub-(symmetric multicategory) of $\ensuremath{\mathcal{C}}$ whose only object is $A$.
\end{example}
If $\ensuremath{\mathcal{C}}$ is the underlying symmetric multicategory of a symmetric monoidal
category, then $\mathop{\rm{End}}(A)_n = \ensuremath{\mathcal{C}}(A^{\otimes n}; A)$ for each $n \in \natural$, and
the actions of the symmetric groups are given by composition with the symmetry
maps.
\begin{defn}
\index{algebra!for a symmetric operad}
\index{morphism!of algebras for a symmetric operad}
Let $P$ be a symmetric operad.
An \defterm{algebra} for $P$ in a multicategory $\ensuremath{\mathcal{C}}$ is an object $A$ and a map $h : P \to \mathop{\rm{End}}(A)$ of symmetric operads.
A \defterm{morphism} $(A, h) \to (A', h')$ of $P$-algebras is an arrow $F: A
\to B$ in \ensuremath{\mathcal{C}}\ such that $h'F = Fh$.
\end{defn}
As with plain operads, the definitions of an algebra for a symmetric operad $P$
and of morphisms between those algebras may be stated in several equivalent
ways.
\section{Finite product operads}
\label{sec:dops}
The definition of categorification in Chapter \ref{ch:categorification} is
couched in terms of operads.
To generalize it, therefore, we might generalize the definition of operad so
that it is capable of expressing every (one-sorted) algebraic theory.
This generalization is not new: our ``finite product operads'' were presented by Tronin
under the name ``{\bf FinSet}-operads''. Our Theorem \ref{thm:DOpCat iso Clone}
appears in \cite{tronin}, and Theorem \ref{thm:dop-clone alg} appears as
Theorem 1.2 in \cite{tronin2}.
A fuller treatment was given by T.~Fiore (who called them ``the functional
forms of theories'') in \cite{fiore}.
Tronin's paper constructs an isomorphism between the category of finite product operads\ and the
category of algebraic clones which commutes with the forgetful functors to
$\cat{Set}^\natural$; Fiore's constructs an equivalence between the
category of finite product operads\ and that of Lawvere theories, and also shows that this
equivalence preserves the categories of algebras.
Let \ensuremath{\bbF}\ be a skeleton of the category of finite sets and functions, with
objects the sets \fs 0, \fs 1, \fs 2, \dots, where $\fs n = \{1,2,\dots,n\}$.
\index{\ensuremath{\bbF}}
\begin{defn}
\index{multicategory!finite product}
\label{dopdef}
A \defterm{finite product multicategory} is:
\begin{itemize}
\item A plain multicategory \ensuremath{\mathcal{C}};
\item for every morphism $f : \fs{n} \to \fs{m}$ in \ensuremath{\bbF}, and for all objects $C_1,
\dots, C_n, D \in \ensuremath{\mathcal{C}}$, a function $f\act - : \ensuremath{\mathcal{C}}(C_1, \dots, C_n;D) \to
\ensuremath{\mathcal{C}}(C_{f(1)}, \dots, C_{f(n)}; D)$
\end{itemize}
satisfying the following axioms:
\begin{itemize}
\item the \ensuremath{\bbF}-action is functorial: $f\act(g\act p) = (f\ocomp g)\act p$, and
$\id_{\fs n}\act p = p$ wherever these equations make sense;
\item the \ensuremath{\bbF}-action and multicategorical composition interact by ``combing
out'':
\[
(f\act p)\cmp{f_1\act p_1, \dots, f_n\act p_n} = (f\cmp{f_1,\dots, f_n)} \act
(p\cmp{p_{f(1)},\dots, p_{f(n)}})
\]
where $(f\cmp{f_1,\dots, f_n})$ is given as follows:
Let $f:\fs n \to \fs m$, and $f_i:{\fs k}_i \to {\fs j_i}$ for $i = 1, \dots,
n$.
Then
\[
\begin{array}{rcccl}
f\cmp{f_1, \dots, f_n} & : & \fs{\sum k_i} & \to & \fs{\sum j_i} \\
f\cmp{f_1,\dots, f_n} & : & \left(\sum_{i=1}^{p-1} k_{f(i)}\right) + h
& \mapsto & \left(\sum_{i=1}^{f(p)-1} j_i\right) + f_p(h)
\end{array}
\]
for all $p \in \{1, \dots, n\}$ and all $h \in \{1, \dots, k_{f(p)}\}$.
See Figure \ref{combfig}.
The small specks represent inputs to the arrow that are ignored.
\end{itemize}
\begin{figure}[h]
\centerline{
\epsfysize=2in
\epsfbox{combing.eps}
}
\caption{``Combing out'' the \ensuremath{\bbF}-action}
\label{combfig}
\end{figure}
\end{defn}
It is now possible to see why we chose to have our symmetries acting on the left
in Definition \ref{def:symm_mcat}: in this more general case, only a left
action is possible.
\begin{defn}
\index{operad!finite product}
A \defterm{finite product operad} is a finite product multicategory\ with only one object.
\end{defn}
We will see in Section \ref{sec:synclass} that finite product operads\ are equivalent in
expressive power to Lawvere theories or clones: hence, every finitary algebraic
theory provides an example of a finite product operad.
As before, the sets $P_n$ contain the $n$-ary\ operations in the theory.
For illustrative purposes, we work out two examples now:
\begin{example}
\index{algebra!for a ring}
\label{ex:Ralg_fp}
Let $R$ be a ring, and $P_n = R[x_1, \dots, x_n]$ (the set of polynomials in
$n$ commuting variables over $R$) for all $n \in \natural$.
If $p \in P_n$ and $q_i \in P_{k_i}$ for $i = 1, \dots, n$, then
\[
(p\ocomp(q_1, \dots, q_n))(x_1, \dots, x_{\sum_{i=1}^n k_i}) =
p(q_1(x_1, \dots, x_{k_1}), \dots, q_n(x_{(\sum_{i=1}^{n-1} k_i) + 1},
\dots, x_{\sum_{i=1}^n k_i}))
\]
and if $f: \fs{n} \to \fs{m}$, then
\[
(f \cdot p)(x_1, \dots, x_m) = p(x_{f(1)}, \dots, x_{f(n)})
\]
\end{example}
\begin{example}
\index{commutative monoids}
\label{ex:comm_monoid_fp}
Let $P_n$ be the set of elements of the free commutative monoid on $n$
variables $x_1, \dots, x_n$.
Elements of $P_n$ are in one-to-one-correspondence with elements of
$\natural^n$.
We call the $n$th component of $p \in P_n$ the \defterm{multiplicity of the
$n$th argument}.
Composition is defined as follows:
\def\vector#1#2{\left[\begin{array}{c} {#1} \\ \vdots \\ {#2}\end{array}\right]}
\def\veclist#1#2{\vector{#1_1}{#1_{#2}}}
\[
\veclist p n \ocomp \left(\veclist {q^1} {k_1}, \dots, \veclist {q^n}
{k_n}\right)
= \vector {p_1 q_1^1} {p_n q^n_{k_n}}
\]
and if $f : \fs{n} \to \fs{m}$,
\[
f \cdot \veclist p n = \vector {\sum_{f(i)=1} p_i} {\sum_{f(i)=m} p_i}
\]
Or, in more familiar notation:
\begin{eqnarray*}
(x_1^{p_1} \dots x_n^{p_n}) \ocomp (x_1^{q^1_1} \dots x_1^{q^1_{k_1}}, \dots
x_{(\sum_{i=1}^{n-1} k_i) + 1}^{q^n_1} \dots x_{\sum_{i=1}^n k_i}^{q^n_{k_n}})
& = &x_1^{p_1 q^1_1} x_2^{p_1 q^1_1} \dots x_{\sum_{i=1}^n k_i}^{p_n q^n_{k_n}}
\\
f\cdot(x_1^{p_1} \dots x_n^{p_n}) & = & x_{f(1)}^{p_1} \dots x_{f(n)}^{p_n} \\
& = & x_1^{\sum_{f(i)=1} p_i} \dots x_m^{\sum_{f(i)=m} p_i}
\end{eqnarray*}
\end{example}
\begin{example}
\label{ex:fpEnd}
\index{endomorphism operad!finite product}
Let $\ensuremath{\mathcal{C}}$ be a finite product category, and $A$ be an object of $\ensuremath{\mathcal{C}}$.
Then there is a finite product operad\ $\mathop{\rm{End}}(A)$, the \defterm{endomorphism operad} of $A$,
where $\mathop{\rm{End}}(A)_n = \ensuremath{\mathcal{C}}(A^n,A)$, and $f\act p$ is $p$ composed with the
appropriate combination of projections to relabel its arguments by $f$.
\end{example}
\begin{defn}
\index{morphism!of finite product multicategories}
Let $M$, $N$ be finite product multicategories.
A \defterm{morphism} $F: M \to N$ consists of
\begin{itemize}
\item for each object $m \in M$, an object $Fm \in N$;
\item for each $n \in \natural$ and all $m_1, \dots, m_n, m \in M$, a map
\[
F_{m_1, \dots, m_n, m}: M(m_1, \dots m_n; m) \to N(Fm_1, \dots, Fm_n; Fm)
\]
commuting with the \ensuremath{\bbF}-action, the unit and composition.
\end{itemize}
\end{defn}
\begin{defn}
\index{algebra!for a finite product multicategory}
Let $M$ be a finite product multicategory.
An \defterm{algebra} for $M$ in a finite product multicategory\ $\ensuremath{\mathcal{C}}$ is a map of finite product multicategories\ $M \to \ensuremath{\mathcal{C}}$.
An \defterm{algebra} for $M$ in a finite product category $\ensuremath{\mathcal{C}}$ is a map of
finite product multicategories\ from $M$ to the underlying finite product multicategory\ of \ensuremath{\mathcal{C}}.
Finite product multicategories\ and their morphisms form a category called {\cat{FP-Multicat}}. \index{{\cat{FP-Operad}}}
\end{defn}
In the special case of finite product operads, these definitions are equivalent to the
following:
\begin{defn}
\index{morphism!of finite product operads}
Let $P$, $Q$ be finite product operads.
A \defterm{morphism} $F: P \to Q$ is a sequence of maps $F_i: P_i \to
Q_i$ commuting with the \ensuremath{\bbF}\ action, the unit and composition.
\end{defn}
\begin{defn}
\index{algebra!for a finite product operad}
Let $P$ be a finite product operad.
An \defterm{algebra} for $P$ in a finite product category $\ensuremath{\mathcal{C}}$ is an object $A
\in \ensuremath{\mathcal{C}}$ and a map of finite product operads\ $P \to \mathop{\rm{End}}(A)$.
\end{defn}
Finite product operads\ and their morphisms form a category called {\cat{FP-Operad}}. \index{{\cat{FP-Operad}}}
\begin{example}
\index{algebra!for a ring}
The algebras in $\ensuremath{\mathcal{C}}$ for the operad described in Example \ref{ex:Ralg_fp} are
associative $R$-algebras in $\ensuremath{\mathcal{C}}$.
\end{example}
\begin{example}
\index{commutative monoids}
The algebras for the operad described in Example \ref{ex:comm_monoid_fp} are
commutative monoid objects in $\ensuremath{\mathcal{C}}$.
\end{example}
\begin{theorem}
\label{thm:DOpCat iso Clone}
${\cat{FP-Operad}} \cong \cat{Clone}$.
\end{theorem}
\begin{proof}
We shall construct a functor $K_{(-)} : {\cat{FP-Operad}} \to \cat{Clone}$, and show that it is
bijective on objects, full and faithful.
If $P$ is a finite product operad, let $K_P$ be the following clone:
\begin{itemize}
\item $(K_P)_n = P_n$ for all $n \in \natural$,
\item composition is given by composition in $P$: if $p \in P_n$ and $p_1,
\dots, p_n \in P_m$, then $p\ccomp(p_1, \dots, p_n) \in (K_P)_m$ is $f \act
(p\ocomp(p_1, \dots, p_n)) \in P_m$, where
\begin{equation}
\label{eqn:KPcomp}
\begin{array}{lccl}
f : &\fs{nm} &\to & \fs{m} \\
& x & \mapsto & ((x - 1) \mod m) + 1,
\end{array}
\end{equation}
\item for all $n \in \natural$ and all $i \in \fs{n}$, the projection
$\delta^i_n$ is $f^i_n \act 1$, where
\begin{equation}
\label{eqn:KPproj}
\begin{array}{lccl}
f^i_n : &\fs{1} &\to & \fs{n} \\
& 1 & \mapsto & i.
\end{array}
\end{equation}
\end{itemize}
It is easily checked that $K_P$ satisfies the axioms for a clone given in
Definition \ref{def:clone}.
On morphisms, $K_{(-)}$ acts trivially: morphisms of clones and of finite product operads\ are
simply maps of signatures commuting with the extra structure, and $K_{(-)}$
preserves the underlying map of signatures.
Let $K$ be a clone.
Let $P_K$ be the finite product operad\ for which
\begin{itemize}
\item $(P_K)_n = K_n$ for all $n \in \natural$,
\item $1 = \delta^1_1$,
\item $p \ocomp (p_1, \dots, p_n) = p \ccomp (p_1 \ccomp (\delta^1_m, \dots,
\delta^{k_1}_{\sum k_i}), \dots, p_n \ccomp (\delta^{k_1 + \dots + k_{n-1} +
1}_{\sum k_i}, \dots, \delta^{\sum k_i}_{\sum k_i}))$ for all $n$ and $k_1,
\dots, k_n \in \natural$, all $p \in K_n$, and all $p_1 \in K_{k_1}, \dots, p_n
\in K_{k_n}$,
\item $f \act p = p \ccomp (\delta^{f(1)}_m, \dots, \delta^{f(n)}_m)$ for all
$n, m \in \natural$, all $f : \fs{n} \to \fs{m}$ and all $p \in K_n$.
\end{itemize}
We will show that $K_{P_K} = K$ for all $K \in \cat{Clone}$, and that $P_{K_P} = P$
for all $P \in {\cat{FP-Operad}}$.
Let $K$ be a clone.
Then $(K_{P_K})_n = (P_K)_n = K_n$ for all $n \in \natural$.
If $n, m \in \natural$, $k \in K_n$ and $k_1, \dots, k_n \in K_m$, then the
composite $k\ccomp(k_1, \dots, k_n)$ in $K_{P_K}$ is given by the composite
$f \act (k\ocomp(k_1, \dots, k_n))$ in $P_K$, where $f$ is given by
(\ref{eqn:KPcomp}) above.
This in turn is given by the composite
\begin{eqnarray*}
(k & \ccomp & (k_1 \ccomp (\delta^1_{nm},
\dots, \delta^{m}_{nm}), \dots, k_n \ccomp (\delta^{(n-1)m + 1}_{nm}, \dots,
\delta^{nm}_{nm}))) \\
& \ccomp & (\delta^1_m, \dots, \delta^m_m, \dots, \delta^1_m, \dots, \delta^m_m)
\end{eqnarray*}
in $K$.
By the associativity law for clones, this is equal to
\[
\begin{array}{rl}
k \ccomp (& k_1 \ccomp (\delta^1_{nm}, \dots, \delta^{m}_{nm})
\ccomp (\delta^1_m, \dots, \delta^m_m, \dots, \delta^1_m, \dots, \delta^m_m), \\
& \dots, \\
& k_n \ccomp (\delta^{(n-1)m + 1}_{nm}, \dots,\delta^{nm}_{nm})
\ccomp (\delta^1_m, \dots, \delta^m_m, \dots, \delta^1_m, \dots, \delta^m_m))
\end{array}
\]
which in turn may be simplified to $k \ccomp (k_1, \dots, k_n)$ as required.
For every $n \in \natural$ and every $i \in \fs{n}$, the projection
$\delta^i_n$ in $K_{P_K}$ is given by $f^i_n \act 1$, where $f^i_n$ is defined
in (\ref{eqn:KPproj}): this in turn is given by $1 \ocomp
(\delta^i_n) = \delta^1_1 \ccomp (\delta^i_n) = \delta^i_n$.
Hence $K_{P_K} = K$.
Conversely, let $P$ be a finite product operad\; we shall show that $P_{K_P} = P$.
For every $n \in \natural$, the set $(P_{K_P})_n$ is equal to $P_n$.
The unit element is given by $1 = \delta^1_1 = f^1_1 \act 1 = 1$.
If $p \in P_n$ and $p_i \in P_{k_i}$ for $i = 1, \dots, n$, then the composite
$p\ocomp(p_1, \dots, p_n)$ is given by $p \ccomp (p_1 \ccomp (\delta^1_m, \dots,
\delta^{k_1}_{\sum k_i}), \dots, p_n \ccomp (\delta^{k_1 + \dots + k_{n-1} +
1}_{\sum k_i}, \dots, \delta^{\sum k_i}_{\sum k_i}))$ in $K_P$, which in turn
is given by (after simplification) $p \ocomp (p_1, \dots, p_n)$ in $P$.
Hence $P = P_{K_P}$, and $K_{(-)}$ is bijective on objects.
The reasoning above also suffices to show that $K_{(-)}$ is well-defined on
morphisms and full (since preserving a finite product operad\ structure amounts exactly to
preserving the associated clone structure).
Since the morphisms of both categories are simply maps of signatures with extra
properties and $K_{(-)}$ commutes with the forgetful functors to
$\cat{Set}^\natural$,
then $K_{(-)}$ is faithful.
Hence $K_{(-)}$ is an isomorphism of categories, and ${\cat{FP-Operad}} \cong \cat{Clone}$.
\end{proof}
\begin{theorem}
\label{thm:dop-clone alg}
Let $P$ be a finite product operad.
Then $\Alg P \cong \Alg{K_P}$.
\end{theorem}
\begin{proof}
Let $(A, (\hat{\phantom{\alpha}}))$ be a $P$-algebra.
Since the elements of the finite product endomorphism operad $\mathop{\rm{End}}(A)$ are
endomorphisms of $A$, and composition is given by composition of morphisms,
then $K_{\mathop{\rm{End}}(A)} = \mathop{\rm{End}}(A)$, the endomorphism clone of $A$.
Since the functor $K_{(-)} : {\cat{FP-Operad}} \to \cat{Clone}$ is an isomorphism, a morphism
of finite product operads\ $P \to \mathop{\rm{End}}(A)$ is exactly a map of clones $K_P \to \mathop{\rm{End}}(A)$.
Hence an algebra for $P$ is exactly an algebra for $K_P$.
A morphism between $P$-algebras is a morphism between their underlying objects
that commutes with $\hat p$ for every $p \in P_n$ and every $n \in \natural$;
this is true iff it commutes with $\hat k$ for every $k \in (K_P)_n$ and every
$n \in \natural$.
\end{proof}
\section{Adjunctions}
\label{sec:adj}
In the next few sections, we shall show that there is a chain of monadic
adjunctions
\index{adjunctions!for operads}
\begin{equation}
\label{eqn:adjs}
\xymatrix{
{\cat{FP-Operad}} \ar@<1.2ex>[d]^{\ensuremath{U^{\rm fp}_{\Sigma}}}
\ar@/r1.4cm/[dd]^{\ensuremath{U^{\rm fp}_{\rm pl}}}
\ar@/r3cm/[ddd]^{\ensuremath{U^{\rm fp}}} \\
\cat{$\Sigma$-Operad} \ar@<1.2ex>[d]^{\ensuremath{U^{\Sigma}_{\rm pl}}} \ar@<1.2ex>[u]^{\ensuremath{F_{\rm fp}^{\Sigma}}}_{\dashv}
\ar@/r1.4cm/[dd]^{\ensuremath{U^{\Sigma}}} \\
\cat{Operad} \ar@<1.2ex>[u]^{\ensuremath{F_{\Sigma}^{\rm pl}}}_{\dashv} \ar@<1.2ex>[d]^{\ensuremath{U^{\rm pl}}}
\ar@/l1.4cm/[uu]^{\ensuremath{F_{\rm fp}^{\rm pl}}} \\
\cat{Set}^\natural \ar@<1.2ex>[u]^{\ensuremath{F_{\rm pl}}}_{\dashv} \ar@/l1.4cm/[uu]^{\ensuremath{F_{\Sigma}}}
\ar@/l3cm/[uuu]^{\ensuremath{F_{\rm fp}}}
}
\end{equation}
The notation is chosen such that $F^x_y \dashv U_x^y$, and $U_x^y U_y^z =
U_x^z$.
The notation is inspired by the exponential notation used for hom-objects: the
source category of one of these functors is determined by its superscript, and
the target category is determined by its subscript.
The ``pl'' stands for ``plain''.
A similar chain of adjunctions (for PROPs rather than operads) was discussed in
\cite{baez_ualg}, pages 51--59.
We refer to the monad $U_x^yF^x_y$ as $T^x_y$.
The right adjoints ${\ensuremath{U^{\rm pl}}}, {\ensuremath{U^{\Sigma}_{\rm pl}}}$ and ${\ensuremath{U^{\rm fp}_{\Sigma}}}$ are found by forgetting
respectively the compositional structure, the symmetric structure, and the
actions of all non-bijective functions, and will not be described further.
By standard properties of adjunctions, the composite functors are adjoint:
${\ensuremath{F_{\Sigma}}} \dashv {\ensuremath{U^{\Sigma}}}$ etc.
\section{Existence and monadicity}
All the left adjoints in (\ref{eqn:adjs}) are examples of a more general
construction.
We shall now investigate this general case, and show that the adjunction which
arises is always monadic.
But first, we have so far only asserted that ${\ensuremath{U^{\rm fp}}}, {\ensuremath{U^{\rm fp}_{\Sigma}}}$ and ${\ensuremath{U^{\rm fp}_{\rm pl}}}$ have
left adjoints.
We shall show that these left adjoints must exist for general reasons.
Let $\cat{FP}$ be the category of small categories with finite products and
product-preserving functors.
\begin{lemma}
\label{lem:frees_exist}
Let $\Csmall$ and $\Dsmall$ be small finite-product categories, let $\ensuremath{\mathcal{C}}$ be
cartesian closed and have all small colimits, and let $Q: \Csmall \to \Dsmall$
preserve finite products.
Then the adjunction
\[
\adjunction{[\Dsmall,\ensuremath{\mathcal{C}}]}{[\Csmall,\ensuremath{\mathcal{C}}]}{Q_!}{Q^*},
\]
where $Q^*$ is composition with $Q$ and $Q_! = \mathop{\rm{Lan}}_Q$, restricts to an
adjunction
\[
\adjunction{\cat{FP}(\Dsmall,\ensuremath{\mathcal{C}})}{\cat{FP}(\Csmall,\ensuremath{\mathcal{C}})}{Q_!}{Q^*},
\]
\label{lem:leftadj}
\end{lemma}
\begin{proof}
\def[\Csmall, \C]{[\Csmall, \ensuremath{\mathcal{C}}]}
\def[\Dsmall, \C]{[\Dsmall, \ensuremath{\mathcal{C}}]}
\def\FP(\Csmall, \C){\cat{FP}(\Csmall, \ensuremath{\mathcal{C}})}
\def\FP(\Dsmall, \C){\cat{FP}(\Dsmall, \ensuremath{\mathcal{C}})}
Certainly $Q^*$ restricts in this way, since $Q$ preserves finite products.
$\FP(\Csmall, \C)$ and $\FP(\Dsmall, \C)$ are full subcategories of $[\Csmall, \C]$
and $[\Dsmall, \C]$, so if we can show that $Q_!$ restricts to a functor $\FP(\Csmall, \C) \to
\FP(\Dsmall, \C)$, then it is automatically left adjoint to the restriction of $Q^*$.
Let $X : \Csmall \to \ensuremath{\mathcal{C}}$ preserve finite products.
We must show that $Q_! X : \Dsmall \to \ensuremath{\mathcal{C}}$ preserves finite products.
We shall proceed by showing that $Q_! X$ preserves terminal objects and binary
products.
Recall that
\[
(Q_! X)(b) \cong \int^a \Dsmall(Q a, b) \times X a
\]
for all $b \in \Dsmall$.
Hence, using $1$ for the terminal objects in $\Dsmall$ and $\ensuremath{\mathcal{C}}$,
\begin{eqnarray*}
(Q_! X)(1) & \cong & \int^a \Dsmall(Q a, 1) \times X a \\
& \cong & \int^a 1 \times X a \\
& \cong & \int^a X a \\
& \cong & X 1 \\
& \cong & 1
\end{eqnarray*}
since $X$ preserves finite products and the colimit of a diagram $D$ over a
category with a terminal object $1$ is simply $D 1$.
Now, let $b_1, b_2 \in \Dsmall$.
Then
\begin{eqnarray}
\label{eq:defQ!}
\nonumber\lefteqn{(Q_! X)(b_1 \times b_2)} \\
& \cong & \int^a \Dsmall(Q a, b_1 \times b_2) \times X a
\\
\label{eq:defprod1}
& \cong & \int^a \Dsmall(Q a, b_1) \times \Dsmall(Q a, b_2) \times X a \\
\label{eq:density1}
& \cong & \int^a
\left(\int^{c_1} \Dsmall(Q c_1, b_1) \times \Csmall(a, c_1) \right)
\times
\left(\int^{c_2} \Dsmall(Q c_2, b_2) \times \Csmall(a, c_2) \right)
\times Xc \\
\label{eq:dist1}
& \cong & \int^{a, c_1, c_2} \Dsmall(Q c_1, b_1) \times \Dsmall(Q c_2, b_2) \times
\Csmall(a, c_1) \times \Csmall(a, c_2) \times X a \\
\label{eq:defprod2}
& \cong & \int^{a, c_1, c_2} \Dsmall(Q c_1, b_1) \times \Dsmall(Q c_2, b_2) \times
\Csmall(a, c_1 \times c_2) \times X a \\
\label{eq:dist2}
& \cong & \int^{c_1, c_2} \Dsmall(Q c_1, b_1) \times \Dsmall(Q c_2, b_2) \times
\left( \int^a \Csmall(a, c_1 \times c_2) \times X a \right) \\
\label{eq:density2}
& \cong & \int^{c_1, c_2} \Dsmall(Q c_1, b_1) \times \Dsmall(Q c_2, b_2) \times
X(c_1 \times c_2) \\
\label{eq:defprod3}
& \cong & \int^{c_1, c_2} \Dsmall(Q c_1, b_1) \times \Dsmall(Q c_2, b_2) \times
X c_1 \times X c_2 \\
\label{eq:Xprod}
& \cong & \int^{c_1, c_2} \Dsmall(Q c_1, b_1) \times X c_1 \times
\Dsmall(Q c_2, b_2) \times X c_2 \\
\label{eq:rearrange}
& \cong & \left (\int^{c_1} \Dsmall(Q c_1, b_1) \times X c_1 \right) \times
\left( \int^{c_2} \Dsmall(Q c_2, b_2) \times X c_2 \right) \\
\label{eq:dist3}
& \cong & (Q_!X)(b_1) \times (Q_!X)(b_2)
\end{eqnarray}
(\ref{eq:defQ!}) is the definition of $Q_!$;
(\ref{eq:defprod1}), (\ref{eq:defprod2}) and (\ref{eq:defprod3}) are from the
definition of products;
(\ref{eq:density1}) and (\ref{eq:density2}) are applications of the Density
Formula;
(\ref{eq:dist1}), (\ref{eq:dist2}) and (\ref{eq:dist3}) use the
distributivity of products over colimits in \ensuremath{\mathcal{C}}\ (since \ensuremath{\mathcal{C}}\ is cartesian
closed),
and (\ref{eq:Xprod}) uses the fact that $X$ preserves finite products.
So $Q_!$ preserves terminal objects and binary products, and hence all finite
products.
\end{proof}
\begin{corollary}
The functors ${\ensuremath{U^{\rm fp}_{\Sigma}}}, {\ensuremath{U^{\rm fp}_{\rm pl}}}$ and ${\ensuremath{U^{\rm fp}}}$ all have left adjoints.
\end{corollary}
\begin{lemma}
\label{lem:monadj}
Let $S$ be a set, whose elements we will call \defterm{sorts}.
Let $T$ and $T'$ be $S$-sorted finite product theories, such that $T'$ is a
subcategory of $T$ and the inclusion of $T'$ into $T$ preserves finite
products.
Let $\Alg T$ be the category of $T$-algebras and morphisms in some finite
product category $\ensuremath{\mathcal{C}}$, and $\Alg{T'}$ be the category of $T'$-algebras and
morphisms in $\ensuremath{\mathcal{C}}$.
Then the free/forgetful adjunction
\[
\adjunction {\Alg{T'}} {\Alg{T}} F U
\]
is monadic, provided the left adjoint $F$ exists.
\end{lemma}
\begin{proof}
\def\prods#1{\prod #1 s_i}
\def\prodr#1{\prod #1 r_j}
\def\prodars#1{\prod #1_{s_i}}
\def\prodarr#1{\prod #1_{r_j}}
We will make use of Beck's theorem to prove monadicity: precisely, we shall
make use of the version in \cite{catwork} VI.7.1, which states that $U$ is
monadic if it has a left adjoint and it strictly creates coequalizers for
$U$-absolute coequalizer pairs.
Recall that a functor $G: \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{D}}$ \defterm{strictly creates coequalizers}
for a diagram $\parallelpair A f g B$ in $\ensuremath{\mathcal{C}}$ if, for every coequalizer $e: GB
\to E$ of $Gf$ and $Gg$ in $\ensuremath{\mathcal{D}}$, there are a unique object $E'$ in $\ensuremath{\mathcal{C}}$ and a
unique arrow $e' : B \to E'$ such that $GE' = E$ and $Ge' = e$, and moreover
that $e'$ is a coequalizer of $\parallelpair A f g B$.
Let $\parallelpair A f g B$ be a $U$-absolute coequalizer pair in $\Alg{T}$,
and $e: U B \to E$ be the coequalizer of $\parallelpair{UA}{Uf}{Ug}{UB}$.
We wish to extend $E$ to a functor $E': T \to \ensuremath{\mathcal{C}}$.
Define $E'$ to be equal to $E$ on objects.
On arrows, we shall define $E'$ using the universal property of $E$ and the
$U$-absolute property of $\parallelpair A f g B$.
For each arrow $\phi : s_1 \times \dots \times s_n \to r_1 \times \dots
\times r_m$ in $T$ (where $s_i, r_j \in S$),
consider the diagram
\begin{equation}
\label{eqn:coeq}
\xymatrixrowsep{4pc}
\xymatrixcolsep{4pc}
\xymatrixnocompile{
\prods{U A} \parallelars{\prodars{U f}}{\prodars{Ug}} \ar[d]_{A\phi}
& \prods{U B} \ar[r]^{\prodars e} \ar[d]_{B\phi}
& \prods{E} \unar[d]^{E'\phi}\\
\prodr{U A} \parallelars{\prodarr{f}}{\prodarr{g}}
& \prodr{U B} \ar[r]^{\prodarr e}
& \prodr E
}
\end{equation}
in $\ensuremath{\mathcal{C}}$, where $e: UB \to E$ is a coequalizer for $\parallelpair {UA}
{Uf} {Ug} {UB}$.
Since $\parallelpair A f g B$ is a $U$-absolute coequalizer pair, $\prodars{e}
: \prods{UB} \to \prods{E}$ is a coequalizer.
Since $f$ and $g$ are $T$-homomorphisms, (\ref{eqn:coeq}) serially commutes, so
$(\prodarr e)\phi$ factors uniquely through $\prodars e$.
Define $E'\phi$ to be this map, as shown (and note that $E'\phi = E\phi$ if
$\phi$ is in $T'$).
This definition straightforwardly makes $E'$ into a functor $T \to \ensuremath{\mathcal{C}}$.
Since $E$ is a $T'$-algebra, and products in $T$ are the same as products in
$T'$, we may deduce that $E': T\to C$ preserves finite products, and thus is
a $T$-algebra.
Clearly, $E'$ is the unique extension of $E$ to a $T$-algebra such that $e$ is
a $T$-algebra morphism.
It remains to show that $e$ is a coequalizer map for $\parallelpair A f g B$ in
$\Alg{T}$.
Suppose $\fork A f g B d D$ is a fork in $\Alg T$.
Then $\fork {UA} {Uf} {Ug} {UB} {Ud} {UD}$ is a fork in $\Alg{T'}$, so $Ud$
factors through $e$; say $Ud = he$.
We must show that $h$ is a $T$-homomorphism.
As before, take $\phi : s_1 \times \dots \times s_n \to r_1 \times \dots \times
r_m$ in $T$, and consider the diagram
\begin{equation}
\xymatrixcolsep{4pc}
\xymatrixnocompile{
\prods{U A} \parallelars{\prodars{U f}}{\prodars{U g}} \ar[dd]_{A\phi}
& \prods{U B} \ar[r]^{\prodars{e}}
\ar[dd]_{B\phi} \ar[dr]_{\prodars{Ud}}
& \prods{E'} \ar@/r1cm/[dd]^{E'\phi} \unar[d]_{\prodars{h}} \\
& & \prods{U D} \ar@/r1cm/[dd]^{D\phi} \\
\prodr{U A} \parallelars{\prodarr f}{\prodarr g}
& \prodr{U B} \ar[r]^{\prodarr e} \ar[dr]_{\prodarr{Ud}}
& \prodr{E'} \unar[d]_{\prodarr{h}} \\
& & \prodr{D}
}
\end{equation}
We must show that the curved square on the far right commutes.
Now $(D\phi) \fcomp (\prodars d) = (D\phi) \fcomp (\prodars h) \fcomp (\prodars
e)$, and $(\prodarr{Ud})\fcomp(UB\phi) = h\fcomp e\fcomp(UB\phi) = h \fcomp
(E\phi) \fcomp (\prodars e)$, since $e$ is a $T$-algebra homomorphism.
But $(\prodarr {Ud})\fcomp \phi = \phi \fcomp (\prod d_i)$, so $D\phi \fcomp
(\prodars{h}) \fcomp (\prodars{e}) = h \fcomp (E\phi) \fcomp (\prodars{e})$.
And $\prodars{e}$ is (regular) epic, so $h \fcomp (E\phi) = (D\phi) \fcomp
(\prodars{h})$.
So $h$ is a $T$-algebra homomorphism.
Hence $U$ strictly creates coequalizers for $U$-absolute coequalizer pairs,
and hence is monadic.
\end{proof}
This result could also have been deduced from the Sandwich Theorem of Manes:
see \cite{manes} Theorem 3.1.29 (page 182).
\begin{theorem}
\label{thm:opd monadic over setN}
\index{\cat{Operad}!monadicity over $\cat{Set}^\natural$}
All the adjunctions in diagram \ref{eqn:adjs}, namely ${\ensuremath{F_{\rm fp}^{\Sigma}}} \dashv {\ensuremath{U^{\rm fp}_{\Sigma}}}, {\ensuremath{F_{\rm fp}^{\rm pl}}} \dashv {\ensuremath{U^{\rm fp}_{\rm pl}}}, {\ensuremath{F_{\rm fp}}} \dashv
{\ensuremath{U^{\rm fp}}}, {\ensuremath{F_{\Sigma}^{\rm pl}}} \dashv {\ensuremath{U^{\Sigma}_{\rm pl}}}, {\ensuremath{F_{\Sigma}}} \dashv {\ensuremath{U^{\Sigma}}}$ and ${\ensuremath{F_{\rm pl}}} \dashv {\ensuremath{U^{\rm pl}}}$, are
monadic.
\end{theorem}
\begin{proof}
Each category mentioned is a category of algebras for some $\natural$-sorted
theory, and the monadicity of each adjunction mentioned is obtained by a simple
application of Lemma \ref{lem:monadj}. For instance, symmetric operads are
algebras for the theory presented by
\begin{itemize}
\item \emph{operations}: one of the appropriate arity for each composition
operation in Definition \ref{def:symmopd}, and an operation $\sigma \cdot -$
for each $n \in \natural$ and each $\sigma$ in $S_n$.
\item \emph{equations}: one for each instance of the axioms in Definition
\ref{def:symmopd}, and an equation $(\sigma \cdot -)\fcomp (\rho \cdot -) =
\sigma\rho \cdot -$ for each $\sigma, \rho \in S_n$ and every $n \in \natural$.
\end{itemize}
\end{proof}
\section{Explicit construction of ${\ensuremath{F_{\rm pl}}}$ and ${\ensuremath{F_{\Sigma}^{\rm pl}}}$}
\label{sec:explicit_Fs}
The previous section showed that ${\ensuremath{F_{\rm pl}}}$ and ${\ensuremath{F_{\Sigma}^{\rm pl}}}$ exist for general reasons,
but it will be useful later to have an explicit construction of these functors.
For this reason, we shall now explicitly construct functors $\cat{Set}^\natural \to
\cat{Operad}$ and $\cat{Operad} \to \cat{$\Sigma$-Operad}$, and prove that they are left adjoint to
${\ensuremath{U^{\rm pl}}}$ and ${\ensuremath{U^{\Sigma}_{\rm pl}}}$.
\begin{defn}
\label{def:srterm}
\index{tree!strongly regular}
\index{strongly regular!tree|see{tree, strongly regular}}
Let $\Phi$ be a signature.
An \defterm{$n$-ary\ strongly regular tree labelled by $\Phi$} is an element of
the set $\mathop{\mbox{tr}}_n \Phi$, which is recursively defined as follows:
\begin{itemize}
\item $|$ is an element of $\mathop{\mbox{tr}}_1 \Phi$.
\item If $\phi \in \Phi_n$, and $\tau_1 \in \mathop{\mbox{tr}}_{k_1} \Phi, \dots, \tau_n \in
\mathop{\mbox{tr}}_{k_n} \Phi$, then $\phi \cmp{\tau_1, \dots, \tau_n} \in
\mathop{\mbox{tr}}_{\sum k_i} \Phi$.
\end{itemize}
\end{defn}
In graph-theoretic terms, all our trees are planar and rooted.
They need not be level.
We shall abuse notation and write $\phi$ instead of $\phi \cmp{|, \dots, |}$,
for $\phi \in \Phi_n$.
Given a signature $\Phi$, the objects of the plain operad $({\ensuremath{F_{\rm pl}}} \Phi)_n$ are
the elements of $\mathop{\mbox{tr}}_n \Phi$, and composition is given by grafting of trees:
\begin{itemize}
\index{grafting!of trees}
\item $| \cmp \tau$ = $\tau$
\item If $\tau_1 \in \mathop{\mbox{tr}}_{k_1} \Phi, \dots, \tau_n \in \mathop{\mbox{tr}}_{k_n} \Phi$, then
\[
(\phi \cmp{\tau_1,\dots, \tau_n}) \cmp{\sigma_1, \dots, \sigma_{\sum k_i}}
= \phi \cmp{\tau_1 \cmp {\sigma_1, \dots, \sigma_{k_1}}, \dots,
\tau_n \cmp {\sigma_{(\sum k_i) - k_n + 1}, \dots, \sigma_{\sum k_i}}}
\]
\end{itemize}
See Figure \ref{fig:grafting}.
\begin{figure}
\centerline{
\epsfxsize=5in
\epsfbox{grafting.eps}
}
\caption{Grafting of trees}
\label{fig:grafting}
\end{figure}
The unary tree $|$ is thus the identity in $({\ensuremath{F_{\rm pl}}} \Phi)$.
${\ensuremath{F_{\rm pl}}}$ acts on arrows as follows.
Let $f: \Phi \to \Psi$ be a map of signatures.
Then:
\begin{itemize}
\item $({\ensuremath{F_{\rm pl}}} f) | = |$
\item $({\ensuremath{F_{\rm pl}}} f) (\phi \ocomp (\tau_1, \dots, \tau_n))
= (f \phi) \ocomp ( ({\ensuremath{F_{\rm pl}}} f)\tau_1, \dots, ({\ensuremath{F_{\rm pl}}} f)\tau_n )$
\end{itemize}
It is readily verified that with this definition ${\ensuremath{F_{\rm pl}}}$ is a functor
$\cat{Set}^\natural \to \cat{Operad}$.
We define natural transformations $\eta: 1_{\cat{Set}^\natural} \to {\ensuremath{U^{\rm pl}}}{\ensuremath{F_{\rm pl}}}$ and
$\epsilon : {\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} \to 1_{\cat{Operad}}$ as follows:
\begin{eqnarray}
\eta_\Phi(\phi) & = &\phi \cmp{|, \dots, |} \\
\epsilon_P(|) & = &1_P \\
\epsilon_P(\phi\cmp{\tau_1, \dots, \tau_n)}
& = & \phi \cmp {\epsilon_P(\tau_1), \dots, \epsilon_P(\tau_n)}
\end{eqnarray}
where $P \in \cat{Operad}, \Phi \in \cat{Set}^\natural, \phi \in \Phi$, and $\tau_1, \dots, \tau_n$ are arrows of $P$.
In other words, $\epsilon_P$ is given by applying composition in $P$ to the
formal tree ${\ensuremath{F_{\rm pl}}} {\ensuremath{U^{\rm pl}}} P$.
\begin{lemma}
$({\ensuremath{F_{\rm pl}}}, {\ensuremath{U^{\rm pl}}}, \eta, \epsilon)$ is an adjunction.
\end{lemma}
\begin{proof}
We proceed by checking the triangle identities.
We require to show that
\begin{eqnarray}
\label{eqn:triarrFp}
\xymatrix{
{\ensuremath{F_{\rm pl}}} \ar[r]^{{\ensuremath{F_{\rm pl}}}\eta} \ar[dr]_{1_{{\ensuremath{F_{\rm pl}}}}}
& {\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}}{\ensuremath{F_{\rm pl}}} \ar[d]^{\epsilon{\ensuremath{F_{\rm pl}}}} \\
& {\ensuremath{F_{\rm pl}}}
} \\
\label{eqn:triarrUp}
\xymatrix{
{\ensuremath{U^{\rm pl}}} \ar[r]^{\eta{\ensuremath{U^{\rm pl}}}} \ar[dr]_{1_{{\ensuremath{U^{\rm pl}}}}}
& {\ensuremath{U^{\rm pl}}}{\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} \ar[d]^{{\ensuremath{U^{\rm pl}}}\epsilon} \\
& {\ensuremath{U^{\rm pl}}}
}
\end{eqnarray}
commute.
For (\ref{eqn:triarrFp}), we proceed by induction on trees.
We shall suppress all subscripts on natural transformations in the interest of
legibility.
For the base case:
\begin{eqnarray*}
\epsilon{\ensuremath{F_{\rm pl}}} ( {\ensuremath{F_{\rm pl}}}\eta (|)) & = & \epsilon{\ensuremath{F_{\rm pl}}} (| \cmp |) \\
& = & | \cmp{\epsilon(|)} \\
& = & | \\
& = & 1_{\ensuremath{F_{\rm pl}}}(|).
\end{eqnarray*}
For the inductive step, let $\Phi$ be a signature, $\phi$ be an $n$-ary\ element of $\Phi$, and $\tau_1, \dots, \tau_n$ be trees labelled by $\Phi$.
Then:
\begin{eqnarray*}
(\epsilon{\ensuremath{F_{\rm pl}}}) ( ({\ensuremath{F_{\rm pl}}}\eta) (\phi \cmp{\tau_1, \dots, \tau_n} ))
& = & \epsilon{\ensuremath{F_{\rm pl}}} (\phi \cmp{\tau_1, \dots, \tau_n}\cmp{|, \dots, |}) \\
& = & \phi \cmp{\tau_1, \dots, \tau_n}
\cmp{(\epsilon{\ensuremath{F_{\rm pl}}})(|), \dots, (\epsilon{\ensuremath{F_{\rm pl}}})(|)} \\
& = & \phi \cmp{\tau_1, \dots, \tau_n} \\
& = & 1_{\ensuremath{F_{\rm pl}}}(\phi \cmp{\tau_1, \dots, \tau_n})
\end{eqnarray*}
Hence $\epsilon{\ensuremath{F_{\rm pl}}} \fcomp {\ensuremath{F_{\rm pl}}}\eta = 1{\ensuremath{F_{\rm pl}}}$, as required.
For (\ref{eqn:triarrUp}), let $P$ be a plain operad, and let $p$ be an $n$-ary\
arrow in $P$.
\begin{eqnarray*}
({\ensuremath{U^{\rm pl}}}\epsilon) ((\eta{\ensuremath{U^{\rm pl}}}) (p)) & = & {\ensuremath{U^{\rm pl}}}\epsilon (p\cmp{|, \dots, |}) \\
& = & p\cmp{({\ensuremath{U^{\rm pl}}}\epsilon)(|), \dots, ({\ensuremath{U^{\rm pl}}}\epsilon)(|)} \\
& = & p\cmp{1, \dots, 1} \\
& = & p \\
& = & 1_{\ensuremath{U^{\rm pl}}}(p)
\end{eqnarray*}
So ${\ensuremath{U^{\rm pl}}}\epsilon \fcomp \eta{\ensuremath{U^{\rm pl}}} = 1_{\ensuremath{U^{\rm pl}}}$, as required.
\end{proof}
\def\calS \times{\ensuremath{\mathcal{S}} \times}
\index{\ensuremath{\mathcal{S}}!relation to ${\ensuremath{F_{\Sigma}^{\rm pl}}}$}
We now consider the ``free symmetric operad'' functor ${\ensuremath{F_{\Sigma}^{\rm pl}}}$.
We shall explicitly define a functor $\calS \times - : \cat{Operad} \to \cat{$\Sigma$-Operad}$ and
show that it is left adjoint to ${\ensuremath{U^{\Sigma}_{\rm pl}}}$, and hence isomorphic to ${\ensuremath{F_{\Sigma}^{\rm pl}}}$.
If $P$ is a plain operad, an element of $(\calS \times P)_n$ is a pair $(\sigma, p)$,
where $p \in P_n$ and $\sigma \in S_n$; i.e., $(\calS \times P)_n = S_n \times P_n$.
Composition is given as follows:
\[
(\sigma, p) \cmp {(\tau_1, q_1), \dots, (\tau_n, q_n)}
= (\sigma \cmp{\tau_1, \dots, \tau_n}, p \cmp{q_{\sigma(1)}, \dots,
q_{\sigma(n)}})
\]
The symmetric group action is given by $\rho \cdot (\sigma, p) = (\rho
\sigma, p)$.
\begin{figure}[h]
\centerline{
\epsfxsize=5in
\epsfbox{freesym.eps}
}
\caption{Composition in $\calS \times P$}
\label{fig:freesym}
\end{figure}
\begin{lemma}
\label{lem:Stimes=Fsigp}
$(\calS \times -) $ is left adjoint to ${\ensuremath{U^{\Sigma}_{\rm pl}}}$.
The unit of the adjunction is given by
\begin{eqnarray*}
\eta: 1 & \to & {\ensuremath{U^{\Sigma}_{\rm pl}}} (\calS \times -) \\
\eta_P: p & \mapsto & (1, p),
\end{eqnarray*}
and the counit is given by
\begin{eqnarray*}
\epsilon : (\calS \times -) {\ensuremath{U^{\Sigma}_{\rm pl}}} & \to & 1 \\
\epsilon_P : (\sigma, p) & \mapsto & \sigma \cdot p.
\end{eqnarray*}
\end{lemma}
\begin{proof}
As before, we proceed by checking the triangle identities.
First, let $P$ be a plain operad, $p$ be an $n$-ary\ arrow in $P$, and $\sigma
\in S_n$.
Then $(\sigma, p)$ is an element of $\calS \times P$.
\begin{eqnarray*}
(\epsilon(\calS \times -)) (\calS \times \eta) (\sigma, p))
& = & (\epsilon(\calS \times -)) (\sigma , (1, p)) \\
& = & \sigma \cdot (1, p) \\
& = & (\sigma \fcomp 1, p) \\
& = & (\sigma, p)
\end{eqnarray*}
Now let $P'$ be a symmetric operad, and $p'$ be an $n$-ary\ arrow in $P'$.
\begin{eqnarray*}
({\ensuremath{U^{\Sigma}_{\rm pl}}}\epsilon) (\eta{\ensuremath{U^{\Sigma}_{\rm pl}}}(p')) & = & ({\ensuremath{U^{\Sigma}_{\rm pl}}}\epsilon)((1, p')) \\
& = & 1 \cdot p' \\
& = & p'
\end{eqnarray*}
So the triangle identities are indeed satisfied, and $(\calS \times -)\dashv {\ensuremath{U^{\Sigma}_{\rm pl}}}$.
\end{proof}
Hence, ${\ensuremath{F_{\Sigma}^{\rm pl}}} = \calS \times -$.
\begin{defn}
\label{def:permtree}
\index{tree!permuted}
Let $\Phi$ be a signature.
An \defterm{$n$-ary\ permuted tree labelled by $\Phi$} is an element of $({\ensuremath{F_{\Sigma}^{\rm pl}}}
{\ensuremath{F_{\rm pl}}} \Phi)_n = ({\ensuremath{F_{\Sigma}}} \Phi)_n$.
\index{tree!finite product}
An \defterm{$n$-ary\ finite product tree labelled by $\Phi$} is an element of
$({\ensuremath{F_{\rm fp}}} \Phi)_n$.
\end{defn}
By Lemma \ref{lem:Stimes=Fsigp}, a permuted tree is a pair $(\sigma, t)$, where
$t \in \mathop{\mbox{tr}}_n \Phi$ and $\sigma \in S_n$, and (by analogous reasoning) an $n$-ary\
finite product tree is a pair $(f, t)$, where $t \in \mathop{\mbox{tr}}_m \Phi$ and $f: \fs{m} \to
\fs{n}$.
\section{Syntactic characterization of the forgetful functors}
There is also a syntactic characterization of the forgetful functor ${\ensuremath{U^{\Sigma}_{\rm pl}}}$.
Given a symmetric operad $P$, we take the signature given by all operations in
$P$ (in other words, the signature ${\ensuremath{U^{\Sigma}}} P$).
We then impose all the plain-operadic equations that are true in $P$, and
take the plain operad corresponding to this strongly regular theory.
This operad is ${\ensuremath{U^{\Sigma}_{\rm pl}}} P$.
We start by making this precise.
\begin{defn}
\label{def:sreqn}
\index{equation!plain-operadic}
Let $\Phi$ be a signature.
A \defterm{plain-operadic equation over $\Phi$ in $n$ variables} is an element
of $(\nterms \Phi)^2$ (that is, a pair of $n$-ary\ strongly regular trees over
$\Phi$), and a \defterm{plain-operadic equation over $\Phi$} is an element of
$\alleqns \Phi$.
\end{defn}
We shall show that a plain-operadic equation over $\Phi$ is the same thing as a strongly regular equation over $\Phi$.
\begin{defn}
\label{def:opd_presentation}
\index{presentation!for a plain operad}
Let $P$ be a plain operad.
A \defterm{presentation} for $P$ is a signature $\Phi$, a signature $E$, and
maps $e_1, e_2 : {\ensuremath{F_{\rm pl}}} E \to {\ensuremath{F_{\rm pl}}} \Phi $ such that, for some $\phi$,
\[
\NoCompileMatrices
\fork {{\ensuremath{F_{\rm pl}}} E} {e_1} {e_2} {{\ensuremath{F_{\rm pl}}} \Phi} \phi P
\]
is a coequalizer.
We say that a regular epi $\phi : {\ensuremath{F_{\rm pl}}} \Phi \to P$ \defterm{generates} $P$, or
that $\phi$ (or, where the choice of $\phi$ is clear, $\Phi$) is a
\defterm{generator} of $P$.
\end{defn}
Presentations and generators for symmetric and finite product operads\ are defined analogously.
We will see how these ``presentations'' are related to presentations of
algebraic theories in Section \ref{sec:synclass}.
We now wish to describe the family of all strongly regular equations that are
true in a given symmetric operad $P$: we will then show that this, together
with the signature given by ${\ensuremath{U^{\Sigma}}} P$, is a presentation for ${\ensuremath{U^{\Sigma}_{\rm pl}}} P$ as
claimed.
\begin{defn}
Let $P$ be a plain operad, and $\phi : {\ensuremath{F_{\rm pl}}} \Phi \to P$ be a generator for $P$.
Let $E$ be a subsignature of $({\ensuremath{U^{\rm pl}}} {\ensuremath{F_{\rm pl}}} \Phi)^2$, so that each $E_n$ is a set of
$n$-ary\ $\Phi$-equations.
Let $i$ be the inclusion map $\xymatrix{E \incar[r] & ({\ensuremath{U^{\rm pl}}} {\ensuremath{F_{\rm pl}}} \Phi)^2 }$, and
$\pi_1, \pi_2$ be the projection maps $({\ensuremath{U^{\rm pl}}} {\ensuremath{F_{\rm pl}}} \Phi)^2 \to {\ensuremath{U^{\rm pl}}} {\ensuremath{F_{\rm pl}}} \Phi$.
Then $P$ \defterm{satisfies all equations in $E$} if the diagram
\[
\NoCompileMatrices
\fork{{\ensuremath{F_{\rm pl}}} E}{\overline{\pi_1 i}}{\overline{\pi_2 i}}{{\ensuremath{F_{\rm pl}}} \Phi} \phi P
\]
is a fork.
\end{defn}
We say that a symmetric or finite product operad\ satisfies a signature of equations if the
analogous condition holds in \cat{$\Sigma$-Operad}\ or {\cat{FP-Operad}}.
Recall the notion of the ``kernel pair'' of a morphism:
\begin{defn}
\index{kernel pair}
Let $f: A \to B$ in some category $\ensuremath{\mathcal{C}}$.
The \defterm{kernel pair} of $f$ is the pair $\parallelpair W p q A$ of maps in
the pullback square
\[
\NoCompileMatrices
\commsquare {W \ar@{}[dr]|-<{\textstyle{\lrcorner}}} p A q f A f B
\]
if this pullback exists.
\end{defn}
\def\srtrees#1{{\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}}#1}
\begin{lemma}
\label{lem:sreqns}
Let $\epsilon$ be the counit of the adjunction ${\ensuremath{F_{\rm pl}}} \dashv {\ensuremath{U^{\rm pl}}}$.
Let $\parallelpair Q {\pi_1} {\pi_2} {{\ensuremath{F_{\rm pl}}}{({\ensuremath{U^{\Sigma}}} P)}}$ be the kernel
pair of the component
\[
\epsilon_{{\ensuremath{U^{\Sigma}_{\rm pl}}} P} : {\ensuremath{F_{\rm pl}}} {\ensuremath{U^{\Sigma}}} P = {\ensuremath{F_{\rm pl}}} {\ensuremath{U^{\rm pl}}} {\ensuremath{U^{\Sigma}_{\rm pl}}} P \to {\ensuremath{U^{\Sigma}_{\rm pl}}} P,
\]
of $\epsilon$.
Let $h$ be the unique map $Q \to ({\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\Sigma}}} P)^2$ induced by $\pi_1, \pi_2$.
Then the image of ${\ensuremath{U^{\rm pl}}} h$ is the largest signature of plain-operadic ${\ensuremath{U^{\Sigma}}}
P$-equations satisfied by $P$.
\end{lemma}
\begin{proof}
$Q, \pi_1, \pi_2$ are given by the diagram
\[
\NoCompileMatrices
\commsquare {Q \ar@{}[dr]|-<{\textstyle{\lrcorner}}} {\pi_1} {{\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\Sigma}}} P}
{\pi_2} \epsilon
{{\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\Sigma}}} P} \epsilon {{\ensuremath{U^{\Sigma}_{\rm pl}}} P}
\]
As a right adjoint, ${\ensuremath{U^{\rm pl}}}$ preserves pullbacks; we take the standard
construction of pullbacks in $\cat{Set}^\natural$ as subobjects of products, in
which case $h$ is an inclusion map.
An element of $({\ensuremath{U^{\rm pl}}} Q)_n$ is then a pair $(e_1, e_2)$ of $n$-ary\ strongly regular
${\ensuremath{U^{\Sigma}}} P$ trees such that $\epsilon(e_1) = \epsilon(e_2)$.
Hence, $Q$ is a signature of plain-operadic ${\ensuremath{U^{\Sigma}}} P$-equations satisfied by
$P$.
Conversely, let $E$ be a signature of plain-operadic ${\ensuremath{U^{\Sigma}}} P$-equations
satisfied by $P$, and let $(e_1, e_2)$ be an element of $E_n$:
then $\epsilon(e_1) = \epsilon(e_2)$ and so $(e_1, e_2)$ is an element of $({\ensuremath{U^{\rm pl}}}
Q)_n$.
\end{proof}
\def\pibar#1{\overline{{\ensuremath{U^{\rm pl}}} \pi_{#1}}}
\begin{corollary}
\label{cor:srpres}
Let $R$ be the plain operad generated by ${\ensuremath{U^{\Sigma}}} P$, satisfying exactly those
plain-operadic equations satisfied by $P$.
Then
\[
\xymatrixcolsep{4pc}
\parallelpair{{\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} Q}{\pibar 1}{\pibar 2}{{\ensuremath{F_{\rm pl}}} {\ensuremath{U^{\Sigma}}} P}
\]
is a presentation for $R$, where the overbars refer to transposition with
respect to the adjunction ${\ensuremath{F_{\rm pl}}} \dashv {\ensuremath{U^{\rm pl}}}$.
\end{corollary}
We recall some standard results.
\begin{lemma}
\label{lem:mon=>regep}
The counit of a monadic adjunction is componentwise regular epi.
\end{lemma}
\begin{proof}
See \cite{acc} 20.15.
\end{proof}
\begin{lemma}
\label{lem:fork.epi}
If $\fork X f g Y h Z$ is a coequalizer in some category, and if $e: W \to X$
is epi, then $\fork W {fe} {ge} Y h Z$ is a coequalizer.
\end{lemma}
\begin{proof}
Suppose $\fork W {fe} {ge} Y i A$ is a fork.
Then $ife = ige$, so $if = ig$ since $e$ is epi.
So $\fork X f g Y i A$ is a fork, and hence $i$ factors uniquely through $h$.
\end{proof}
\begin{lemma}
\label{lem:coeq-kerpair}
In categories with all kernel pairs, every regular epi is the coequalizer of
its kernel pair.
\end{lemma}
\begin{proof}
Let $\ensuremath{\mathcal{C}}$ have all kernel pairs, and $\fork A f g B e C$ be a coequalizer
diagram in $\ensuremath{\mathcal{C}}$.
Let $\parallelpair W p q B$ be the kernel pair of $e$.
We will show that $\fork W p q B e C$ is a coequalizer diagram.
Since $ef = eg$, we may uniquely factor $(f,g)$ through W:
\[
\xymatrix{
A \ar[ddr]_g \ar[rrd]^f \unar[dr]|i \\
& W \ar@{}[dr]|-<{\textstyle{\lrcorner}} \ar[r]_p \ar[d]^q
& B \ar[d]_e \\
& B \ar[r]^e
& C
}
\]
Suppose $\fork W p q B h D$ is a fork.
$hp = hq$, so $hpi = hqi$, so $hf = hg$.
By the universal property of $e$, we may factor $h$ uniquely through $e$.
So $\fork W p q B e C$ is a coequalizer diagram, as required.
\end{proof}
\def\epsilon_{{\ensuremath{U^{\Sigma}_{\rm pl}}} P}{\epsilon_{{\ensuremath{U^{\Sigma}_{\rm pl}}} P}}
\begin{lemma}
Let $P$ be a symmetric operad, and let $Q, \pi_1, \pi_2$ be as in Lemma
\ref{lem:sreqns}.
Then the coequalizer of the diagram
\[
\parallelpair Q {\pi_1} {\pi_1} {{\ensuremath{F_{\rm pl}}} {\ensuremath{U^{\Sigma}}} P}
\]
is ${\ensuremath{U^{\Sigma}_{\rm pl}}} P$.
\end{lemma}
\begin{proof}
Let $\epsilon'$ be the unit of the adjunction ${\ensuremath{F_{\Sigma}^{\rm pl}}} \dashv {\ensuremath{U^{\Sigma}_{\rm pl}}}$.
This adjunction is monadic, so $\epsilon_{{\ensuremath{U^{\Sigma}_{\rm pl}}} P}$ is regular epi by Lemma
\ref{lem:mon=>regep}.
By Lemma \ref{lem:coeq-kerpair}, $\epsilon_{{\ensuremath{U^{\Sigma}_{\rm pl}}} P}$ is the coequalizer of its kernel pair, i.e.
\[
\NoCompileMatrices
\fork Q {\pi_1} {\pi_2} {\allterms{{\ensuremath{U^{\Sigma}}} P}} {\epsilon_{{\ensuremath{U^{\Sigma}_{\rm pl}}} P}} {{\ensuremath{U^{\Sigma}_{\rm pl}}} P}
\]
is a coequalizer diagram.
\end{proof}
\begin{theorem}
\index{{\ensuremath{U^{\Sigma}_{\rm pl}}}!explicit construction}
Let $P$ be a symmetric operad.
Then ${\ensuremath{U^{\Sigma}_{\rm pl}}} P$ is the plain operad whose operations are those in $P$,
satisfying exactly those plain-operadic equations which are true in $P$.
\end{theorem}
\begin{proof}
The adjunction ${\ensuremath{F_{\rm pl}}} \dashv {\ensuremath{U^{\rm pl}}}$ is monadic, so if $\epsilon'$ is its
counit, then $\epsilon'_Q : {\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} Q \to Q$ is (regular) epi by Lemma
\ref{lem:mon=>regep}.
Hence, by Lemma \ref{lem:fork.epi},
\def\pieps#1{\pi_{#1} \epsilon'_Q}
\[
\fork {{\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\rm pl}}} Q} {\pieps 1} {\pieps 2} {{\ensuremath{F_{\rm pl}}}{\ensuremath{U^{\Sigma}}} P} {\epsilon_{{\ensuremath{U^{\Sigma}_{\rm pl}}} P}} {{\ensuremath{U^{\Sigma}_{\rm pl}}} P}
\]
is a coequalizer.
But $\pieps 1 = \overline{{\ensuremath{U^{\rm pl}}} \pi_1}$, and similarly
$\pieps 2 = \overline{{\ensuremath{U^{\rm pl}}} \pi_2}$.
Hence, by Corollary \ref{cor:srpres}, ${\ensuremath{U^{\Sigma}_{\rm pl}}} P$ is the plain operad generated by ${\ensuremath{U^{\Sigma}}} P$, satisfying all plain-operadic equations true in $P$.
Since ${\ensuremath{U^{\Sigma}}} P = {\ensuremath{U^{\rm pl}}} {\ensuremath{U^{\Sigma}_{\rm pl}}} P$, the $n$-ary\ operations of ${\ensuremath{U^{\Sigma}_{\rm pl}}} P$ are exactly
the $n$-ary\ operations of $P$.
\end{proof}
We may generalize this as follows:
\begin{theorem}
Let $\ensuremath{\mathcal{C}}$ be a category with pullbacks, and $T$ be a monad on $\ensuremath{\mathcal{C}}$.
Let $(\halg T A a) \in \ensuremath{\mathcal{C}}^T$.
Let $\parallelpair E {\phi_1} {\phi_2} {TA}$ be the kernel pair of $a$ in $\ensuremath{\mathcal{C}}$.
Then
\[
\fork {F_T E} {\bar \phi_1} {\bar \phi_2} {F_T A} {\epsilon_{(A,a)}} {(A,a)}
\]
is a coequalizer in $\ensuremath{\mathcal{C}}^T$, where $F_T : \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{C}}^T$ is the free functor, and $\epsilon$ is the counit of the adjunction $F_T \dashv U_T$.
\end{theorem}
\begin{proof}
As above.
\end{proof}
\begin{corollary}
\index{{\ensuremath{U^{\rm fp}_{\Sigma}}}!explicit construction}
Let $P$ be a finite product operad.
Then ${\ensuremath{U^{\rm fp}_{\Sigma}}} P$ is the symmetric operad whose operations are given by those of
$P$, satisfying all linear equations that are true in $P$, and ${\ensuremath{U^{\rm fp}_{\rm pl}}} P$ is the
plain operad whose operations are given by those in $P$, satisfying all
strongly regular equations that are true in $P$.
\end{corollary}
\section{Operads and syntactic classes of theories}
\label{sec:synclass}
We have defined notions of algebras for plain, symmetric and finite product operads.
We might ask how these are related to the algebraic theories of Chapter
\ref{ch:theories}: are the algebras for an operad $P$ algebras for some
algebraic theory $\mathcal{T}_P$?
If so, what can we say about the theories that so arise?
We will show the following:
\begin{itemize}
\item Plain operads are equivalent in expressive power to \emph{strongly
regular} theories.
\item Symmetric operads are equivalent in expressive power to \emph{linear}
theories.
\item Finite product operads\ are equivalent in expressive power to general algebraic theories.
\end{itemize}
The first equivalence is proved in \cite{hohc}.
The second has long been folklore (see, for instance, \cite{baez_ualg} page
50), but as far as I know no proof has appeared before.
An (independently found) proof does appear in an unpublished paper of Ad\'amek
and Velebil, who also consider the enriched case.
\index{Ad\'amek, Ji\v r\'i} \index{Velebil, Ji\v r\'i}
The third equivalence was proved in two stages by Tronin, in \cite{tronin} and
\cite{tronin2}.
Recall the definitions of strongly regular and linear terms from
Definition \ref{def:sr_linterms}, and the definitions of strongly regular,
permuted and finite product trees (Definitions \ref{def:srterm} and
\ref{def:permtree}).
Let $\Phi$ be a signature.
We will show that there is an isomorphism between the set $({\ensuremath{T_{\rm fp}}} \Phi)_n$ and
the set of $n$-ary\ words in $\Phi$, and that this isomorphism restricts to
further isomorphisms as follows:
\begin{equation}
\label{eq:termtypes}
\xymatrix{
({\ensuremath{T_{\rm fp}}} \Phi)_n & \cong
& \{\mbox{$n$-ary\ words in $\Phi$}\} \\
({\ensuremath{T_{\Sigma}}} \Phi)_n \incar[u] & \cong
& \{\mbox{$n$-ary\ linear words in $\Phi$}\} \incar[u] \\
({\ensuremath{T_{\rm pl}}} \Phi)_n \incar[u] & \cong
& \{\mbox{$n$-ary\ strongly regular words in $\Phi$}\} \incar[u] \\
}
\end{equation}
The maps in the left-hand column can be viewed as inclusions between different
sets of finite product trees, or equivalently as maps arising from the units of
the adjunctions ${\ensuremath{F_{\rm fp}^{\Sigma}}} \dashv {\ensuremath{U^{\rm fp}_{\Sigma}}}$ and ${\ensuremath{F_{\Sigma}^{\rm pl}}} \dashv {\ensuremath{U^{\Sigma}_{\rm pl}}}$.
Let $\Phi$ be a signature.
Observe that trees in $\Phi$ give rise to terms according to the following
recursive algorithm:
\begin{itemize}
\item Let $\tau$ be an $n$-ary\ strongly regular tree, and $Y = (y_1, y_2,
\dots, y_n)$ a finite sequence of variables.
The \defterm{term $\mathop{\rm{term}}(\tau,Y)$ arising from $\tau$ with working alphabet
$Y$} is given as follows:
\begin{itemize}
\item If $\tau = |$, then $\mathop{\rm{term}}(\tau,Y) = y_1$.
\item If $\tau = \phi\ocomp(\tau_1, \dots, \tau_n)$, then
\[
\mathop{\rm{term}}(\tau,Y) =
\phi(\mathop{\rm{term}}(\tau_1, (y_1, \dots, y_{i_1})), \dots, \mathop{\rm{term}}(\tau_n, (y_{1 +
i_{n-1}}, \dots, y_{i_n}))),
\]
where $i_1$ is the arity of $\tau_1$, and $i_j - i_{j-1}$ is the arity of
$\tau_j$ for $j > 1$.
\end{itemize}
\item The \defterm{term $\mathop{\rm{term}}(\tau)$ arising from $\tau$} is $\mathop{\rm{term}}(\tau,
(x_1, x_2, \dots, x_n))$.
\item Let $\sigma\cdot \tau$ be a permuted tree.
Then $\mathop{\rm{term}}(\sigma\cdot \tau) = \mathop{\rm{term}}(\tau, (x_{\sigma 1}, x_{\sigma 2},
\dots, x_{\sigma n}))$.
\item Let $f \cdot \tau$ be a finite product tree.
Then $\mathop{\rm{term}}(f\cdot \tau) = \mathop{\rm{term}}(\tau, (x_{f(1)}, x_{f(2)}, \dots, x_{f(n)})$
\end{itemize}
\begin{defn}
Let $t$ be a $\Phi$-term.
We define a plain-operadic tree $\mathop{\rm{tree}}(t)$ recursively:
\begin{itemize}
\item if $t$ is a variable, let $\mathop{\rm{tree}}(t) = |$.
\item if $t = \phi (t_1, \dots, t_n)$, let
$\mathop{\rm{tree}}(t) = \phi \ocomp (\mathop{\rm{tree}}(t_1), \dots, \mathop{\rm{tree}}(t_n))$.
\end{itemize}
\end{defn}
\begin{lemma}
\label{lem:term<->tree}
\index{tree!finite product!equivalence to terms}
Every $\Phi$-term $t$ is equal to $\mathop{\rm{term}}(f \cdot \tau)$ for a unique finite
product tree $(f\cdot \tau)$.
\end{lemma}
\begin{proof}
We will show
\begin{enumerate}
\item
\label{term->tree->term}
if $t$ is a $\Phi$-term, then $\mathop{\rm{term}}(\mathop{\rm{label}}(t)\cdot \mathop{\rm{tree}}(t)) = t$;
\item
\label{tree->term->tree}
if $(f \cdot \tau)$ is a finite product tree, then
$f = \mathop{\rm{label}}(\mathop{\rm{term}}(f \cdot \tau))$ and $\tau = \mathop{\rm{tree}}(\mathop{\rm{term}}(f \cdot \tau))$.
\end{enumerate}
(\ref{term->tree->term}) Let $t$ be a $\Phi$-term.
Let $f = \mathop{\rm{label}}(t)$, and $\tau = \mathop{\rm{tree}}(t)$.
Then $\mathop{\rm{term}}(f \cdot \tau)$ is $\mathop{\rm{term}}(\tau, (x_{f(1)}, \dots, x_{f(n)}))$.
We proceed by induction.
\begin{itemize}
\item if $t = x_i$, then $\mathop{\rm{term}}(f\cdot \tau)$ is $\mathop{\rm{term}}(|, (x_i))$, which is
$x_i$.
\item if $t = \phi (t_1, \dots, t_n)$, where each $t_i$ has arity $k_i$,
then
\begin{eqnarray*}
\mathop{\rm{term}}(f \cdot \tau) &= &\mathop{\rm{term}}(\phi \ocomp (\mathop{\rm{tree}}(t_1), \dots, \mathop{\rm{tree}}(t_n)),
(x_{f(1)}, \dots, x_{f(n)})) \\
& = & \phi (\mathop{\rm{term}}(\mathop{\rm{tree}}(t_1), (x_{f(1)}, \dots, x_{f(k_1)})), \dots, \\
& & \phantom{\phi (} \mathop{\rm{term}}(\mathop{\rm{tree}}(t_n), (x_{f((\sum_{i=1}^{n-1} k_i) + 1)},
\dots, x_{f(\sum_{i=1}^n k_i)}))) \\
& = & \phi (t_1, \dots, t_n) \\
& = & t.
\end{eqnarray*}
\end{itemize}
(\ref{tree->term->tree})
Let $\tau$ be an $n$-ary\ plain-operadic tree in $\Phi$, and $f$ a function of
finite sets with codomain $\fs{m}$.
Let $t = \mathop{\rm{term}}(f \cdot \tau)$.
We proceed as usual by induction on $\tau$.
\begin{itemize}
\item If $\tau = |$, then $t = x_{f(1)}$; then $\mathop{\rm{tree}}(t) = | = \tau$ and
$\mathop{\rm{label}}(t)$ is the function $\fs 1 \to \fs{m}$ sending 1 to $f(1)$, i.e. $\mathop{\rm{label}}(t)
= f$.
\item If $\tau = \phi \ocomp (\tau_1, \dots, \tau_n)$, then
\begin{eqnarray*}
t &= &\mathop{\rm{term}}(\phi \ocomp (\tau_1, \dots, \tau_n), (x_{f(1)},
\dots, x_{f(\sum k_i)})) \\
&= &\phi (\mathop{\rm{term}}(\tau_1,(x_{f(1)}, \dots, x_{f(k_1)})), \dots,
\mathop{\rm{term}}(\tau_n, (x_{f((\sum_{i=1}^{n-1} k_i) + 1)}) \dots,
x_{f(\sum_{i=1}^n k_i)})))
\end{eqnarray*}
By induction, $\mathop{\rm{var}}(t) = (x_{f(1)}, \dots, x_{f(\sum k_i)})$, so $\mathop{\rm{label}}(t) =
f$, and $\mathop{\rm{tree}}(t) = \phi \ocomp(\tau_1, \dots, \tau_n) = \tau$ as required.
\end{itemize}
\end{proof}
We have now established the isomorphism in the top line of \ref{eq:termtypes}.
If we use this isomorphism to identify finite product operads\ with finitary monads on \cat{Set}, we
may view the functor ${\ensuremath{F_{\rm fp}^{\rm pl}}}$ as the well-known functor sending a plain operad
to its associated monad on \cat{Set}.
\begin{lemma}
\label{lem:syntactic_semantic_equiv}
\index{tree!permuted!equivalence to linear terms}
\index{tree!plain-operadic!equivalence to strongly regular terms}
Let $t$ be a $\Phi$-term.
Then $t$ is linear iff $t = \mathop{\rm{term}}(\tau)$ for some permuted tree $\tau$, and
strongly regular iff $t = \mathop{\rm{term}}(\tau)$ for some strongly regular tree $\tau$.
\end{lemma}
\begin{proof}
In Lemma \ref{lem:term<->tree}, we factored every $\Phi$-term $t$ into a
strongly regular tree $\mathop{\rm{tree}}(t)$ and a labelling function $\mathop{\rm{label}}(t)$.
By definition, $t$ is linear iff $\mathop{\rm{label}}(t)$ is a bijection, which occurs iff
$t = \mathop{\rm{term}}(\sigma \cdot \tau)$ for some plain-operadic tree $\tau$ and some
bijection $\sigma$.
Hence, the linear terms and permuted trees are in one-to-one correspondence.
Similarly, strongly regular terms and plain-operadic trees are in one-to-one
correspondence.
\end{proof}
The commutativity of \ref{eq:termtypes} now follows from our explicit
construction of ${\ensuremath{F_{\Sigma}}}$ and ${\ensuremath{F_{\rm pl}}}$ in Section \ref{sec:adj}.
\begin{lemma}
\label{lem:linpresfactor}
Let $(\Phi, E)$ be a presentation of an algebraic theory.
Then $(\Phi, E)$ is linear if and only if the projection maps $\parallelpair E
{\pi_1} {\pi_2} \mathop{\rm{term}}(\Phi)$ may be factored through the map
$\xymatrix{{\ensuremath{T_{\Sigma}}} \Phi \ar[r]^\eta & {\ensuremath{T_{\rm fp}}} \Phi \ar[r]^{\sim} & \mathop{\rm{term}}{\Phi}}$:
\[
\xymatrix{
E \parallelarsdir{\pi_1}{\pi_2}{rr} \uparallelarsdir{}{}{dr}
& & {\ensuremath{T_{\rm fp}}} \Phi \ar[r]^-{\sim} & \mathop{\rm{term}} \Phi \\
& {\ensuremath{T_{\Sigma}}} \Phi \ar[r]^-{\sim} \ar[ur]_{\eta}
& \{\mbox{\rm linear $\Phi$-terms}\} \incar[ur]
}
\]
\end{lemma}
\begin{proof}
By definition, the presentation is linear iff $\pi_1, \pi_2$ factor
through the signature of linear $\Phi$-terms.
By Lemma \ref{lem:syntactic_semantic_equiv}, this signature is isomorphic to
${\ensuremath{T_{\Sigma}}} \Phi$, so we are done.
\end{proof}
\begin{theorem}
Let $Q \in {\cat{FP-Operad}}$.
Then
\begin{enumerate}
\item $Q$ is strongly regular iff there exists a $P \in \cat{Operad}$ such that $Q \cong {\ensuremath{F_{\rm fp}^{\rm pl}}} P$;
\item $Q$ is linear iff there exists a $P \in \cat{$\Sigma$-Operad}$ such that $Q \cong
{\ensuremath{F_{\rm fp}^{\Sigma}}} P$;
\end{enumerate}
\end{theorem}
\begin{proof}
We will consider the linear case; the strongly regular case is proved
analogously.
If $Q$ is linear, then there exists a linear presentation $\parallelpair E {}{}
{{\ensuremath{F_{\rm fp}}} \Phi}$ for $Q$.
We may regard $E$ as a subobject of the signature of $\Phi$-equations.
By assumption, $E$ consists only of linear equations; by Lemma
\ref{lem:syntactic_semantic_equiv}, every $(s,t) \in E$ is $(\mathop{\rm{term}}(\sigma_1
\cdot \tau_1), \mathop{\rm{term}}(\sigma_2 \cdot \tau_2))$ for some pair $(\sigma_1 \cdot
\tau_1, \sigma_2 \cdot \tau_2)$ of permuted trees.
So the diagram $\parallelpair E {} {} {{\ensuremath{F_{\rm fp}}} \Phi}$ in {\cat{FP-Operad}}\ is the image
under ${\ensuremath{F_{\rm fp}^{\Sigma}}}$ of a diagram
\[
\parallelpair {E'} {} {} {{\ensuremath{F_{\Sigma}}} \Phi}
\]
in $\cat{$\Sigma$-Operad}$.
This diagram has a coequalizer: call it $P$.
The functor ${\ensuremath{F_{\rm fp}^{\Sigma}}}$ is a left adjoint, and thus preserves coequalizers:
hence, $Q$ is the image under ${\ensuremath{F_{\rm fp}^{\Sigma}}}$ of $P$.
Now suppose $Q = {\ensuremath{F_{\rm fp}^{\Sigma}}} P$ for some symmetric operad $P$.
We may take the canonical presentation of $P$:
\[
\xymatrixcolsep{4pc}
\fork{{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}}{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} P}{\epsilon{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}}}{{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}}\epsilon}
{{\ensuremath{F_{\Sigma}}} {\ensuremath{U^{\Sigma}}} P}\epsilon P
\]
and apply ${\ensuremath{F_{\rm fp}^{\Sigma}}}$ to it:
\[
\xymatrixcolsep{4pc}
\fork{{\ensuremath{F_{\rm fp}}}({\ensuremath{U^{\Sigma}}}{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} P)}{{\ensuremath{F_{\rm fp}^{\Sigma}}}\epsilon{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}}}
{{\ensuremath{F_{\rm fp}}}{\ensuremath{U^{\Sigma}}}\epsilon}{{\ensuremath{F_{\rm fp}}} {\ensuremath{U^{\Sigma}}} P}{{\ensuremath{F_{\rm fp}^{\Sigma}}}\epsilon}{{\ensuremath{F_{\rm fp}^{\Sigma}}} P = Q}
\]
Since ${\ensuremath{F_{\rm fp}^{\Sigma}}}$ is a left adjoint, it preserves coequalizers, so the transpose
of this parallel pair is a presentation for $Q$.
Take this transpose:
\[
\xymatrixcolsep{4pc}
\xymatrix{
{\ensuremath{U^{\Sigma}}}{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} P
\parallelars{\overline{\epsilon{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}}}}
{\overline{{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}}\epsilon}}
& {\ensuremath{U^{\Sigma}}}{\ensuremath{F_{\Sigma}}}{\ensuremath{U^{\Sigma}}} P \ar[r]^{\eta'}
& {{\ensuremath{U^{\rm fp}}} {\ensuremath{F_{\rm fp}}} {\ensuremath{U^{\Sigma}}} P} \ar[r]^{{\ensuremath{U^{\rm fp}}}{\ensuremath{F_{\rm fp}^{\Sigma}}}\epsilon}
& {\ensuremath{U^{\rm fp}}} Q
}
\]
where $\eta'$ is the unit of the adjunction ${\ensuremath{F_{\rm fp}}} \dashv {\ensuremath{U^{\rm fp}}}$, and the bars
refer to transposition with respect to the adjunction ${\ensuremath{F_{\Sigma}}} \dashv {\ensuremath{U^{\Sigma}}}$.
This is in precisely the form required for Lemma \ref{lem:linpresfactor}.
\end{proof}
\begin{example}
\index{monoids}
\index{pointed sets}
\index{commutative monoids}
The theories of monoids and pointed sets are strongly regular, because the finite product operads\ corresponding to these theories are in the image of ${\ensuremath{F_{\rm fp}^{\rm pl}}}$; the theory of
commutative monoids is linear but not strongly regular, because the finite product operad\ whose
algebras are commutative monoids is in the image of ${\ensuremath{F_{\rm fp}^{\Sigma}}}$ but not in the
image of ${\ensuremath{F_{\rm fp}^{\rm pl}}}$.
\end{example}
There is a little more to be said about these classes of theories.
\begin{defn}
\index{wide pullback}
A \defterm{wide pullback} is a limit of a (possibly infinite) diagram of the
form
\[
\xymatrixrowsep{1pc}
\xymatrixcolsep{3pc}
\xymatrix{
\bullet \ar[ddr] \\
\bullet \ar[dr] \\
\vdots & \bullet \\
\bullet \ar[ur] \\
\vdots
}
\]
\end{defn}
\begin{defn}
\index{cartesian!natural transformation}
A natural transformation $\alpha : F \to G$ is \defterm{cartesian} if every
naturality square
\[
\commsquare {FA} {Ff} {FB} {\alpha_A} {\alpha_B} {GA} {Gf} {GB}
\]
is a pullback square.
\end{defn}
\begin{defn}
\index{cartesian!monad}
A monad $(T, \mu, \eta)$ is \defterm{cartesian} if $T$ preserves pullbacks and
$\mu, \eta$ are cartesian natural transformations.
\end{defn}
\begin{theorem}
\label{thm:opds<=>cart. monad}
A plain operad is equivalent to a cartesian monad on \cat{Set}\ equipped with a
cartesian map of monads to the free monoid monad.
\end{theorem}
\begin{proof}
See \cite{hohc} 6.2.4.
Let 1 be the terminal plain operad; algebras for 1 are monoids.
Since 1 is terminal, every plain operad $P$ comes equipped with a map $!: P \to
1$.
This induces a cartesian map of monads $T_! : T_P \to T_1$, and $T_1$ is the
free monoid monad.
\end{proof}
\begin{lemma}
Let $T, S$ be endofunctors on a category $\Acat$, let $\alpha : T \to S$ be a
cartesian natural transformation, and let $S$ preserve wide pullbacks.
Then $T$ preserves wide pullbacks.
\end{lemma}
\begin{proof}
This follows from the facts that wide pullbacks are products in slice
categories and that the functor $f^* : \Acat/B \to \Acat/A$ induced by a map $f
: A \to B$ is product-preserving.
\end{proof}
\begin{corollary}
\label{cor:wpb}
Let $P$ be a plain operad.
Then the functor part of the monad $T_P$ arising from $P$ preserves wide
pullbacks.
\end{corollary}
\begin{defn}
\index{familial representability}
A functor $F : \ensuremath{\mathcal{C}} \to \cat{Set}$ is \defterm{familially representable} if $F$ is a
coproduct of representable functors.
A monad $(T, \mu, \eta)$ on $\cat{Set}$ is \defterm{familially representable} if $T$
is familially representable.
\end{defn}
\begin{theorem} (Carboni-Johnstone)
Let $\ensuremath{\mathcal{C}}$ be a complete, locally small, well-powered category with a small
cogenerating set, and let $F : \ensuremath{\mathcal{C}} \to \cat{Set}$ be a functor.
The following are equivalent:
\begin{enumerate}
\item $F$ is familially representable;
\item $F$ preserves wide pullbacks.
\end{enumerate}
\end{theorem}
\begin{proof}
See \cite{c+j1}, Theorem 2.6.
\end{proof}
\begin{corollary}
The monad associated to a strongly regular theory is familially representable.
\end{corollary}
However, the inclusion is only one-way: there exist cartesian monads $(T, \mu,
\eta)$ such that $T$ is familially representable but the induced theory
is not strongly regular.
For instance, take the theory of involutive monoids:
\begin{defn}
\index{involutive monoid}
\index{monoid with involution|see{involutive monoid}}
An \defterm{involutive monoid} (or \defterm{monoid with involution}) is a
monoid $(M,.,1)$ equipped with an involution $i : M \to M$,
satisfying $i(a.b) = i(b).i(a)$.
\end{defn}
The theory of involutive monoids is familially representable, but not strongly
regular --- see \cite{c+j2}.
\section{Enriched operads and multicategories}
\label{sec:enriched}
In the previous sections we considered operads $P$ where $P_0, P_1,
\dots \in \cat{Set}$, and composition was given by functions.
It is possible to consider operads where $P_0, P_1, \dots$ lie in some other
category; the resulting objects are called \emph{enriched operads}.
Enriched operads have many applications and a rich theory: for instance,
topologists often consider operads enriched in \cat{Top}\ or in some category of
vector spaces.
Our treatment here will be brief, sufficient only to set up the definitions of
Chapter \ref{ch:categorification}: for more on enriched operads, see
\cite{mss}.
Throughout this section, let $(\ensuremath{\mathcal{V}}, \otimes, I, \alpha, \lambda, \rho, \tau)$ be
a symmetric monoidal category.
\begin{defn}
\index{multicategory!plain!enriched}
A \defterm{$\ensuremath{\mathcal{V}}$-multicategory} $\ensuremath{\mathcal{C}}$ consists of the following:
\begin{itemize}
\item a collection $\ensuremath{\mathcal{C}}_0$ of \emph{objects},
\item for all $n \in \natural$ and all $c_1, \dots, c_n,
d \in \ensuremath{\mathcal{C}}_0$, an object $\ensuremath{\mathcal{C}}(c_1, \dots, c_n; d) \in \ensuremath{\mathcal{V}}$ called the
\emph{arrows} from $c_1, \dots, c_n$ to $d$,
\item for all $n, k_1, \dots k_n \in \natural$ and $c_1^1 \dots, c_{k_n}^n,
d_1, \dots, d_n, e \in \ensuremath{\mathcal{C}}_0$, an arrow in $\ensuremath{\mathcal{V}}$ called \emph{composition}
\[\ocomp
: \ensuremath{\mathcal{C}}(d_1, \dots, d_n;e) \otimes
\ensuremath{\mathcal{C}}(c^1_1, \dots, c^1_{k_1};d_1) \otimes \dots \otimes
\ensuremath{\mathcal{C}}(c^n_1, \dots, c^n_{k_n};d_n)
\to
\ensuremath{\mathcal{C}}(c^1_1, \dots, c^n_{k_n};e)\]
\item for all $c \in \ensuremath{\mathcal{C}}$, a \emph{unit} $u_c: I \to \ensuremath{\mathcal{C}}(c;c)$
\end{itemize}
satisfying the following axioms:
\begin{itemize}
\item \emph{Associativity:} For all $b\udot\ldseq, c\seq\udot, d_\bullet, e \in \ensuremath{\mathcal{C}}$, the
following diagram commutes:
\[
\xymatrixcolsep{1.5pc}
\xymatrix{
*\txt{$\ensuremath{\mathcal{C}}(d_\bullet; e) \otimes \ensuremath{\mathcal{C}}(c^1_\bullet; d_1) \otimes \dots \otimes
\ensuremath{\mathcal{C}}(c^n_\bullet; d_n)$ \\
$\otimes \ensuremath{\mathcal{C}}(b^1_{1\bullet}; c^1_1) \otimes
\dots \otimes \ensuremath{\mathcal{C}}(b^{n}_{k_n\bullet}; c^n_{k_n})$}
\ar[r]^-\ocomp
\ar[dd]^-\ocomp
&
\ensuremath{\mathcal{C}}(d_\bullet; e) \otimes \ensuremath{\mathcal{C}}(b^1_{\bullet\bullet}; d_1) \otimes \dots \otimes
\ensuremath{\mathcal{C}}(b^n_{\bullet\bullet}; d_n)
\ar[dd]^-\ocomp
\\ \\
\ensuremath{\mathcal{C}}(c\seq\udot; e) \otimes \ensuremath{\mathcal{C}}(b^1_{1\bullet}; c^1_1) \otimes
\dots \otimes \ensuremath{\mathcal{C}}(b^{n}_{k_n\bullet}; c^n_{k_n})
\ar[r]^-\ocomp
&
\ensuremath{\mathcal{C}}(b\udot\ldseq ; e)
}
\]
\item \emph{Units:} For all $c_\bullet, d \in \ensuremath{\mathcal{C}}$, the following diagram commutes:
\[
\xymatrixcolsep{1.3pc}
\xymatrix{
\ensuremath{\mathcal{C}}(c_\bullet ; d)
\ar[r]^\lambda
\ar[ddrr]^1
\ar[d]_{\rho^n}
&
I \otimes \ensuremath{\mathcal{C}}(c_\bullet; d)
\ar[dr]^{u \otimes 1}
\\
\ensuremath{\mathcal{C}}(c_\bullet ; d) \otimes I \otimes \dots \otimes I
\ar[dr]_{1\otimes u^{\otimes n}}
&
&
\ensuremath{\mathcal{C}}(d;d) \otimes \ensuremath{\mathcal{C}}(c_\bullet;d)
\ar[d]^{\ocomp}
\\
&
\ensuremath{\mathcal{C}}(c_\bullet ; d) \otimes \ensuremath{\mathcal{C}}(c_1;c_1) \otimes \dots \otimes \ensuremath{\mathcal{C}}(c_n;c_n)
\ar[r]^-{\ocomp}
&
\ensuremath{\mathcal{C}}(c_\bullet ; d)
}
\]
\end{itemize}
(We suppress the symmetry maps in \ensuremath{\mathcal{V}}\ for clarity).
\end{defn}
\begin{defn}
\label{def:symmVmcat}
\index{multicategory!symmetric!enriched}
A \defterm{symmetric \ensuremath{\mathcal{V}}-multicategory} is a \ensuremath{\mathcal{V}}-multicategory \ensuremath{\mathcal{C}}\ and, for every
$n \in \natural$, every $\sigma \in S_n$, and every $a_1, \dots a_n, b \in \ensuremath{\mathcal{C}}$,
an arrow
\[
\sigma \cdot - : \ensuremath{\mathcal{C}}(a_1, \dots, a_n; b) \longrightarrow
\ensuremath{\mathcal{C}}(a_{\sigma 1}, \dots, a_{\sigma n}; b)
\]
in \ensuremath{\mathcal{V}}\ such that
\begin{itemize}
\item for each $n \in \natural$ and each $a_1, \dots, a_n, b \in \ensuremath{\mathcal{C}}$, the arrow
$1_n \act - : \ensuremath{\mathcal{C}}(a_1, \dots, a_n;b) \to \ensuremath{\mathcal{C}}(a_1, \dots, a_n;b)$ is the identity
arrow on $\ensuremath{\mathcal{C}}(a_1, \dots, a_n;b)$,
\item for each $\sigma, \rho \in S_n$,
\[
(\rho \cdot - )(\sigma \cdot -) = (\rho\sigma) \cdot -
\]
\item for each $n, k_1, \dots, k_n \in n$, each $\sigma \in S_n$ and $\rho_i
\in S_{k_i}$ for $i = 1, \dots, n$, and for all $a^1_1, \dots, a^n_{k_n}, b_1,
\dots, b_n, c \in \ensuremath{\mathcal{C}}$, the diagram
\[
\xymatrixcolsep{-5pc}
\xymatrix{
& \ensuremath{\mathcal{C}}(b_1, \dots, b_n; c) \otimes \bigotimes_{i=1}^n{\ensuremath{\mathcal{C}}(a^i_1,\dots, a^i_n; b_i)}
\ar[dl]_{(\sigma\act-) \otimes \bigotimes_{i=1}^n{(\rho_i\act -)}}
\ar[ddr]^\ocomp \\
\ensuremath{\mathcal{C}}(b_{\sigma 1}, \dots, b_{\sigma n};c)\otimes
\bigotimes_{i=1}^n{\ensuremath{\mathcal{C}}(a^i_{\rho_i 1}, \dots, a^i_{\rho_i n};b_i)}
\ar[dd]_{1\otimes \sigma_*} \\
& & \ensuremath{\mathcal{C}}(a^1_1, \dots, a^n_{k_n}; c) \ar[ddl]^{\sigma \ocomp(\rho_1,\dots,\rho_n)}
\\
\ensuremath{\mathcal{C}}(b_{\sigma 1}, \dots, b_{\sigma n};c) \otimes \bigotimes_{i=1}^n \ensuremath{\mathcal{C}}(a^{\sigma
i}_{\rho_{\sigma i} 1}, \dots, a^{\sigma i}_{\rho_{\sigma i} n} ; b_{\sigma i})
\ar[dr]_\ocomp \\
& \ensuremath{\mathcal{C}}(a^{\sigma 1}_{\rho_{\sigma 1} 1}, \dots, a^{\sigma n}_{\rho_{\sigma n}
k_n}; c)
}
\]
commutes, where $\sigma\ocomp(\listn{\rho_#1})$ is as defined in Example
\ref{ex:symmopd}.
\end{itemize}
\end{defn}
In the case $\ensuremath{\mathcal{V}} = \cat{Set}$ (with the cartesian monoidal structure), this is
equivalent to Definition \ref{def:symm_mcat}.
Let \ensuremath{\bbF}\ be a skeleton of the category of finite sets and functions, with
objects the sets \fs 0, \fs 1, \fs 2, \dots, where $\fs n = \{1,2,\dots,n\}$.
\begin{defn}
\index{multicategory!finite product!enriched}
A \defterm{finite product \V-multicategory} is
\begin{itemize}
\item A plain \ensuremath{\mathcal{V}}-multicategory \ensuremath{\mathcal{C}};
\item for every function $f : \fs{n} \to \fs{m}$ in \ensuremath{\bbF}, and for all objects $C_1,
\dots, C_n, D \in \ensuremath{\mathcal{C}}$, a morphism $f\act - : \ensuremath{\mathcal{C}}(C_1, \dots, C_n;D) \to
\ensuremath{\mathcal{C}}(C_{f(1)}, \dots, C_{f(n)}; D)$ in \ensuremath{\mathcal{V}}\,
\end{itemize}
satisfying the conditions given in Definition \ref{def:symmVmcat}, where
$(f\cmp{f_1,\dots, f_n})$ is as given in Definition \ref{dopdef}.
\end{defn}
In the case $\ensuremath{\mathcal{V}} = \cat{Set}$, this is equivalent to Definition \ref{dopdef}.
\begin{defn}
\index{operad!enriched}
A (\defterm{plain}, \defterm{symmetric}, \defterm{finite product})
\defterm{\ensuremath{\mathcal{V}}-operad} is a (plain, symmetric, finite product) \ensuremath{\mathcal{V}}-multicategory
with only one object.
\end{defn}
\begin{defn}
\index{morphism!of enriched multicategories}
Let $\ensuremath{\mathcal{C}}$, $\ensuremath{\mathcal{D}}$ be plain \ensuremath{\mathcal{V}}-multicategories.
A \defterm{morphism} $F: \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{D}}$ is
\begin{itemize}
\item for each object $C \in \ensuremath{\mathcal{C}}$, a choice of object $FC \in \ensuremath{\mathcal{D}}$,
\item for each $n \in \natural$ and all collections of objects $A_1, \dots,
A_n, B \in \ensuremath{\mathcal{C}}$, an arrow
$ \ensuremath{\mathcal{C}}(A_1, \dots, A_n;B) \to \ensuremath{\mathcal{D}}(FA_1, \dots, FA_n; FB) $ in $\ensuremath{\mathcal{V}}$
\end{itemize}
such that
\begin{itemize}
\item for all $A \in \ensuremath{\mathcal{C}}$, the diagram
\[
\xymatrix{
& I \ar[ddl]_u \ar[ddr]^u \\ \\
\ensuremath{\mathcal{C}}(A;A) \ar[rr]^F & & \ensuremath{\mathcal{D}}(FA;FA)
}
\]
commutes;
\item for all $n, k_1, \dots, k_n \in \natural$ and all $C, B_1, \dots, B_n, A^1_1, \dots, A^n_{k_n} \in \ensuremath{\mathcal{C}}$, the diagram
\[
\xymatrixrowsep{4pc}
\xymatrix{
\ensuremath{\mathcal{C}}(B_\bullet;C) \otimes \bigotimes_{i=1}^n \ensuremath{\mathcal{C}}(A^i_\bullet;B_i)
\ar[r]^-{\ocomp} \ar[d]_{F\otimes \dots \otimes F}
& \ensuremath{\mathcal{C}}(A\seq\udot;C) \ar[d]^F \\
\ensuremath{\mathcal{D}}(FB_\bullet;FC)\otimes \bigotimes_{i=1}^n \ensuremath{\mathcal{D}}(F A^i_\bullet;FB_i)
\ar[r]^-\ocomp
& \ensuremath{\mathcal{D}}(FA\seq\udot;FC)
}
\]
commutes.
\end{itemize}
\end{defn}
Suppose that $\ensuremath{\mathcal{V}}$ is cocomplete.
Let $Q$ be a (plain, symmetric, finite product) $\ensuremath{\mathcal{V}}$-operad, and $A$ an object of $\ensuremath{\mathcal{V}}$.
Let $Q \kel A$ denote the coend
\[
\int^{n\in \ensuremath{\mathcal{C}}} Q_n \times A^n
\]
where \ensuremath{\mathcal{C}}\ is
\begin{itemize}
\item the discrete category on $\natural$ if $Q$ is a plain operad;
\item a skeleton $\Bcat$ of the category of finite sets and bijections if $Q$
is a symmetric operad;
\index{\ensuremath{\Bcat}}
\item a skeleton $\ensuremath{\bbF}$ of the category of finite sets and all functions if $Q$
is a finite product operad.
\end{itemize}
This notation is taken from Kelly's papers \cite{kelly} and \cite{kellycalc} on
clubs.
\index{Kelly, Max}
The various endomorphism operads defined in Examples \ref{ex:plEnd},
\ref{ex:symmEnd} and \ref{ex:fpEnd} transfer straightforwardly to the
\ensuremath{\mathcal{V}}-enriched setting.
An \defterm{algebra} for a (plain, symmetric, finite product) \ensuremath{\mathcal{V}}-operad $P$ in
a (plain, symmetric, finite product) \ensuremath{\mathcal{V}}-multicategory $\ensuremath{\mathcal{C}}$ is an object $A \in
\ensuremath{\mathcal{C}}$ and a morphism $(\hat{\phantom{\alpha}}) : P \to \mathop{\rm{End}}(A)$ of the appropriate type.
\index{algebra!for an enriched operad}
Equivalently, an algebra for $P$ in $\ensuremath{\mathcal{C}}$ is an object $A \in \ensuremath{\mathcal{C}}$ and a morphism
$h: P \kel A \to A$ such that the diagram
\[
\xymatrix{
P \kel P \kel A \ar[r]^{1 \kel h} \ar[d]_{\ocomp} & P \kel A \ar[d]^h \\
P \kel A \ar[r]^h & A
}
\]
commutes, and $h(1_P, -)$ is the identity on $A$.
\begin{remark}
\index{multicategory!internal}
\index{operad!internal}
There is another possibility, that of considering \emph{internal}
multicategories in the category \ensuremath{\mathcal{V}}, which gives a different notion: now
$\ensuremath{\mathcal{C}}_0$ is an object in \ensuremath{\mathcal{V}}\ rather than a collection.
An internal operad in \ensuremath{\mathcal{V}}\ is an internal multicategory \ensuremath{\mathcal{C}}\ such that $\ensuremath{\mathcal{C}}_0$ is
terminal in \ensuremath{\mathcal{V}}.
We shall not consider internal multicategories or operads further.
\end{remark}
We shall in particular consider the case $\ensuremath{\mathcal{V}} = \cat{Cat}$, and \cat{Cat}-operads again
have a simple concrete description:
\begin{lemma}
\index{operad!enriched!in \cat{Cat}}
A {(plain) \cat{Cat}-operad} $Q$ is a sequence of categories $Q_0, Q_1,
\dots$, a family of composition functors $\ocomp:Q_n \times Q_{k_1}
\times \ldots \times Q_{k_n} \to Q_{\sum k_i}$ and an identity
$1_Q \in Q_1$, satisfying (strict) functorial versions of the axioms given in
\ref{lem:operad_description}.
\end{lemma}
\begin{lemma}
A {symmetric \cat{Cat}-operad} is a plain \cat{Cat}-operad $Q$ with a left group
action of each symmetric group $S_n$ on the corresponding category $Q_n$,
strictly satisfying equations as in Definition \ref{def:symmopd}.
\end{lemma}
\begin{lemma}
A {finite product \Cat-operad} is a plain \cat{Cat}-operad $Q$ equipped with functors $f \act - :
Q_n \to Q_m$ for each function $f: \fs{n} \to \fs{m}$ of finite sets, strictly
satisfying equations as in Definition \ref{dopdef}.
\end{lemma}
All of these lemmas can be established by a straightforward check of
the definitions.
Just as 2-category theory has a special flavour distinct from the theory of
$\ensuremath{\mathcal{V}}$-categories in the case $\ensuremath{\mathcal{V}}=\cat{Cat}$, so the theories of $\cat{Cat}$-operads and
$\cat{Cat}$-multicategories have unique features:
\begin{defn}
\label{laxmorphdef}
\index{morphism!of algebras for a finite product \Cat-operad}
Let $Q$ be a finite product \Cat-operad, and let $Q \kel A \stackrel\alpha\to A , Q \kel B
\stackrel\beta\to B$ be algebras for $Q$ in $\cat{Cat}$. A
\defterm{lax morphism of $Q$-algebras} $A \to B$ consists of a 1-cell $F: A \to
B$ and a 2-cell $\phi: \beta F \to F \alpha$ satisfying the
following conditions:
\begin{equation}
\label{eq:laxmorph1}
\xymatrixrowsep{3pc}
\xymatrix{
A \ar[r]^F \ar[d]_\eta
\ddlowertwocell<-20>_1{=}
\drtwocell\omit{=}
& B \ar[d]^\eta
\dduppertwocell<20>^1{=} \\
Q\kel A \ar[r]^{1\kel F} \ar[d]_\alpha
\drtwocell\omit{\phi}
& Q\kel B \ar[d]^\beta
& & = & A \rtwocell^F_F{=} & B \\
A \ar[r]_F & B
}
\end{equation}
\begin{equation}
\label{eq:laxmorph2}
\xymatrixcolsep{1.4pc}
\xymatrixrowsep{3pc}
\xymatrix{
Q \kel Q \kel A \ar[d]_\mu \ar[r]^{1\kel 1\kel F}
\drtwocell\omit{=}
& Q \kel Q \kel B \ar[d]^\mu
& & & Q \kel Q \kel A \ar[dl]_\mu \ar[r]^{1\kel 1\kel F}
\ar[d]_{1\kel \alpha}
\drtwocell\omit{\phi}
& Q \kel Q \kel B \ar[dr]^\mu \ar[d]^{1\kel \beta} \\
Q \kel A \ar[r]^{1\kel F} \ar[d]_{\alpha}
\drtwocell\omit{\phi}
& Q \kel B \ar[d]^{\beta}
& =
& Q \kel A \ar[dr]_{\alpha}
\rtwocell\omit{=}
& Q \kel A \ar[r]^{1\kel F} \ar[d]_{\alpha}
\drtwocell\omit{\phi}
& Q \kel B \ar[d]^{\beta}
\rtwocell\omit{=}
& Q \kel B \ar[dl]^{\beta} \\
A \ar[r]_F
& B
& & & A \ar[r]_F
& B
}
\end{equation}
\begin{equation}
\label{eq:laxmorph3}
\xymatrixcolsep{1.4pc}
\xymatrixnocompile{
& Q_m \times A^n \ar[dl]_{1 \times f^*} \ar[dr]^{1 \times F^n}
\ddtwocell\omit{=}
& & & & Q_m \times A^n \ar[dl]_{1 \times f^*} \ar[dr]^{1 \times F^n}
\ar[dd]^{f_* \times 1}
\\
Q_m \times A^m \ar[dd]^\alpha \ar[dr]^{1 \times F^m}
& & Q_m \times B^n \ar[dl]_{1 \times f^*} \ar[dd]^{f_* \times 1}
& & Q_m \times A^m \ar[dd]^\alpha
\drtwocell\omit{=}
& & Q_m \times B^n \ar[dd]^{f_* \times 1} \\
\drtwocell\omit{\phi}
& Q_n \times B^m \ar[dd]^\beta
\drtwocell\omit{=}
& & =
& & Q_n \times A^n \ar[dl]_\alpha \ar[dr]_{1 \times F^n}
\urtwocell\omit{=}
\ddtwocell\omit{\phi} \\
A \ar[dr]_F
& & Q_n \times B^n \ar[dl]^\beta
& & A \ar[dr]_F
& & Q_n \times B^n \ar[dl]^\beta \\
& B
& & & & B
}
\end{equation}
for all functions $f : \fs{m} \to \fs{n}$.
A morphism $(F,\phi)$ is \defterm{weak} if $\phi$ is invertible, and
\defterm{strict} if $\phi$ is an identity.
\index{weak!morphism of algebras for a finite product \Cat-operad}
\index{strict!morphism of algebras for a finite product \Cat-operad}
\index{morphism!of algebras for a symmetric \cat{Cat}-operad}
\index{lax!morphism of algebras for a symmetric \cat{Cat}-operad}
\index{weak!morphism of algebras for a symmetric \cat{Cat}-operad}
\index{strict!morphism of algebras for a symmetric \cat{Cat}-operad}
Lax morphisms for algebras of plain \cat{Cat}-operads are required to satisfy
\ref{eq:laxmorph1} and \ref{eq:laxmorph2}, and lax morphisms for algebras of
symmetric \cat{Cat}-operads are required to satisfy \ref{eq:laxmorph1},
\ref{eq:laxmorph2} and the restriction of \ref{eq:laxmorph3} to the case where
$f$ is a bijection.
\end{defn}
We shall make use of a more explicit formulation in the plain case.
\begin{lemma}
\label{lem:wkfunc_explicit}
Let $Q$ be a plain $\cat{Cat}$-operad, and let $(A,h)$ and $(B,h')$ be $Q$-algebras.
A lax map of $Q$-algebras $(A,h) \to (B,h')$ is a pair $(G, \psi)$,
where $G:A \to B$ is a functor and $\psi$ is a sequence of natural
transformations $\psi_i : h'_i (1 \times G^i) \to G h_i$, called the
\defterm{coherence maps}, such that the following equation holds, for all $n,
k_1, \dots, k_n \in \natural$:
\[
\xymatrixrowsep{4pc}
\xymatrixcolsep{4pc}
\xymatrix{
Q_n \times \prodkn Q \times A^{\sum k_i}
\ar[d]_{1\times 1^n \times G^{\sum k_i}}
\ar[r]^-{\prodkn{h}}
\drtwocell\omit{^*{!(-1,-1.5)\object{\prodkn{\psi}}}}
& Q_n \times A^n
\ar[d]|{1\times G^n}
\ar[r]^{h_n}
\drtwocell\omit{^*{!(-1,-1.5)\object{\psi_n}}}
& A
\ar[d]^G
\\
Q_n \times \prodkn Q \times A^{\sum k_i}
\ar[r]_-{\prodkn{h'}}
& Q_n \times B^n
\ar[r]_{h'}
& B \\
}
\hskip 1in
\]
\begin{equation}
\label{weakfunctordef}
\hskip 2in
\midlabel{=}
\xymatrix{
Q_n \times \prodkn Q \times A^{\sum k_i}
\ar[d]_{1\times 1^n \times G^{\sum k_i}}
\ar[r]^-{h_{\sum k_i}}
\drtwocell\omit{^*{!(-1.5,-2.5)\object{\psi_{\sum k_i}}}}
& A
\ar[d]^G
\\
Q_n \times \prodkn Q \times B^{\sum k_i}
\ar[r]_-{h'_{\sum k_i}}
& B \\
}
\end{equation}
and the diagram
\begin{equation}
\label{weakfunctordef2}
\xymatrixrowsep{3pc}
\xymatrixcolsep{3pc}
\xymatrix{
G a \ar[r]^{\delta'_1} \ar[d]_1
& h'(1_P, G a) \ar[d]^{\psi_1} \\
G a \ar[r]_{G \delta_1}
& G h(1_P, a)
}
\end{equation}
commutes.
The morphism is weak if every $\psi$ is invertible, and
strict if every $\psi$ is an identity.
\end{lemma}
\begin{proof}
This can be established by a straightforward check of the definition.
\end{proof}
\begin{defn}
\label{transfdef}
\index{transformation!between morphisms of algebras for a \cat{Cat}-operad}
Let $Q, A, B$ etc.\ be as above, and let $(F,\phi), (G,\gamma)$ be lax
morphisms of $Q$-algebras $A \to B$.
A \defterm{$Q$-transformation} $F \to G$ is a natural transformation $\sigma:
F \to G$ such that
\begin{equation}
\xymatrixrowsep{1pc}
\xymatrix{
Q \kel A \rtwocell^{1\kel F}_{1\kel G}{\sigma} \ar[dd]_\alpha
\ddrtwocell\omit{\gamma}
& Q \kel B \ar[dd]^\beta
& & Q \kel A \ar[r]^{1 \kel F} \ar[dd]_\alpha
\ddrtwocell\omit{\phi}
& Q \kel B \ar[dd]^\beta \\
& & = \\
A \ar[r]_G
& B
& & A \rtwocell^F_G{\sigma}
& B
}
\end{equation}
\end{defn}
\begin{lemma}
\label{inv <=> Q-inv}
A $Q$-transformation $\sigma: (F, \phi) \to (G, \psi)$ is invertible as a
natural transformation if and only if it is invertible as a $Q$-transformation.
\end{lemma}
\begin{proof}
``If'' is obvious: we concentrate on ``only if''. It is enough to show that
$\sigma^{-1}$ is a $Q$-transformation, which is to say that
\begin{equation}
\xymatrix{
h(q, Ga\seq)
\ar[r]^\psi
\ar[d]_{h(q,\sigma^{-1}_{a\seq})} &
Gh(q,a\seq)
\ar[d]^{\sigma^{-1}_{h(q,a\seq)}} \\
h(q, Fa\seq)
\ar[r]^\phi &
Fh(q,a\seq)
}
\end{equation}
commutes for all $(q,a\seq) \in Q \kel A$, and this follows
from the fact that ${\sigma_{h(q,a\seq)}} \fcomp \phi = \psi \fcomp
{h(q,\sigma_{a\seq})}$.
\end{proof}
Finite product \Cat-operads, their morphisms and transformations form a 2-category
called {\cat{\Cat-FP-Operad}}.
Similarly, there is a 2-category $\cat{Cat-Operad}$ of plain \cat{Cat}-operads, their
morphisms and transformations, and a 2-category $\cat{Cat-$\Sigma$-Operad}$, of symmetric
operads, their morphisms and transformations.
\begin{theorem}
There is a chain of monadic adjunctions
\index{adjunctions!for \cat{Cat}-operads}
\begin{equation}
\label{eqn:cat_adjs}
\xymatrixrowsep{3pc}
\xymatrix{
{\cat{\Cat-FP-Operad}} \ar@<1.2ex>[d]^{\ensuremath{U^{\rm fp}_{\Sigma}}}
\ar@/r1.7cm/[dd]^{\ensuremath{U^{\rm fp}_{\rm pl}}}
\ar@/r3cm/[ddd]^{\ensuremath{U^{\rm fp}}} \\
\cat{Cat-$\Sigma$-Operad} \ar@<1.2ex>[d]^{\ensuremath{U^{\Sigma}_{\rm pl}}} \ar@<1.2ex>[u]^{\ensuremath{F_{\rm fp}^{\Sigma}}}_{\dashv}
\ar@/r1.7cm/[dd]^{\ensuremath{U^{\Sigma}}} \\
\cat{Cat-Operad} \ar@<1.2ex>[u]^{\ensuremath{F_{\Sigma}^{\rm pl}}}_{\dashv} \ar@<1.2ex>[d]^{\ensuremath{U^{\rm pl}}}
\ar@/l1.7cm/[uu]^{\ensuremath{F_{\rm fp}^{\rm pl}}} \\
\cat{Cat}^\natural \ar@<1.2ex>[u]^{\ensuremath{F_{\rm pl}}}_{\dashv} \ar@/l1.7cm/[uu]^{\ensuremath{F_{\Sigma}}}
\ar@/l3cm/[uuu]^{\ensuremath{F_{\rm fp}}}
}
\end{equation}
\end{theorem}
\begin{proof}
This follows from Lemmas \ref{lem:frees_exist} and \ref{lem:monadj}, via an
application of the argument of Theorem \ref{thm:opd monadic over setN}.
\end{proof}
Since operads can be considered as one-object multicategories, a \cat{Cat}-operad
$P$ (of whatever type) is really a 2-dimensional structure.
We will therefore refer to the objects and morphisms of the categories $P_i$
as \defterm{1-cells} and \defterm{2-cells} of $P$, respectively.
\index{1- and 2-cells}
\section{Maps of algebras as algebras for a multicategory}
\label{sec:Pbar}
Let $P$ be a plain operad.
We form a multicategory $\bar P = \cat{2} \times P$, where $\cat{2}$ is
the category $(\xymatrix{\cdot\ar[r] &\cdot})$.
We may describe $\bar P$ as follows: there are two objects, labelled
0 and 1; the hom-sets $\bar P(0,\dots, 0; 0)$ and $\bar P(x_1, \dots, x_n; 1)$
are copies of $P_n$, for $x_i \in \{0,1\}$, and $\bar P(x_1, \dots, x_n; 0) =
\emptyset$ if any of the $x_i$s are 1.
Composition is given by composition in $P$.
An algebra for $\bar P$ is a pair $A_0, A_1$ of $P$-algebras, and a morphism of $P$-algebras $A_0 \to A_2$.
See \cite{markl}, Example 2.4 for more details.
We can extend this construction by defining a multicategory $\bar{\bar P} =
\cat{3} \times P$, whose algebras are composable pairs of maps of
$P$-algebras, a multicategory $\bar {\bar {\bar P}} = \cat{4} \times P$
whose algebras are composable triples of maps of $P$-algebras, and so on.
With the obvious face and degeneracy maps, these multicategories form a
cosimplicial object in the category of plain multicategories.
The same construction can be performed for symmetric and enriched operads, and
the result continues to hold.
\chapter{Other Approaches}
\label{ch:others}
\section{Pseudo-algebras for 2-monads}
We begin by recalling some standard notions of 2-monad theory.
\begin{defn}
\index{2-monad}
A \defterm{2-monad} is a monad object in the 2-category of 2-categories, in the
sense of \cite{street}; that is to say, a 2-category $\ensuremath{\mathcal{C}}$, a strict 2-functor
$T: \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{C}}$, and 2-transformations $\mu : T^2 \to T$ and $\eta : 1_\ensuremath{\mathcal{C}} \to T$
satisfying the usual monad laws:
\begin{equation}
\xymatrix{
& T^3 \ar[dr]^{\mu T} \ar[dl]_{T \mu} \\
T^2 \ar[dr]_{\mu} & & T^2 \ar[dl]^\mu \\
& T
}
\end{equation}
\begin{equation}
\xymatrix{
T \ar[r]^{T\eta} \ar[dr]_{1_T}
& T^2 \ar[d]^\mu
& T \ar[l]_{\eta T} \ar[dl]^{1_T} \\
& T
}
\end{equation}
\end{defn}
As is common for ordinary 1-monads, we will usually refer to a 2-monad $(\ensuremath{\mathcal{C}}, T,
\mu, \eta)$ as simply $T$.
The usual notion of an algebra for a monad carries over simply to this case:
\begin{defn}
\index{algebra!for a 2-monad}
Let $(\ensuremath{\mathcal{C}}, T, \mu, \eta)$ be a 2-monad.
A \defterm{strict algebra} for $T$ is an object $A \in \ensuremath{\mathcal{C}}$ and a 1-cell $a : T
A \to A$ satisfying the following axioms:
\begin{equation}
\xymatrix{
& T^2 A \ar[dr]^{\mu} \ar[dl]_{T a} \\
T A \ar[dr]_a & & T A \ar[dl]^a \\
& A
}
\end{equation}
\begin{equation}
\xymatrix{
A \ar[r]^{\eta} \ar[dr]_{1_A}
& T A \ar[d]^a \\
& A
}
\end{equation}
\end{defn}
For our purposes, it is more interesting to consider the well-known
\defterm{pseudo-algebras} for a 2-monad.
These are algebras ``up to isomorphism'':
\begin{defn}
\index{pseudo-algebra!for a 2-monad}
Let $(\ensuremath{\mathcal{C}}, T, \mu, \eta)$ be a 2-monad.
A \defterm{pseudo-algebra} for $T$ is an object $A \in \ensuremath{\mathcal{C}}$, a 1-cell $a: TA
\to A$, and invertible 2-cells
\[
\xymatrixnocompile{
& T^2 A \ar[dr]^{\mu} \ar[dl]_{T a}
\ddtwocell\omit{^*{!(0, -0.5)\object{\alpha}}} \\
T A \ar[dr]_a & & T A \ar[dl]^a \\
& A
}
\hskip 2cm
\xymatrixnocompile{
A \ar[r]^{\eta} \ar[dr]_{1_A}
\drtwocell\omit{^<-1.7>*{!(0.5,0.5)\object{\beta}}}
& T A \ar[d]^a \\
& A
}
\]
satisfying the equations
\begin{equation}
\xymatrixnocompile{
& T^3 A \ar[dl]_{T^2 a} \ar[dr]^{\mu T} \ar[dd]^{T \mu}
& & & & T^3 A \ar[dl]_{T^2 a} \ar[dr]^{\mu T} \ddtwocell\omit{=} \\
T^2 A \ar[dd]_{Ta}
& {} \drtwocell\omit{=} & T^2 A \ar[dd]^{\mu}
& & T^2 A \ar[dd]_{Ta} \ar[dr]_{\mu}
& & T^2 A \ar[dd]^{\mu} \ar[dl]^{Ta} \\
\urtwocell\omit{*{!(-1, -1.5)\object{T \alpha}}}
& T^2 A \ar[dl]_{T a} \ar[dr]_{\mu}
& {} & =
& & T A \ar[dd]^a
& \\
T A \ar[dr]_a
& & T A \ar[dl]^a
& & T A \ar[dr]_a
& \ultwocell\omit{\alpha} \urtwocell\omit{\alpha}
& T A \ar[dl]^a \\
& A \uutwocell\omit{\alpha}
& & & & A
}
\end{equation}
\begin{equation}
\xymatrix{
& T A \ar@/u0.8cm/[ddl]_1 \ar@/u0.8cm/[ddr]^1 \ar[d]|{T \eta}
& & & TA \ar@/l0.5cm/[dd]_a \ar@/r0.5cm/[dd]^a \ddtwocell\omit{=}
& & & TA \ar[dl]_a \ar@/ur1cm/[ddrr]^1 \ar[dr]^{\eta T} \ddtwocell\omit{=} \\
& T^2 A \ar[dl]_{T a} \ar[dr]^\mu
& & =
& & =
& A \ar[dr]^\eta \ar@/dl1cm/[ddrr]_1
\drlowertwocell\omit{^<1.7>*{!(-0.5,-0.5)\object{\beta}}}
& & T^2 A \ar[dr]^\mu \ar[dl]_{T a}\\
TA \ar[dr]_a
& & TA \ar[dl]^a
& & A
& & & TA \ar[dr]_a
& & TA \ar[dl]^a\\
& A \uutwocell\omit{\alpha}
& & & & & & & A \uutwocell\omit{\alpha}
}
\end{equation}
\end{defn}
\begin{defn}
\index{pseudo-algebra!for a 2-monad!pseudo-morphisms}
Let $(\ensuremath{\mathcal{C}}, T, \mu, \eta)$ be a 2-monad, and let $(A, a, \alpha_1, \alpha_2)$ and
$(B, b, \beta_1, \beta_2)$ be pseudo-algebras for $T$.
A \defterm{pseudo-morphism of pseudo-algebras} $(A, a, \alpha_1, \alpha_2)$ to
$(B, b, \beta_1, \beta_2)$ is a pair $(f, \phi)$, where $f : A \to B$ is a
1-cell in $\ensuremath{\mathcal{C}}$ and $\phi$ is an invertible 2-cell:
\[
\xymatrix{
TA \ar[r]^{Tf} \ar[d]_a \drtwocell\omit{\phi}
& TB \ar[d]^b \\
A \ar[r]_f & B
}
\]
satisfying the axioms
\begin{eqnarray}
\xymatrixrowsep{3pc}
\xymatrixcolsep{3pc}
\xymatrix{
T^2 A \ar[r]^{T^2 f} \ar[d]_{\mu_A}\drtwocell\omit{=}
& T^2 B \ar[d]^{\mu_B} \\
T A \ar[r]^{Tf} \ar[d]_a \drtwocell\omit{\phi}
& T B \ar[d]^b \\
A \ar[r]_f
& B
}
\xymatrix{ {} \\ = }
\xymatrix{
T^2 A \ar[r]^{T^2 f} \ar[d]_{T a}
\drtwocell\omit{*{!(-2,0)\object{T \phi}}}
& T^2 B \ar[d]^{T b} \\
T A \ar[r]^{Tf} \ar[d]_a \drtwocell\omit{\phi}
& T B \ar[d]^b \\
A \ar[r]_f
& B
}
\\
\xymatrixrowsep{3pc}
\xymatrixcolsep{3pc}
\xymatrix{
A \ar[r]^f \ar[d]_{\eta_A}
\ddlowertwocell<-13>_{1_A}{*{!(0,1)\object{\alpha_2^{-1}}}}
\drtwocell\omit{=}
& B \ar[d]^{\eta_B}
\dduppertwocell<13>^{1_B}{*{!(0,1)\object{\beta_2^{-1}}}} \\
T A \ar[r]^{T f} \ar[d]_a \drtwocell\omit{\phi}
& T B \ar[d]^b \\
A \ar[r]_f
& B
}
\xymatrix{ {} \\ = }
\xymatrix{
\\
A \rtwocell^f_f{*{!(-0.5,0)\object{1_f}}} & B
}
\end{eqnarray}
\end{defn}
This gives rise to a category $\PsAlg{T}$ for any 2-monad $T$.
\index{$\PsAlg{T}$}
Every cartesian monad $T$ on \cat{Set}\ gives rise to a 2-monad $\bar T$ on \cat{Cat}\ in
an obvious way, and (as we saw in Theorem \ref{thm:opds<=>cart. monad}) every
plain operad $P$ gives rise to a cartesian monad $T_P$ on \cat{Set}.
So an alternative definition of ``weak $P$-category'' might be ``pseudo-algebra
for $\bar T_P$''.
In order to explore the connections between this idea and the notion of weak
$P$-category given in previous chapters, we shall need some theorems from
\cite{bkp} and related papers.
\begin{theorem} (Blackwell, Kelly, Power)
\label{thm:ps-classifier}
Let $T$ be a 2-monad with rank on a cocomplete 2-category $K$, let $\Alg T\strictness{str}$
be the 2-category of strict $T$-algebras and strict morphisms, and $\Alg T\strictness{wk}$
be the 2-category of strict $T$-algebras and weak morphisms.
Then the inclusion $J : \Alg T\strictness{str} \to \Alg T\strictness{wk}$ has a left adjoint $L$.
Thus every strict $T$-algebra $A$ has a \defterm{pseudo-morphism classifier} $p
: A \to A'$ (where $A' = JLA$), such that for all $B \in K$, and every
pseudo-morphism $f: A \to B$, we may express $f$ uniquely as the composite of
$p$ and a strict morphism:
\[
\xymatrix{
A \ar[r]^p \ar[dr]_f & A' \unar[d]^{J\bar f} \\ & B
}
\]
\end{theorem}
\begin{proof}
See \cite{bkp}, Theorem 3.13.
\end{proof}
\begin{theorem} (Blackwell, Kelly, Power)
\label{thm:f* adj}
Let $f:S \to T$ be a strict map between 2-monads with rank on a cocomplete
2-category $K$.
Then the induced map $f^* : \Alg T\strictness{str} \to \Alg S\strictness{str}$ has a left adjoint, and
the induced map $f^* : \Alg T\strictness{wk} \to \Alg S\strictness{wk}$ has a left biadjoint.
\end{theorem}
\begin{proof}
See \cite{bkp}, Theorem 5.12.
\end{proof}
\begin{corollary}
\label{cor:ps-alg adj}
Composing this left adjoint with the left adjoint of Theorem
\ref{thm:ps-classifier} gives us an adjunction
\[
\NoCompileMatrices
\adjunction{\Alg S\strictness{wk}}{\Alg T\strictness{str}} F U
\]
\end{corollary}
\begin{theorem} (Power, Lack)
\label{thm:ps-classifier as boff}
Let $T$ be a 2-monad with rank on a cocomplete 2-category $K$ of the form
$\cat{Cat}^X$ for some set $X$, and let $T$ preserve pointwise
bijectivity-on-objects.
Let $(A, a)$ be a strict $T$-algebra.
Then the pseudo-morphism classifier $A'$ for $A$ may be found by factorizing the
structure map $a : TA \to A$ as a pointwise bijective-on-objects map followed
by a locally full-and-faithful map:
\[
\xymatrix{
TA \ar[rr]^a \booar[dr] & & A \\ & A' \lffar[ur]
}
\]
\end{theorem}
\begin{proof}
The construction is given in Power's paper \cite{power}, and the universal
property of the algebra constructed is proved in Lack's paper \cite{lack}.
\end{proof}
This argument is due to Steve Lack (private communication).
\begin{theorem}
\label{thm:ps-algs are weak algs}
Let $P$ be a plain operad.
Let $T_P$ be the monad induced by $P$ on \cat{Set}.
Then a pseudo-algebra for $\bar T_P$ is a weak $P$-category in the sense of
Definition \ref{def:weakening}.
Furthermore, there is an isomorphism of categories $\PsAlg{\bar T_P} \cong
\Algwk{P}$.
\end{theorem}
\begin{proof}
\cat{Cat-Operad}\ is monadic over $\cat{Cat}^\natural$ via one of the special monads of
Theorem \ref{thm:ps-classifier as boff}, and hence, for every plain
$\cat{Cat}$-operad $P$, the pseudo-morphism classifier of $P$ is none other than
$\Wk{P}$.
Hence, if $A$ is a category, then a strict map of \cat{Cat}-operads $\Wk{P} \to
\mathop{\rm{End}}(A)$ is precisely a weak map $P \to \mathop{\rm{End}}(A)$, or equivalently a
$\bar T_P$-pseudo-algebra structure on $A$.
\end{proof}
We may also use these ideas to provide a simple proof of the strictification
result in Theorem \ref{mainthm}.
The map $P \to \Wk{P}$ given by Theorem \ref{thm:ps-classifier} is pseudo, but it
has a strict retraction $q: \Wk{P} \to P$.
This is equivalent to a strict map of monads $T_\Wk{P} \to T_P$.
By Corollary \ref{cor:ps-alg adj}, this induces a 2-functor $\Alg P\strictness{str} \to \Alg
\Wk{P}\strictness{wk}$ with a left adjoint.
This functor is simply the inclusion of the 2-category of strict
$P$-categories, strict $P$-functors and $P$-transformations into the 2-category
of weak $P$-(categories, functors, transformations), and its left adjoint is
the functor ${\rm\textbf{st} }$ constructed in Section \ref{sec:strictification}.
The fact that any weak $P$-category $A$ is equivalent to $\st A$ is a
consequence of the fact that any pseudo $P$-algebra is equivalent to a strict
one, and this holds by the General Coherence Result of Power.
However, pseudo-algebras are less useful in the case of linear theories.
Since the monads arising from symmetric operads are not in general cartesian,
we may not perform the construction given above.
We may, however, use the existence of colimits in \cat{Cat}, and consider the
2-monad
\[
A \mapsto \int^{n \in \bbB} P_n \times A^n
\]
for any symmetric operad $P$.
If $P$ is the free symmetric operad on a plain operad $P'$, this 2-monad is
equal to $\bar T_{P'}$.
Yet this coend construction also leads to problems.
Let $T$ be the ``free commutative monoid'' monad on \cat{Set}, and $S$ be the ``free
monoid'' monad on \cat{Set}.
Since these both arise from symmetric operads, we may lift them to 2-monads
$T', S'$ on \cat{Cat}\ as described above.
$T'$ is the free commutative monoid monad on \cat{Cat}, which is to say the free
strict symmetric monoidal category 2-monad; similarly, $S'$ is the free strict
monoidal category 2-monad.
For each category $A$, there is a functor $\pi_A: S'A \to T'A$ which is full
and surjective-on-objects; hence, if $(A, a, \alpha_1,\alpha_2)$ is a
pseudo-algebra for $T'$, we obtain an $S'$-pseudo-algebra structure by
precomposing with
$\pi_A$:
\[
\xymatrix{
S' A \ear[d]^{\pi_A} \\
T' A \ar[d]^a \\
A
}
\hskip 1cm
\xymatrix{
& S'^2 A \ear[d]^{(\pi * \pi)_A} \\
& T'^2 A \ar[dl]_{T'a} \ar[dr]^{\mu} \ddtwocell\omit{^\alpha_1} \\
T'A \ar[dr]_a & & T'A \ar[dl]^a \\
& A
}
\hskip 1cm
\xymatrix{
A \ar[d]_{1_A} \ar[r]^\eta & S'A \ear[d]^{\pi_A} \\
A \ar[r]^\eta \ar[dr]_{1_A}
\druppertwocell\omit{^<-2>*{!(1,0)\object{\alpha_2}}}
& T'A \ar[d]^a \\
& A
}
\]
The $S'$-pseudo-algebra structure so obtained is uniquely determined.
Since $\pi_A$ is full and surjective-on-objects, it is epic, and so every
pseudo-algebra for $S'$ is a pseudo-algebra for $T'$ in at most one way.
Hence we may view all pseudo-algebras for $T'$ as pseudo-algebras for $S'$
(that is, as monoidal categories) with extra properties.
But there exist monoidal categories with several choices of symmetric structure on them.
For instance, consider the category of graded Abelian groups, with tensor
product $(A\otimes B)_n = \bigoplus_{i + j = n} A_i \otimes B_j$.
As well as the obvious symmetry, there is another given by $\tau_{AB}(a\otimes
b) = (-1)^{ij} b\otimes a$, where $a \in A_i, b \in B_j$.
We can say at least something about the extra properties that pseudo-algebras
for $T'$ must have:
\begin{theorem}
\index{commutative monoids}
A pseudo-algebra for $T'$ is a symmetric monoidal category $A$ in which $x
\otimes y = y \otimes x$ for all $x, y \in A$.
\end{theorem}
\begin{proof}
Recall our construction of the finite product operad\ whose algebras are commutative monoids in
Example \ref{ex:comm_monoid_fp}.
From this, we may deduce that if $A$ is a set, then an element of $T_P A$ is a
function $A \to \natural$ assigning each element of $A$ its multiplicity: in
other words, a multiset of elements of $A$.
Let $(A, a, \alpha, \beta)$ be a pseudo-algebra for $T'$ in \cat{Cat}.
Then we have a binary tensor product:
\[
x\otimes y := a(x^1y^1)
\]
where $x^1y^1$ is the function $A \to \natural$ sending $x$ and $y$ to 1 and
all other objects of $A$ to 0.
The tensor is defined analogously on morphisms.
The components of $\alpha$ and $\beta$ give us associator, symmetry and unit
maps, and it can be shown that they satisfy the axioms for a monoidal category.
However, since the function $x^1y^1$ is equal to the function $y^1x^1$ for all
$x, y \in A$, it must be the case that $x \otimes y = y \otimes x$.
\end{proof}
Since not all symmetric monoidal categories satisfy this condition, it is
apparent that a na\"\i ve approach to categorification based on pseudo-algebras
is doomed to fail, and that more sophistication is required.
In fact, I conjecture that a stronger condition holds: that the symmetry maps
are all identities.
In the specific case of symmetric monoidal categories, we may remedy the
situation as follows. \index{symmetric monoidal category}
Let $T$ be the ``free symmetric strict monoidal category'' 2-monad.
Then pseudo-algebras for $T$ are precisely symmetric monoidal categories.
\section{Laplaza sets}
This notion was introduced by T. Fiore, P. Hu and I. Kriz in
\cite{fhk_laplaza}, as a generalization of Laplaza's categorification of rigs
in \cite{laplaza}.
It was introduced as an attempt to correct an error in the earlier definition
of categorification proposed in \cite{fiore}; the error in question is
essentially that discussed in Section \ref{sec:evaluation} above.
\index{Fiore, Tom}\index{Hu, Po}\index{Kriz, Igor}\index{Laplaza, Miguel}
\begin{defn}
\index{Laplaza set}
Let $T$ be a finite product operad.
A \defterm{Laplaza set} for $T$ is a subsignature of ${\ensuremath{U^{\rm fp}}} T$.
\end{defn}
Concretely, a Laplaza set $S$ for $T$ is a sequence $S_0 \subset T_0, S_1
\subset T_1, \dots$ of subsets of $T_0, T_1, \dots$.
\begin{defn}
\index{pseudo-algebra!for a finite product operad\ with Laplaza set}
Let $T$ be a finite product operad, and $S$ be a Laplaza set for $T$.
A \defterm{$(T,S)$-pseudo algebra} is
\begin{itemize}
\item a category \ensuremath{\mathcal{C}}\,
\item for each $\phi \in T_n$, a functor $\hat \phi : \ensuremath{\mathcal{C}}^n \to \ensuremath{\mathcal{C}}$,
\item \defterm{coherence morphisms} witnessing all equations that are true in
$T$,
\end{itemize}
such that, if
\begin{itemize}
\item $s_1, s_2, t_1$ and $t_2$ are elements of $({\ensuremath{F_{\rm fp}}}{\ensuremath{U^{\rm fp}}} T)_n$,
\item $\delta_1: \hat s_1 \to \hat t_1$ and $\delta_2: \hat s_2 \to \hat t_2$
are coherence morphisms,
\item $\epsilon(s_1) = \epsilon(s_2) \in S$ and $\epsilon(t_1) = \epsilon(t_2)
\in S$,
\end{itemize}
then $\delta_1 = \delta_2$.
\end{defn}
This definition can be recast in terms of strict algebras for a finite product \Cat-operad.
By judicious choice of Laplaza set, one can recover the classical notion of
symmetric monoidal category and Laplaza's categorification of the theory of
rigs.
\section{Non-algebraic definitions}
\index{Rosick\'y, Ji\v r\'i}
\index{Leinster, Tom}
Various definitions have appeared that are inspired by the notions of homotopy
monoids etc. in topology.
In \cite{hty_opd}, Leinster proposes a definition of a ``homotopy $P$-algebra
in $M$'' for any plain operad $P$ and any monoidal category $M$; his shorter
paper \cite{hty_mon} explores this definition in the case $P = 1$.
Related (but more general) is Rosick\'y's work described in \cite{rosicky}.
These definitions stand roughly in relation to ours as do the ``non-algebraic''
definitions of $n$-category in relation to the ``algebraic'' ones: see
\cite{cheng+lauda}.
\chapter{Theories}
\label{ch:theories}
The first step will be to obtain a mathematical description of the notion of an
algebraic theory, of which the familiar theories of groups, rings, modules etc.
are examples.
In this chapter, we present some standard ways of doing this, and prove that
they are equivalent.
The most convenient description for our purposes will be the notion of
\emph{clone}, which appears to have been introduced by Philip Hall in
unpublished lecture notes in the 1960s, and may be found on \cite{cohn} page
132, under the name ``abstract clone''.
The treatment here follows \cite{johnstone}.
The remainder of the material in this chapter is all well-known, and may be
found in e.g. \cite{borceux} chapters 3 and 4, or \cite{a+r} chapter 3.
In the next chapter, we shall describe \emph{operads},
which allow us to capture certain algebraic theories in an especially simple
way, suitable for categorification, and we shall show how operads relate to the
clones described in this chapter.
\section{Syntactic approach}
The most traditional way of formalizing algebraic theories is syntactic.
In this approach, we abstract from the standard ``operations plus equations''
description (used to describe e.g. the theory of groups) to create
\defterm{presentations of algebraic theories}, and define a notion of an
\defterm{algebra} for a presentation.
\begin{defn}
\index{signature}
A \defterm{signature} $\Phi$ is an object of $\cat{Set}^\natural$.
\end{defn}
In other words, a signature is a sequence of sets $\Phi_0, \Phi_1, \Phi_2,
\dots$.
Fix a countably infinite set $X = \{x_1, x_2, \dots, \}$, whose elements we
call \defterm{variables}.
\index{variable}
Throughout, let $\Phi$ be a signature.
\begin{defn}
\label{def:term}
\index{term}
Let $n \in \natural$.
An \defterm{$n$-ary\ $\Phi$-term} is defined by the following inductive clauses:
\begin{itemize}
\item $x_1, x_2, \dots, x_n$ are $n$-ary\ terms.
\item If $\phi \in \Phi_m$ and $t_1, \dots, t_m$ are $n$-ary\ terms, then $\phi
(t_1, \dots, t_m)$ is an $n$-ary\ term.
\end{itemize}
A \defterm{$\Phi$-term} is an $n$-ary\ $\Phi$-term for some $n \in \natural$.
\end{defn}
\begin{defn}
\index{var}
Let $t$ be an $n$-ary\ $\Phi$-term.
Then $\mathop{\rm{var}}(t)$ is the sequence of elements of $\{x_1, \dots, x_n\}$ given as
follows:
\begin{itemize}
\item $\mathop{\rm{var}}(x_i) = (x_i)$,
\item $\mathop{\rm{var}}(\phi (t_1, \dots, t_n)) =
\mathop{\rm{var}}(t_1) \mathrel{++} \mathop{\rm{var}}(t_2) \mathrel{++} \dots \mathrel{++} \mathop{\rm{var}}(t_n)$,
\end{itemize}
where $\mathrel{++}$ is concatenation.
\end{defn}
\begin{defn}
\index{supp}
Let $t$ be an $n$-ary\ $\Phi$-term.
Then $\mathop{\rm{supp}}(t)$, the \defterm{support of $t$}, is the subset of $\{x_1, \dots,
x_n\}$ given as follows:
\begin{itemize}
\item $\mathop{\rm{supp}}(x_i) = \{x_i\}$,
\item $\mathop{\rm{supp}}(\phi (t_1, \dots, t_n)) =
\mathop{\rm{supp}}(t_1) \cup \mathop{\rm{supp}}(t_2) \cup \dots \cup \mathop{\rm{supp}}(t_n)$,
\end{itemize}
\end{defn}
\begin{defn}
\index{labelling function}
Let $t$ be an $n$-ary\ $\Phi$-term, with $\mathop{\rm{var}}(t) = (x_{i_1}, \dots, x_{i_m})$.
The \defterm{labelling function} $\mathop{\rm{label}}(t)$ of $t$ is the function $\fs{m} \to \fs{n}$
sending $j$ to $i_j$.
\end{defn}
\begin{defn}
\index{equation}
An \defterm{$n$-ary\ $\Phi$-equation} is a pair $(s,t)$ of $n$-ary\ $\Phi$-terms.
A \defterm{$\Phi$-equation} is an $n$-ary\ $\Phi$-equation for some $n \in
\natural$.
\end{defn}
\begin{defn}
\label{def:sr_linterms}
\index{linear!term}\index{strongly regular!term}
An $n$-ary\ term $t$ is \defterm{linear} if $\mathop{\rm{label}}(t)$ is a bijection, and
\defterm{strongly regular} if $\mathop{\rm{label}}(t)$ is an identity.
\index{linear!equation}\index{strongly regular!equation}
An equation $(s,t)$ is \defterm{linear} if both $s$ and $t$ are linear, and
\defterm{strongly regular} if both $s$ and $t$ are strongly regular.
\end{defn}
In other words, a term is linear if every variable is used exactly once, and
strongly regular if every variable is used exactly once in the order $x_1,
\dots, x_n$.
Up to trivial relabellings, an equation is linear if every variable is used
exactly once on both sides, though not necessarily in the same order: an
example is the commutative equation $x_1.x_2 = x_2.x_1$.
An equation is strongly regular if every variable is used exactly once in the
same order on both sides.
An example is the associative equation $x_1 . (x_2 . x_3) = (x_1.x_2).x_3$,
though some care is needed.
Strictly, a $\Phi$-equation is a pair $(n, (s,t))$ where $n \in \natural$ and
$s, t$ are $n$-ary\ $\Phi$-terms.
The equation $(3, ((x_1.x_2).x_3, x_1.(x_2.x_3)))$ is strongly regular, but the
equation $(4, ((x_1.x_2).x_3, x_1.(x_2.x_3)))$ is not.
Classically, an $n$-ary\ equation $(s,t)$ is \defterm{regular} \index{regular} if
$\mathop{\rm{label}}(t)$ and $\mathop{\rm{label}}(s)$ are surjections: however, we will not consider
regular equations further.
The term ``linear'' is borrowed from linear logic, and the term ``strongly
regular'' is due to Carboni and Johnstone (from \cite{c+j1}).
\begin{defn}
\index{presentation!of an algebraic theory}
A \defterm{presentation of a (one-sorted) algebraic theory} is
\begin{itemize}
\item a signature $\Phi$,
\item a set $E$ of $\Phi$-equations.
\end{itemize}
\index{generating operation}
Elements of $\Phi_n$ are called ($n$-ary) \defterm{generating operations}.
\end{defn}
\begin{defn}
\index{linear!presentation}
\index{strongly regular!presentation}
Let $P = (\Phi, E)$ be a presentation of an algebraic theory.
$P$ is \defterm{linear} if every equation in $E$ is linear, and
\defterm{strongly regular} if every equation in $E$ is strongly regular.
\end{defn}
We will return to the consideration of linear and strongly regular
presentations once we have defined operads.
\begin{defn}
\index{algebra!for a signature}
Let $\Phi$ be a signature.
An \defterm{algebra} for $\Phi$ is
\begin{itemize}
\item a set $A$,
\item for each $n$-ary\ operation $\phi$, a map $\phi_A : A^n \to A$.
\index{primitive operation}
These are called the \defterm{primitive operations} of the algebra $A$.
\end{itemize}
\end{defn}
Let $\Phi$ be a signature, and $A$ a $\Phi$-algebra.
Each $n$-ary\ $\Phi$-term $t$ gives rise to an \defterm{$n$-ary\
derived operation} \index{derived operation} $t_A :A^n \to A$, defined
recursively as follows:
\begin{itemize}
\item if $t = x_i$, then $t_A$ is projection onto the $i$th factor,
\item if $t = \phi(t_1, \dots, t_m)$, then $t_A$ is the composite
\[
\xymatrix {
A^n \ar[rr]^{((t_1)_A, \dots, (t_m)_A)} && A^m \ar[r]^{\phi_A} & A
}.
\]
\end{itemize}
Let $\mathop{\rm{term}}_n{\Phi}$ denote the set of $n$-ary\ derived operations over $\Phi$.
Then $\mathop{\rm{term}}{\Phi}$ is a signature for every signature $\Phi$.
A morphism of signatures $f : \Phi \to \Psi$ induces a map $\bar f : \mathop{\rm{term}}\Phi
\to \mathop{\rm{term}} \Psi$.
Indeed, $\mathop{\rm{term}}$ is an endofunctor on $\cat{Set}^\natural$, and in Section
\ref{sec:synclass} we shall show that it is actually a monad.
\begin{defn}
\index{algebra!for a presentation}
Let $P = (\Phi, E)$ be a presentation of an algebraic theory.
A \defterm{$P$-algebra} is a $\Phi$-algebra $A$ such that, for every equation
$(s,t)$ in $E$, the derived operations $s_A, t_A$ are equal.
\end{defn}
An algebra for $\Phi$ is an algebra for $(\Phi, \{\})$.
Conversely, every algebra for $(\Phi, E)$ is an algebra for $\Phi$.
\begin{defn}
\index{morphism!of algebras for a signature}
Let $\Phi$ be a signature, and $A$ and $B$ be $\Phi$-algebras.
A \defterm{morphism of $\Phi$-algebras} $f : A \to B$ is a map $f:A \to B$
which commutes with every primitive operation:
\[
\commsquare {A^n} {f^n} {B^n} {\phi_A} {\phi_B} A f B
\]
for every $n \in \natural$ and every $n$-ary\ primitive operation $\phi$.
If $P = (\Phi, E)$ is a presentation, then a \defterm{morphism of
$P$-algebras} is a morphism of $\Phi$-algebras.
\end{defn}
By an easy induction, a morphism of $\Phi$-algebras will commute with every
derived operation too.
Given a presentation $P$, there is a category $\Alg{P}$ \index{$\Alg{P}$} whose
objects are $P$-algebras and whose arrows are $P$-algebra morphisms.
We shall call a category \ensuremath{\mathcal{C}}\ a \defterm{variety of algebras} (or simply a
\defterm{variety}) if \ensuremath{\mathcal{C}}\ is isomorphic to $\Alg P$ for some presentation $P$.
We will need to consider \defterm{closures} of sets of equations; the
idea is that the closure of $E$ contains the members of $E$ and all of their
consequences.
\begin{defn}
\index{grafting!of terms}
Let $t$ be an $n$-ary\ $\Phi$-term, and $t_1, \dots, t_n$ be $\Phi$-terms.
Then the \defterm{graft} $t (t_1, \dots, t_n)$ is the $\Phi$-term defined
recursively as follows.
\begin{itemize}
\item If $t = x_i$, then $t (t_1, \dots, t_n) = t_i$.
\item If $t = \phi(s_1, \dots, s_m)$, where $\phi \in \Phi_m$ and $s_1, \dots,
s_m$ are $n$-ary\ $\Phi$-terms, then
$t (t_1, \dots, t_n)
= \phi ((s_1 (t_1, \dots, t_n)), \dots, s_m (t_1, \dots, t_n))$.
\end{itemize}
\end{defn}
\begin{defn}
\index{closure}
Let $\Phi$ be a signature and $E$ be a set of $\Phi$-equations.
The \defterm{closure} $\bar E$ of $E$ is the smallest equivalence relation on
$\mathop{\rm{term}} \Phi$ which contains $E$ and is closed under grafting of terms:
\begin{itemize}
\item if $(s,t) \in \bar E$, then $(s (t_1, \dots, t_n), t (t_1, \dots, t_n)) \in \bar E$ for all $t_1, \dots, t_n$.
\item if $(s_i, t_i) \in \bar E$ for $i = 1, \dots, n$, then $(t (s_1,
\dots, s_n), t (t_1, \dots, t_n)) \in \bar E$ for all $t$.
\end{itemize}
\end{defn}
\section{Clones}
Clones attempt to capture theories directly: a clone is to a presentation of an
algebraic theory as a group is to a presentation of that group.
\begin{defn}
\label{def:clone}
\index{clone}
A \defterm{clone} $K$ is
\begin{itemize}
\item a sequence of sets $K_0, K_1, \dots$,
\item for all $m,n \in \natural$, a function $\ccomp : K_n \times (K_m)^n \to
K_m$,
\item for each $n \in \natural$ and each $i \in \{1,\dots,n\}$, an
element $\delta^i_n \in K_n$
\end{itemize}
such that
\begin{itemize}
\item for each $f \in K_n$, $g_1,\dots,g_n \in K_m$, $h_1, \dots, h_{m} \in
K_p$,
\[
f \ccomp(g_1 \ccomp (h_1, \dots, h_m), \dots, g_n \ccomp (h_1, \dots,
h_m))
= (f \ccomp (g_1, \dots, g_n)) \ccomp (h_1, \dots, h_{m})
\]
\item for all $n$, all $i \in {1, \dots, n}$ and all $f_1, \dots, f_n \in K_m$,
\[
\delta^i_n \ccomp (f_1, \dots, f_n) = f_i
\]
\item for all $n$ and $f \in K_n$,
\[
f \ccomp (\delta_n^1, \dots, \delta_n^n) = f
\]
\end{itemize}
\end{defn}
\begin{example}
\index{endomorphism clone}
Let $\ensuremath{\mathcal{C}}$ be a finite product category, and $A$ be an object of \ensuremath{\mathcal{C}}.
The \defterm{endomorphism clone} of $A$, $\mathop{\rm{End}}(A)$, is defined as follows:
\begin{itemize}
\item $\mathop{\rm{End}}(A)_n = \ensuremath{\mathcal{C}}(A^n, A)$ for each $n \in \natural$,
\item for all $n \in \natural$ and $i \in \{1, \dots, n\}$, the map
$\delta^i_n$ is the projection of $A^n$ onto its $i$th factor,
\item for all $n, m \in \natural$, all $f \in \mathop{\rm{End}}(A)_n$, and all $g_1, \dots,
g_n \in \mathop{\rm{End}}(A)_m$, the morphism $f\ccomp(g_1, \dots, g_n)$ is the composite
$fh$, where $h$ is the unique arrow $A^m \to A^n$ induced by the maps $g_1,
\dots, g_n$ and the universal property of $A^n$.
\end{itemize}
\end{example}
\begin{defn}
\index{morphism!of clones}
A \defterm{morphism} of clones $f: K \to K'$ is a map in $\cat{Set}^\natural$
which commutes with the composition operations and $\delta$s.
\end{defn}
\begin{defn}
\index{algebra!for a clone}
Let $K$ be a clone, and \ensuremath{\mathcal{C}}\ a finite product category with specified finite
powers.
An \defterm{algebra} for $K$ in \ensuremath{\mathcal{C}}\ is an object $A \in \ensuremath{\mathcal{C}}$ and a morphism of
clones $K \to \mathop{\rm{End}}(A)$.
\end{defn}
Equivalently, an algebra for a clone $K$ in a finite product category \ensuremath{\mathcal{C}}\ with
specified powers is
\begin{itemize}
\item an object $A$ of \ensuremath{\mathcal{C}},
\item for each $n \in \natural$ and each $k \in K_n$, a morphism $\hat k : A^n
\to A$
\end{itemize}
such that
\begin{itemize}
\item for all $n \in \natural$ and all $i \in \{1, \dots n\}$, the morphism
$\widehat \delta^i_n$ is the projection of $A^n$ onto its $i$th factor;
\item for all $n, m \in \natural$, all $f \in K_n$, and all $g_1, \dots, g_n
\in K_m$, the diagram
\[
\xymatrix{
& A^m \ar[ddl]_{\widehat{g_1}} \unar[ddd]^h \ar[ddr]^{\widehat{g_n}}
\ar@/r3cm/[dddd]^{\widehat{f\ccomp(g_1, \dots, g_n)}} \\ \\
A & & A \\
& A^n \ar[ul]^{\widehat \delta^1_n} \ar[ur]_{\widehat \delta^n_n}
\ar[d]^f \\
& A
}
\]
commutes, where $h$ is the unique arrow induced by the universal property of
$A^n$.
\end{itemize}
\begin{defn}
\index{morphism!of algebras for a clone}
Let $A$ and $B$ be algebras for a clone $K$ in a finite product category $\ensuremath{\mathcal{C}}$
with specified finite powers.
A \defterm{morphism} of algebras $A \to B$ is a morphism $F : A \to B$ in \ensuremath{\mathcal{C}}\
such that the diagram
\[
\commsquare{A^n}{F^n}{B^n}{\hat k}{\hat k} A F B
\]
commutes for all $n \in \natural$ and all $k \in K_n$.
\end{defn}
Algebras for a clone and their morphisms form a category: we call this category
$\cat{Alg}_\ensuremath{\mathcal{C}}(K)$, or $\Alg{K}$ in the case where $\ensuremath{\mathcal{C}} = \cat{Set}$.\index{\cat{Alg}}
Clones can be enriched in any finite product category $\ensuremath{\mathcal{V}}$ in an obvious
way: the sequence of sets $K_0, K_1, \dots$ becomes a sequence of objects of
$\ensuremath{\mathcal{V}}$, and so on.
\section{Lawvere theories}
Lawvere theories are a particularly elegant approach to describing algebraic
theories, introduced by Lawvere in his thesis \cite{lawthesis}.
\index{Lawvere, F. William}
Like a clone, a Lawvere theory (sometimes called a \defterm{finite product
theory}) is an object that represents the semantics of the theory directly; in
Lawvere theories, the data are encoded into a category.
Algebras for the theory are then certain functors from the Lawvere theory to
\cat{Set}.
\begin{defn}
\index{Lawvere theory}
A \defterm{Lawvere theory} is a category $\mathcal{T}$ whose objects form a
denumerable set $\{\bf{0}, \bf{1}, \bf{2}, \dots \}$, such that $\bf{n}$ is the
$n$-th power of $\bf{1}$.
\index{morphism!of Lawvere theories}
A \defterm{morphism} of Lawvere theories $\mathcal{T} \to \mathcal{S}$ is an identity-on-objects
functor $\mathcal{T} \to \mathcal{S}$ which preserves projection maps.
The category of Lawvere theories and their morphisms is called $\cat{Law}$.
\index{\cat{Law}}
An \defterm{algebra} for $\mathcal{T}$ is a functor $F:\mathcal{T} \to \cat{Set}$ which preserves
finite products.
\index{algebra!for a Lawvere theory}
A \defterm{morphism} of algebras is a natural transformation.
\index{morphism!of algebras for a Lawvere theory}
The \defterm{category of $\mathcal{T}$-algebras} is the full subcategory of $[\mathcal{T}, \cat{Set}]$
whose objects are finite-product-preserving functors.
\end{defn}
Lawvere theories encode algebraic theories by storing the $n$-ary\ operations
of the theory as morphisms $\bf{n} \to \bf{1}$.
We can consider algebras for Lawvere theories in categories other than $\cat{Set}$:
an algebra for a Lawvere theory $\mathcal{T}$ in a finite product category $\ensuremath{\mathcal{C}}$ is just
a finite-product-preserving functor $\mathcal{T} \to \ensuremath{\mathcal{C}}$.
This captures our usual notions of, for instance, topological groups: a
topological group is just an algebra for the Lawvere theory of groups in the
category $\cat{Top}$.
Much the same could be said for clones and presentations, of course, but in this
case the definition is especially economical.
We may generalize this definition as follows:
\begin{defn}
\index{finite product theory, multi-sorted}
Let $S$ be a set.
An \defterm{$S$-sorted finite product theory} is a small finite product
category whose underlying monoidal category is strict and whose monoid of
objects is the free monoid on $S$.
Elements of $S$ will be called \defterm{sorts}.
Algebras and morphisms of algebras are defined as above.
\end{defn}
\section{Finitary monads}
\index{monad}
Recall that a \defterm{monad} on a category $\ensuremath{\mathcal{C}}$ is a monoid object in the
category $[\ensuremath{\mathcal{C}}, \ensuremath{\mathcal{C}}]$ of endofunctors on $\ensuremath{\mathcal{C}}$.
Concretely, a monad is a triple $(T, \mu, \eta)$ where
\begin{itemize}
\item $T: \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{C}}$ is a functor,
\item $\mu : T^2 \to T$ is a natural transformation,
\item $\eta : 1_\ensuremath{\mathcal{C}} \to T$ is a natural transformation,
\end{itemize}
and $\mu, \eta$ satisfy coherence axioms which are analogues of the usual
associativity and unit laws for monoids, namely
\begin{equation}
\label{eqn:monad_assoc}
\xymatrix{
& T^3 \ar[dl]_{\mu T} \ar[dr]^{T\mu} \\
T^2 \ar[dr]_\mu & & T^2 \ar[dl]^\mu \\
& T
}
\end{equation}
\begin{equation}
\label{eqn:monad_unit}
\xymatrix{
T \ar[r]^{T\eta} \ar[dr]_{1_T}
& T^2 \ar[d]^\mu
& T \ar[l]_{\eta T} \ar[dl]^{1_T} \\
& T
}
\end{equation}
We shall often abuse notation and refer to the monad $(T, \mu, \eta)$ as simply
$T$.
\begin{defn}
\label{def:monad_morphism}
\index{morphism!of monads}
Let $(T_1, \mu_1, \eta_1), (T_2, \mu_2, \eta_2)$ be monads on a category $\ensuremath{\mathcal{C}}$.
A \defterm{morphism of monads} $(T_1, \mu_1, \eta_1) \to (T_2, \mu_2, \eta_2)$
is a natural transformation $\alpha : T_1 \to T_2$ such that the diagrams
\begin{equation}
\label{eqn:monad_morph_mu}
\xymatrix{
T_1^2 \ar[r]^{\mu_1} \ar[d]_{\alpha * \alpha} & T_1 \ar[d]^\alpha \\
T_2^2 \ar[r]_{\mu_2} & T_2
}
\end{equation}
\begin{equation}
\label{eqn:monad_morph_eta}
\xymatrix{
1 \ar[r]^{\eta_1} \ar[dr]_{\eta_2} & T_1 \ar[d]^{\alpha} \\
& T_2
}
\end{equation}
commute.
\end{defn}
Monads on \ensuremath{\mathcal{C}}\ and monad morphisms form a category $\cat{Mnd}(\ensuremath{\mathcal{C}})$. \index{$\cat{Mnd}(\ensuremath{\mathcal{C}})$}
This notion (or rather, a 2-categorical version) was introduced and studied by
Street in \cite{street}. \index{Street, Ross}
\begin{defn}
\index{filtered!category}
A category $\ensuremath{\mathcal{C}}$ is \defterm{filtered} if every finite diagram in $\ensuremath{\mathcal{C}}$ admits a
cocone.
\end{defn}
Equivalently, $\ensuremath{\mathcal{C}}$ is filtered if:
\begin{itemize}
\item $\ensuremath{\mathcal{C}}$ is nonempty;
\item for every pair of parallel arrows $\parallelpair A f g B$ in $\ensuremath{\mathcal{C}}$, there
is an arrow $h: B \to C$ such that $hf = hg$;
\item for every pair of objects $A, B$, there is an object $C$ and arrows
\[
\xymatrix{
A \ar[dr]^f \\
& C \\
B \ar[ur]_g
}
\]
\end{itemize}
Filteredness generalizes the notion of directedness for posets (a
\defterm{directed poset} is a poset in which every finite subset has an upper
bound). \index{directed}
A filtered category which is also a poset is precisely a directed poset.
\begin{defn}
\index{filtered!colimit}
A \defterm{filtered colimit} in a category $\ensuremath{\mathcal{C}}$ is the colimit of a diagram $D: \I \to \ensuremath{\mathcal{C}}$, where $\I$ is a filtered category.
\end{defn}
\begin{theorem}
Every object in \cat{Set}\ is a filtered colimit of finite sets.
\end{theorem}
\begin{proof}
Let $X \in \cat{Set}$, and consider the subcategory $\I$ of $\cat{Set}$ whose objects are
finite subsets of $X$ and whose morphisms are inclusions.
This is a directed poset, and thus a filtered category.
$X$ is the colimit of the inclusion of $\I$ into $\cat{Set}$.
\end{proof}
\begin{theorem}
Let $\I$ be a small category.
Colimits of shape $\I$ in $\cat{Set}$ commute with all finite limits iff $\I$ is
filtered.
\end{theorem}
\begin{proof}
See \cite{m+m}, Corollary VII.6.5.
\end{proof}
\begin{defn}
\index{finitary!functor}
A functor $F: \ensuremath{\mathcal{C}} \to \ensuremath{\mathcal{D}}$ is \defterm{finitary} if it preserves filtered
colimits.
\end{defn}
\begin{defn}
\index{finitary!monad}
A monad $(T, \mu, \eta)$ on $\ensuremath{\mathcal{C}}$ is \defterm{finitary} if $T$ is finitary.
\end{defn}
A finitary monad on \cat{Set}\ is determined by its behaviour on finite sets, in the
following sense: since every set $X$ is a filtered colimit of its finite
subsets, then $TX$ has to be the colimit of the images under $T$ of the finite
subsets of $X$.
\section{Equivalences}
Let $(\Phi, E)$ be a presentation of an algebraic theory.
We define $K_{(\Phi, E)}$ to be the clone whose operations are elements of the
quotient signature $(\mathop{\rm{term}}\Phi)/\bar E$, with composition given by grafting,
and $\delta^i_n = x_i$ for all $i, n \in \natural$.
By definition of $\bar E$, grafting gives a well-defined family of composition
functions on $K_{(\Phi, E)}$.
Conversely, given a clone $K$, we may define a presentation of an algebraic
theory $(\Phi_K, E_K)$, by taking $(\Phi_K)_n = K_n$ for all $n \in \natural$,
and for all $n, m \in \natural$,
all $k \in K_n$ and all $k_1, \dots, k_n \in K_m$, letting $E_m$ contain the
equation
\[
k(k_1(x_1, \dots, x_m), \dots, k_n(x_1, \dots, x_m))
= k\ccomp(k_1,\dots, k_n)(x_1, \dots, x_m).
\]
\begin{lemma}
Let $K$ be a clone.
Then $K_{(\Phi_K, E_K)}$ is isomorphic to $K$.
\end{lemma}
\begin{proof}
See \cite{johnstone}, Lemma 1.7.
\end{proof}
\begin{lemma}
Let $(\Phi, E)$ be a presentation of an algebraic theory.
Let $(\Phi', E')$ be the presentation obtained from the clone $K_{(\Phi, E)}$.
Then the category $\Alg{\Phi, E}$ is isomorphic to the category $\Alg{\Phi',
E'}$
\end{lemma}
\begin{proof}
See \cite{johnstone}, Lemma 1.8.
\end{proof}
\begin{defn}
\index{linear!clone}
\index{strongly regular!clone}
Let $K$ be a clone.
We say that $K$ is \defterm{strongly regular} (resp. \defterm{linear}) if
there exists a strongly regular (resp. linear) presentation $P$ such that
$K = K_{(\Phi, E)}$.
\end{defn}
Given a clone $K$, we construct a Lawvere theory $\mathcal{T}_K$ for which
$\mathcal{T}_K({\bf n},{\bf m}) = (K_n)^m$.
Suppose $f = (f_1, \dots, f_m) \in \mathcal{T}_K(\bf{n}, \bf{m})$ and $g = (g_1, \dots,
g_p) \in \mathcal{T}_K(\bf{m}, \bf{p})$, then the composite $gf$ is $(g_1\ccomp (f_1, \dots, f_m), \dots, g_p \ccomp (f_1, \dots, f_m))$.
By the axioms for a clone, this is a category, with the identity map on
$\bf{n}$ being $(\delta^1_n, \dots, \delta^n_n)$.
It remains to show that ${\bf n}$ is the $n$th power of $\bf 1$ for every $n
\in \natural$.
The $i$th projection of $\bf n$ onto $1$ is evidently $\delta^i_n$: we must
show that these have the requisite universal property.
Take $m, n \in \natural$, and $n$ maps $f_1, \dots, f_n : {\bf m} \to {\bf 1}$
in $\mathcal{T}_K$.
The diagram
\[
\xymatrixcolsep{3pc}
\xymatrix{
& {\bf m} \ar@/u0.8cm/[ddl]_{f_1} \unar[d]^h \ar@/u0.8cm/[ddr]^{f_n} \\
& {\bf n} \ar[dl]^{\delta^1_n} \ar[dr]_{\delta^n_n} \\
{\bf 1} & \dots & {\bf 1}
}
\]
commutes if and only if $h = (f_1, \dots, f_n)$, and hence ${\bf n}$ is indeed
the $n$th power of ${\bf 1}$, and so $\mathcal{T}_K$ is a Lawvere theory.
The Lawvere theories so constructed evidently respect isomorphisms of clones.
Furthermore, the diagram
\[
\xymatrixrowsep{3pc}
\xymatrix{
\cat{Clone} \ar[rr]^{\mathcal{T}_{(-)}} \ar[dr]_\cat{Alg} & & \cat{Law} \ar[dl]^\cat{Alg} \\
& \cat{CAT}^{\rm{op}}
}
\]
commutes up to equivalence:
\begin{theorem}
Let $K$ be a clone.
Then $\Alg K \simeq \Alg{\mathcal{T}_K}$.
\end{theorem}
\begin{proof}
Let $A$ be a $K$-algebra.
We define a $\mathcal{T}_K$-algebra $F_A$ as follows:
\begin{itemize}
\item $F_A {\bf n} = A^n$ for all $n \in \natural$;
\item If $k \in \mathcal{T}_K({\bf n}, 1) = K_n$, then $F_Ak = \hat k$;
\item if $(k_1, \dots, k_n) : {\bf m} \to {\bf n}$ in $\mathcal{T}_K$, then
$F_A (k_1, \dots, k_n)$ is the unique arrow $A^m \to A^n$ such that the diagram
\[
\xymatrix{
& A^m \ar[ddl]_{\widehat k_1} \ar[ddr]^{\widehat k_n} \unar[d] \\
& A^n \ar[dl]^{\widehat{\delta^1_n}} \ar[dr]_{\widehat{\delta_n^n}} \\
A & \dots & A
}
\]
commutes.
\end{itemize}
Let $f : A \to B$ be a morphism of $K$-algebras.
Then the diagram
\[
\commsquare{A^n}{f^n}{B^n}{\widehat k}{\widehat k}A f B
\]
commutes for all $n \in \natural$ and all $k \in K_n$.
By the universal property of $B^m$, the diagram
\[
\commsquare{A^n}{f^n}{B^n}{F_A(k_1, \dots, k_m)}{F_B(k_1, \dots, k_m)}
{A^m}{f^m}{B^m}
\]
commutes for all $n, m \in \natural$ and all $(k_1, \dots, k_m) : {\bf n} \to {\bf m}$ in $\mathcal{T}_K$.
Hence $F_f = (f^n)_{{\bf n}}$ is a natural transformation $F_A \to F_B$, and
hence a morphism of $\mathcal{T}_K$-algebras.
This defines a functor $F_{(-)} : \Alg{K} \to \Alg{\mathcal{T}_K}$; we wish to show that
it is an equivalence.
For every $\mathcal{T}_K$-algebra $G$, we may define a $K$-algebra $A$ by setting $A =
G{\bf 1}$ and $\hat k = G({\bf n} \stackrel{k}{\longrightarrow} {\bf 1})$ for
all $k \in K_n$ and all $n \in \natural$.
Then $F_A$ is isomorphic as a $\mathcal{T}_K$-algebra to $G$, and hence the functor
$F_{(-)} : \Alg{K} \to \Alg{\mathcal{T}_K}$ is essentially surjective on objects.
We shall show further that it is full and faithful.
Let $A$ and $B$ be $K$-algebras, and let $\alpha_n : F_A \to F_B$ be
a morphism between their associated $\mathcal{T}_K$-algebras.
Since the diagram
\[
\commsquare{A^n}{\alpha_n}{B^n}{\widehat{\delta^i_n}}{\widehat{\delta^i_n}}
A {\alpha_1} B
\]
commutes for all $n \in \natural$ and all $i \in \{1, \dots, n\}$, it must be
the case that $\alpha_n = \alpha_1^n$ for all $n$.
Hence, the diagram
\[
\commsquare{A^n}{\alpha_1^n}{B^n}{\widehat k}{\widehat k}A {\alpha_1} B
\]
must commute for all $n \in \natural$ and all $k \in K_n$.
So $\alpha_1$ is a $K$-algebra morphism, and $\alpha_n = F_{\alpha_1}$.
Hence $F_{(-)}$ is full.
Suppose $F_f = F_g$; then $(F_f)_{\bf 1} = (F_g)_{\bf 1}$, so $f = g$.
Hence $F_{(-)}$ is faithful; and hence it is an equivalence of categories.
\end{proof}
Given a Lawvere theory $\mathcal{T}$, we can construct a clone $K_\mathcal{T}$, as follows:
\begin{itemize}
\item Let $(K_\mathcal{T})_n = \mathcal{T}(\bf{n},\bf{1})$ for all $n \in \natural$.
\item For all $n, m \in \natural$, all $f \in (K_\mathcal{T})_n$ and all $g_1, \dots,
g_n \in (K_\mathcal{T})_m$, let $f \ccomp (g_1, \dots, g_n) = f \fcomp (g_1 + \dots
+ g_n) \fcomp \Delta$,
where $(g_1 + \dots + g_n)$ is the unique map ${\bf mn} \to
{\bf n}$ in $\mathcal{T}$ such that the diagram
\[
\xymatrixcolsep{3pc}
\xymatrix{
& {\bf mn} \unar[dd]|{g_1 + \dots + g_n}
\ar[dl] \ar[dr] \\
{\bf m} \ar[dd]^{g_1}
& & {\bf m} \ar[dd]_{g_n} \\
& {\bf n} \ar[dl] \ar[dr] \\
{\bf 1} & \dots & {\bf 1}
}
\]
commutes, and $\Delta : {\bf m} \to {\bf mn}$ is the diagonal map (or
equivalently, the image of the codiagonal function ${\bf mn} \to {\bf m}$ under
the contravariant embedding of $\ensuremath{\bbF}$ into $\mathcal{T}$).
\item For all $n \in \natural$ and all $i \in {1, \dots, n}$, let $\delta_n^i$
be the $i$th projection $\bf{n} \to \bf{1}$.
\end{itemize}
This extends to a functor $K_{(-)} : \cat{Law} \to \cat{Clone}$, as follows: given
Lawvere theories $\mathcal{T}_1$ and $\mathcal{T}_2$, and a morphism of Lawvere theories $F: \mathcal{T}_1
\to \mathcal{T}_2$, let $K_F$ be the map of signatures sending $k \in (K_{\mathcal{T}_1})_n =
\mathcal{T}_1(\bf{n}, \bf{1})$ to $Fk \in (K_{\mathcal{T}_2})_n = \mathcal{T}_2(\bf{n}, \bf{1})$.
Since $F$ is a functor, and thus commutes with composition in $\mathcal{T}_1, \mathcal{T}_2$,
then $K_F$ must commute with composition in $K_{\mathcal{T}_1}$ and $K_{\mathcal{T}_2}$.
Since $F$ preserves finite products, it commutes with the projection maps in
$K_{\mathcal{T}_1}$ and $K_{\mathcal{T}_2}$.
Thus, $K_F$ is a morphism of clones.
\begin{theorem}
The functor $K_{(-)}$ is pseudo-inverse to the functor $\mathcal{T}_{(-)}$.
\end{theorem}
\begin{proof}
Since every object in a Lawvere theory is a copower of ${\bf 1}$, a Lawvere
theory $\mathcal{T}$ is entirely determined (up to isomorphism) by the hom-sets $\mathcal{T}({\bf
n}, {\bf 1})$, and thus by $K_\mathcal{T}$.
The theorem follows straightforwardly.
\end{proof}
Given a Lawvere theory $\mathcal{T}$, we construct a monad $(T, \mu, \eta)$ on \cat{Set}\ as
follows:
\begin{itemize}
\item If $X$ is a set, let $T X = \int^{n\in \ensuremath{\bbF}} \mathcal{T}({\bf n}, {\bf 1})
\times X^n$.
\item If $x \in X$, then $\eta(x) = (1,x) \in T X$.
\item If $f : {\bf n} \to {\bf 1}$ in $T$ and $(f_i, x^i_\bullet) \in \mathcal{T}({\bf k_i},
{\bf 1}) \times X^{k_i}$ for $i = 1, \dots, n$, then
\[
\mu(f,((f_1,x^1_\bullet), \dots, (f_n, x^n_\bullet)))
= (f\fcomp(f_1+ \dots+ f_n), x\seq\udot)
\]
\end{itemize}
\begin{theorem}
The monad so constructed is finitary.
\end{theorem}
\begin{proof}
See \cite{a+r}, Theorems 3.18 and 1.5, and Remarks 3.4(4) and 3.6(6).
\end{proof}
Given a finitary monad $T$ on \cat{Set}, we can construct a Lawvere theory $\mathcal{T}$.
Take the full subcategory $\ensuremath{\bbF}_T$ of the Kleisli category $\cat{Set}_T$ whose objects
are finite sets.
Now let $\mathcal{T}$ be the skeleton of the dual of $\ensuremath{\bbF}_T$.
The monad induced by this Lawvere theory is isomorphic to the original monad:
see \cite{a+r}, Remark 3.17 and Theorem 3.18.
The moral of the above theorems is that presentations, clones, Lawvere theories
and finitary monads on \cat{Set}\ all capture the same notion, and may be used
interchangeably.
Further, the notion that is captured corresponds to our usual intuitive
understanding of equational algebraic theories.
The equivalence between (finitary monads on \ensuremath{\mathcal{C}}) and (monads on \ensuremath{\mathcal{C}}\ that may be
described by a finitary presentation) may actually be generalized to the
case where \ensuremath{\mathcal{C}}\ is an arbitrary finitely presentable category: see \cite{k+p}.
|
\section{Introduction and results}
Consider the Bessel differential equation
\begin{equation}\label{1}
z^2\frac{d^2\mathcal{C}_\nu(z)}{dz^2}+z\frac{d\mathcal{C}_\nu(z)}{dz}+(z^2-\nu^2)\mathcal{C_{\nu}}(z)=0.
\end{equation}
The general solution is
$\mathcal{C}_\nu(z)=J_\nu(z)\cos\alpha-Y_\nu(z)\sin\alpha$, a
linear combination of $J_\nu(z)$ and $Y_\nu(z)$ being the Bessel
functions of the first and second kind (defined e.g. in Ref.
\cite{Abramowitz}).
When $\nu$ is real, the Bessel functions $J_{\nu}(x)$,
$Y_{\nu}(x)$, their derivatives $J'_{\nu}(x)$, $Y'_{\nu}(x)$ each
have an infinite number of positive real zeros, all of which are
simple with the possible exception of $x=0$ \cite{Abramowitz} (see
also \cite{Watson}, Bessel-Lommel and Rolle's theorems). For
non-negative order $\nu$ the $s$th positive real zeros of these
functions are denoted by $j_{\nu,s}$, $y_{\nu,s}$, $j'_{\nu,s}$,
$y'_{\nu,s}$, respectively, except that $x=0$ is counted as the
first zero of $J'_{0}(x)$. Since $J'_{0}(x)=-J_{1}(x)$, it follows
that $j'_{0,1}=0,j'_{0,s}=j_{1,s-1}, (s=2,3,...)$
\cite{Abramowitz}.
The following results are widely known
\cite{Abramowitz,Watson,LiuZou} about the zeros of the Bessel
functions.
\begin{theorem}\label{thm:1}
The positive real zeros of the functions $J_\nu(x)$,
$J_{\nu+1}(x)$, $Y_\nu(x)$, $Y_{\nu+1}(x)$, $J'_\nu(x)$ and
$Y'_\nu(x)$ interlace according to the three distinct
inequalities:
\begin{equation}\label{3}
j_{\nu,1}<j_{\nu+1,1}<j_{\nu,2}<j_{\nu+1,2}<j_{\nu,3}<\ldots
\end{equation}
\begin{equation}\label{4}
y_{\nu,1}<y_{\nu+1,1}<y_{\nu,2}<y_{\nu+1,2}<y_{\nu,3}<\ldots
\end{equation}
\begin{equation}\label{5}
\nu\leq j'_{\nu,1}<y_{\nu,1}<y'_{\nu,1}<j_{\nu,1}<j'_{\nu,2}
<y_{\nu,2}<y'_{\nu,2}<j_{\nu,2}<j'_{\nu,3}<\ldots
\end{equation}
Furthermore, the positive real zeros of $J'_\nu(x)$, $J'_{\nu+1}(x)$ and that of $Y'_\nu(x)$, $Y'_{\nu+1}(x)$ are interlaced:
\begin{equation}\label{lz1}
j'_{\nu,1}<j'_{\nu+1,1}<j'_{\nu,2}<j'_{\nu+1,2}<j'_{\nu,3}<\ldots
\end{equation}
\begin{equation}\label{lz2}
y'_{\nu,1}<y'_{\nu+1,1}<y'_{\nu,2}<y'_{\nu+1,2}<y'_{\nu,3}<\ldots
\end{equation}
\end{theorem}
Note that Eqs. (\ref{lz1})-(\ref{lz2}) were established only
recently in Ref. \cite{LiuZou} where a particular inverse
scattering problem was studied.
In addition to Theorem \ref{thm:1} the following auxiliary relations can be found.
\begin{proposition}\label{prop}
For non-negative orders, i.e. $\nu\geq0$
\begin{eqnarray}
j_{\nu+1,s}<j'_{\nu,s+1},\quad s=1,2,\ldots\label{palz}\\
y_{\nu+1,s}<y'_{\nu,s},\quad s=1,2,\ldots\label{propeq}
\end{eqnarray}
hold.
\end{proposition}
Eq. (\ref{palz}) emerged previously when studying a particular inverse scattering problem \cite{PA3}, and also in Ref. \cite{LiuZouTR} independently of the authors.
Now, with the aid of Proposition \ref{prop} it is possible to unify Eqs. (\ref{3}), (\ref{4}) and (\ref{5}) into a single one. In addition we obtain a simple proof of Eqs. (\ref{lz1}) and (\ref{lz2}).
We shall formulate our main result in a slightly generalized way, including also an interesting breaking condition.
\begin{theorem}[Interlacing of positive real zeros of the Bessel functions]\label{thm:3}
The positive real zeros of the Bessel functions $J_{\nu}(x)$,
$J'_{\nu}(x)$, $Y_{\nu}(x)$, $Y'_{\nu}(x)$,
$J_{\nu+\varepsilon}(x)$, $Y_{\nu+\varepsilon}(x)$,
$0<\varepsilon\leq1$, are interlaced according to the inequalities
\begin{equation}\label{ineq}
j'_{\nu,s}<y_{\nu,s}<y_{\nu+\varepsilon,s}<y'_{\nu,s}<j_{\nu,s}<j_{\nu+\varepsilon,s}<j'_{\nu,s+1}
<\ldots\quad s=1,2,\ldots,\,\nu\geq 0.
\end{equation}
For $\varepsilon>1$ this interlacing property is destroyed.
\end{theorem}
Eqs. (\ref{lz1}) and (\ref{lz2}) can be generalized to
\begin{equation}\label{jder}
j'_{\nu,1}<j'_{\nu+\varepsilon,1}<j'_{\nu,2}<j'_{\nu+\varepsilon,2}<j'_{\nu,3}<\ldots
\end{equation}
\begin{equation}\label{yder}
y'_{\nu,1}<y'_{\nu+\varepsilon,1}<y'_{\nu,2}<y'_{\nu+\varepsilon,2}<y'_{\nu,3}<\ldots
\end{equation}
with $0<\varepsilon\leq1$ and $\nu\geq0$. We note that the latter
two inequalities cannot be integrated with our unified interlacing
inequality (\ref{ineq}). While both
$j'_{\nu,s}<j'_{\nu+\varepsilon,s}<y_{\nu+\varepsilon,s}$ and
$y'_{\nu,s}<y'_{\nu+\varepsilon,s}<j_{\nu+\varepsilon,s}$ hold,
numerical counterexamples can easily be constructed for the
non-existence of a uniform inequality between both
$j'_{\nu+\varepsilon,s}$ and $y_{\nu,s}$ (for which
$j'_{\nu,s}<y_{\nu,s}<y_{\nu+\varepsilon,s}$ applies), and
$y'_{\nu+\varepsilon,s}$ and $j_{\nu,s}$ (for which
$y'_{\nu,s}<j_{\nu,s}<j_{\nu+\varepsilon,s}$ applies) for all
$s=1,2,\ldots$ and $\nu>0$, $0<\varepsilon\leq1$.
\section{Proofs}
We start by proving Eqs. (\ref{jder}) and (\ref{yder}) depending on Theorem \ref{thm:3}. The
first inequality is trivial due to the monotonicity of $j'_{\nu,s}$
in $\nu$ for $\nu\geq0$ (see \cite{Watson}). For proving the second inequality take the sequence of
Theorem \ref{thm:3} at some arbitrary $s$ for $\nu$, $\nu+1$ and for
$\nu+1$, $\nu+2$ with $\varepsilon=1$:
\begin{eqnarray*}
j'_{\nu,s}<y_{\nu,s}<y_{\nu+1,s}<y'_{\nu,s}<j_{\nu,s}<j_{\nu+1,s}<j'_{\nu,s+1}\\
j'_{\nu+1,s}<y_{\nu+1,s}<y_{\nu+2,s}<y'_{\nu+1,s}<j_{\nu+1,s}<j_{\nu+2,s}<j'_{\nu+1,s+1}
\end{eqnarray*}
From the first one we have $j_{\nu+1,s}<j'_{\nu,s+1}$ and from the
second one we have $j'_{\nu+1,s}<j_{\nu+1,s}$. Combining these two
yields the first interlacing property $j'_{\nu+1,s}<j'_{\nu,s+1}$
of Eq. (\ref{lz1}) for the derivative function $J'_\nu(x)$. For
$\varepsilon<1$ the relation follows from the the monotonicity of
$j'_{\nu,s}$ in $\nu$.
For the positive zeros of the derivative function $Y'_\nu(x)$ a
similar reasoning can be presented.
$y'_{\nu,s}<y'_{\nu+\varepsilon,s}$ is trivial due to the
monotonicity in $\nu$. Use Theorem \ref{thm:3} for $\nu$, $\nu+1$ and for
$\nu+1$, $\nu+2$ with $\varepsilon=1$
\begin{eqnarray*}
\ldots<y_{\nu+1,s+1}<y'_{\nu,s+1}<\ldots\\
\ldots<y'_{\nu+1,s}<j_{\nu+1,s}<j'_{\nu+1,s+1}<y_{\nu+1,s+1}<\ldots
\end{eqnarray*}
and combine them to get the relations $y'_{\nu+1,s}<y'_{\nu,s+1}$
contained in Eq. (\ref{lz2}) . Again the monotonicity of
$y_{\nu,s}$ in $\nu$ implies the non-trivial inequalities
$y'_{\nu+\varepsilon,s}<y'_{\nu,s+1}$ for $\varepsilon<1$.
Eq. (\ref{palz}) of Proposition \ref{prop} was already proven,
independently of each other, in Refs. \cite{PA3} and
\cite{LiuZouTR} therefore its proof is omitted here. The proof of
Eq. (\ref{propeq}) is elementary and based on the analysis of
intervals on which both $Y_{\nu+1}(x)$ and $Y_\nu(x)$ take the
same and the opposite sign. Of course, by definition,
$Y_{\nu+1}(x)$ and $Y_{\nu}(x)$ each keeps sign in the intervals
defined by their two consecutive zeros, respectively. That is
$Y_{\nu+1}(x)$ keeps the sign in the interval
\begin{equation*}
y_{\nu+1,s}<x<y_{\nu+1,s+1},\qquad s=1,2,\ldots,
\end{equation*}
and $Y_{\nu}(x)$ does it in
\begin{equation*}
y_{\nu,s}<x<y_{\nu,s+1},\qquad s=1,2,\ldots.
\end{equation*}
However, since $Y_{\nu}(x\to 0)=-\infty$ for $\nu\geq0$ and using
Eq. (\ref{4}) [which implies that $y_{\nu,s}<y_{\nu+1,s} $ and
$y_{\nu,s+1}<y_{\nu+1,s+1} $], one concludes by induction that
both $Y_{\nu+1}(x)$ and $Y_\nu(x)$ keep the same sign in the
common intervals
\begin{equation}\label{17}
y_{\nu+1,s}<x<y_{\nu,s+1},\qquad s=1,2,\ldots,
\end{equation}
whereas in
\begin{equation}\label{18}
y_{\nu,s}<x<y_{\nu+1,s},\qquad s=1,2,\ldots,
\end{equation}
the signs do differ. Now let us take the recurrence relation
$$
\mathcal{C}'_\nu(x)=-\mathcal{C}_{\nu+1}(x)+\frac{\nu}{x}\mathcal{C}_\nu(x)
$$
with $\mathcal{C}=Y$ at $x=y'_{\nu,s}$. It yields
\begin{equation}\label{19}
Y_{\nu+1}(y'_{\nu,s})=\frac{\nu}{y'_{\nu,s}}Y_{\nu}(y'_{\nu,s}),
\end{equation}
i.e. the signs of $Y_{\nu+1}(x)$ and $Y_{\nu}(x)$ coincide at
$x=y'_{\nu,s}$. But, because of (\ref{5}) [which tells that
$y'_{\nu,s}$ lies within the interval $y_{\nu,s}<x<y_{\nu,s+1}$],
the content of Eq. (\ref{19}) means also that $y'_{\nu,s}$ must be
in the common intervals given above by (\ref{17}). This completes
the proof of Proposition \ref{prop} for $y_{\nu+1,s}<y'_{\nu,s}$.
The proof of the first part of Theorem \ref{thm:3} is also elementary and follows from the
application of the two relations of Proposition \ref{prop} [being
previously unknown] in conjunction with the three distinct
inequalities (\ref{3}), (\ref{4}) and (\ref{5}) [being already known, i.e. from
Ref. \cite{Abramowitz}]. The case $0<\varepsilon<1$ is immediately implied
by the well-known property of $j_{\nu,s}$'s, and $y_{\nu,s}$'s that, for a
fixed $s$, they are strictly increasing functions of $\nu$ if $\nu\geq0$ \cite{Watson}.
The negative statement for $\varepsilon>1$ can be deduced from
$y_{\nu+\varepsilon,s}>j_{\nu,s}$, that is from the violation of
the prescribed relation between the third and fifth term in the
inequality sequence of Theorem \ref{thm:3}. In Ref. \cite{PA3} it
was proven that the Wronskian $W_{\nu\mu}(x)\equiv
J_\nu(x)Y_\mu'(x)-J_\nu'(x)Y_\mu
(x)\neq0$ for $x\in(0,\infty)$ if
and only if $0<|\nu-\mu|\leq1$ is maintained ($\nu\neq\mu$). One
of the ideas in that proof is that the set of extremal points of
$W_{\nu\mu}(x)$ is
$\{j_{\nu,s}\}_{s=1}^\infty\cup\{y_{\mu,s}\}_{s=1}^\infty$ and it
has been unveiled that the inequality sequences
$y_{\mu,s}<j_{\nu,s}<y_{\mu,s+1}<j_{\nu,s+1}$, $s=1,2,\ldots$
hold
if and only if all the local extrema are of the same sign, i.e.
$W_{\nu\mu}(x)\neq0$, $x\in\mathbb{R}^+$.
These relations are exactly the same that we are studying here
with $\mu=\nu+\varepsilon$. Since for $\varepsilon>1$
$|\nu-\mu|\leq1$ cannot hold and there is at least one root of
$W_{\nu\mu}(x)$ the inequalities must be violated for some $s$.
Now our proof is complete.
|
\section{Introduction}
The purpose of this paper is to study Sobolev space mapping properties of the
direct and inverse scattering maps for the one-dimensional Schr\"{o}dinger
equation with a potential of low regularity and no bound states. One of our
motivations is to use the scattering maps for the Schr\"{o}dinger equation to
construct and study solutions of the KdV equation on the line with initial
data of low regularity using the inverse scattering method. This paper
presents first steps toward this goal which we will continue in
\cite{HMP:2011}.
In this paper, we will describe a new representation for singular potentials
on the line, the Riccati representation, inspired by the work of Kappeler,
Perry, Shubin, and Topalov \cite{KPST:2005} on the Miura map \cite{Miura:1968}%
. As we will see, the Sobolev mapping properties of the scattering map are
particularly transparent when this representation is used. An analogous
representation for Schr\"{o}dinger operators on the circle appears in the work
of Kappeler and Topalov \cite{KT:2003,KT:2005a,KT:2005b,KT:2006} on well-posedness of the periodic KdV and mKdV equations.
If $q$ is a real-valued distribution on the real line belonging to the space
$H^{-1}(\mathbb{R})$, the Schr\"{o}dinger operator $-d^{2}/dx^{2}+q$ may be
defined as the self-adjoint operator associated to the closure of the
semibounded quadratic form
\begin{equation}
\mathfrak{q}(\varphi)=\int\left\vert \varphi^{\prime}(x)\right\vert
^{2}dx+\left\langle q,\left\vert \varphi\right\vert ^{2}\right\rangle
\label{eq:qf}%
\end{equation}
with domain $C_{0}^{\infty}(\mathbb{R})$ (see Appendix B in \cite{KPST:2005}
and references therein). It is natural to begin by considering such singular
potentials without negative-energy bound states, i.e., distributions $q$ for
which the quadratic form (\ref{eq:qf}) is non-negative. As shown in
\cite{KPST:2005}, such a distribution can be presented in the form
\[
q=u^{\prime}+u^{2}%
\]
where $u\in L_{\mathrm{loc}}^{2}(\mathbb{R})$ is the logarithmic derivative of
a positive solution $y\in H_{\mathrm{loc}}^{1}(\mathbb{R})$ of the zero-energy
Schr\"{o}dinger equation $-y^{\prime\prime}+qy=0$. The function $u$ is called
a \emph{Riccati representative} for the distribution $q$.
There is a one-to-one correspondence between Riccati representatives $u$ and
strictly positive solutions $y$ to the zero-energy Schr\"{o}dinger equation,
normalized so that $y(0)=1$. This latter set consists either of a single point
or a one-parameter family of solutions. Explicitly, $y=\theta y_{-}
+(1-\theta)y_{+}$, where $y_{\pm}$ are the unique, normalized, positive
solutions with the property that
\[
\int_{0}^{\infty}\frac{ds}{y^{2}_{+}(s)}
= \int_{-\infty}^{0}\frac{ds}{y^{2}_{-}(s)}
= \infty
\]
(see \S 5 of \cite{KPST:2005}). If we set $u_{\pm}=\frac{d}{dx}\log y_{\pm}$,
these \textquotedblleft extremal\textquotedblright\ Riccati representatives
$u_{\pm}$ have the property that $v:=u_{-}-u_{+}$ is a nonnegative, H\"{o}lder
continuous function and is either strictly positive, if $u_{+}\neq u_{-}$, or
identically zero, if $u_{+}=u_{-}$.
We can now describe the class of potentials we will study and define the
Riccati representation for such potentials that will play a central role in
our work. Denote by $\mathcal{Q}$ the set of real-valued distributions $q\in
H^{-1}(\mathbb{R})$ with the properties that
(i) the quadratic form (\ref{eq:qf}) is non-negative, and
(ii) the Riccati representatives $u_{\pm}$ obey $u_{\pm}\in L^{1}%
(\mathbb{R}^{\pm})\cap L^{2}(\mathbb{R})$.\newline We have $\mathcal{Q}%
=\mathcal{Q}_{0}\cup\mathcal{Q}_{>}$ where $\mathcal{Q}_{0}$ is the set of all
$q\in Q$ with $v(0)=0$, and $\mathcal{Q}_{>}$ is the set of all such
distributions with $v(0)>0$. This class includes the usual Faddeev--Marchenko
class\footnote{That is, real-valued measurable functions $q$ with $\int\left(
1+\left\vert x\right\vert \right) \left\vert q(x)\right\vert ~dx<\infty$.}
but also positive measures with suitable decay, certain highly oscillating
potentials, and sums of delta functions with positive weight (see \S 1 of
\cite{FHMP:2009} and \S 2 of \cite{HMP:2010} for further examples). The set
$\mathcal{Q}_{0}$ is very unstable under perturbations so that potentials in
the sets $\mathcal{Q}_{0}$ and $\mathcal{Q}_{>}$ are referred to respectively
as \textquotedblleft exceptional\textquotedblright\ and \textquotedblleft
generic\textquotedblright\ potentials.
A distribution $q\in\mathcal{Q}$ is uniquely determined by the data%
\begin{equation}
\left( \left. u_{-}\right\vert _{(-\infty,0)},\left. u_{+}\right\vert
_{(0,\infty)},v(0)\right) \in X^{-}\times X^{+}\times\lbrack0,\infty
),\label{eq:vbl.riccati}%
\end{equation}
where $X^{\pm}=L^{2}(\mathbb{R}^{\pm})\cap L^{1}(\mathbb{R}^{\pm})$ (see
\cite{HMP:2010}, Lemma 2.3). We will call the triple $\left( \left.
u_{-}\right\vert _{(-\infty,0)},\left. u_{+}\right\vert _{(0,\infty
)},v(0)\right) $ the \emph{Riccati representation} of $q$. Note that
$q\in\mathcal{Q}_{0}$ has a unique Riccati representative $u\in L^{1}%
(\mathbb{R})\cap L^{2}(\mathbb{R})$.
For $q\in\mathcal{Q}$, it was shown in \cite{FHMP:2009} (for $q\in
\mathcal{Q}_{0}$) and \cite{HMP:2010} (for $q\in\mathcal{Q}_{>}$) that the
usual formulation of scattering theory for the Schr\"{o}dinger equation
carries over. First, there exist Jost solutions $f_{\pm}(x,k)$, asymptotic as
$x\rightarrow\pm\infty$ to $\exp\left( \pm ikx\right) $. Second, one can
use these solutions to define reflection coefficients $r_{\pm}(k)$ that
describe scattering. The scattering maps $\mathcal{S}_{\pm}\,$\ are then
defined as%
\[
\mathcal{S}_{\pm}:q \mapsto r_{\pm}.
\]
We will study the scattering maps, parameterizing their domain using the
Riccati representation.
The Riccati representation connects the scattering problem for the
Schr\"{o}dinger equation to the scattering problem for the ZS--AKNS\ system
(see Zakharov--Shabat \cite{ZS:1971} and Ablowitz--Kaup--Newell--Segur
\cite{AKNS:1974}):%
\begin{equation}
\frac{d}{dx}\Psi=ik\sigma_{3}\Psi+Q(x)\Psi,\label{eq:AKNS}%
\end{equation}
where%
\[
\sigma_{3}=\left(
\begin{array}
[c]{cc}%
1 & 0\\
0 & -1
\end{array}
\right)
\]
and
\begin{equation}
Q(x)=\left(
\begin{array}
[c]{cc}%
0 & u(x)\\
u(x) & 0
\end{array}
\right) ,\label{eq:Q}%
\end{equation}
where $u$ is a Riccati representative for $q$. If $q\in\mathcal{Q}_{0},$ then
the Schr\"{o}dinger scattering problem is in fact equivalent to the scattering
problem for (\ref{eq:AKNS}) with potential (\ref{eq:Q}), and the scattering
maps can be studied using techniques developed for the ZS--AKNS\ system (see
\cite{Frayer:2008} and \cite{FHMP:2009}). On the other hand, if $q\in
\mathcal{Q}_{>}$, one can construct Jost solutions $f_{+}$ and $f_{-}$ for the
Schr\"{o}dinger equation from scattering solutions associated to
ZS--AKNS\ systems (\ref{eq:AKNS}), where the potential $Q$ is given by
(\ref{eq:Q}) respectively with $u=u_{+}$ and $u=u_{-}$.
The Riccati representation is particularly well-suited to studying Sobolev
space mapping properties of the scattering map. We first consider the case of
$q\in\mathcal{Q}_{0}$, where $q$ is specified uniquely by a single real-valued
Riccati representative $u\in X$, with $X$ denoting the Banach space
$L^{1}(\mathbb{R})\cap L^{2}(\mathbb{R})$ with norm
\[
\left\Vert u\right\Vert _{X}=\left\Vert u\right\Vert _{L^{1}(\mathbb{R}%
)}+\left\Vert u\right\Vert _{L^{2}(\mathbb{R})}.
\]
We will write $X_{\mathbb{R}}$ for the real Banach space of real-valued
functions $u\in X$. Denote by $\widehat{X}$ and $\widehat{X_{\mathbb{R}}}$ the
images of $X$ and $X_{\mathbb{R}}$ under the Fourier transform, set
$\left\Vert \widehat{v}\right\Vert _{\widehat{X}}=\left\Vert v\right\Vert
_{X}$, and let%
\[
\widehat{X}_{1}:=\left\{ r\in\widehat{X_{\mathbb{R}}}:\left\Vert r\right\Vert
_{\infty}<1\right\} .
\]
Note that $r(-k)=\overline{r(k)}$ for any $r\in\widehat{X_{\mathbb{R}}}$. It
was shown in \cite{Frayer:2008}, \cite{FHMP:2009} that the scattering maps
$\mathcal{S}_{\pm}$ are invertible, locally bi-Lipschitz maps from
$X_{\mathbb{R}}$ onto $\widehat{X}_{1}$. Since the maps $\mathcal{S}_{\pm}$ in
the Riccati variable are scattering maps for the ZS--AKNS\ system, one can use
techniques of Zhou \cite{Zhou:1998} to prove the following refined Sobolev
mapping property. For $s\geq0$, let
\[
L^{2,s}(\mathbb{R}):=\left\{ u\in L^{2}(\mathbb{R}):\left( 1+\left\vert
x\right\vert \right) ^{s}u\in L^{2}(\mathbb{R})\right\}
\]
and denote by $H^{s}(\mathbb{R})$ the image of $L^{2,s}(\mathbb{R})$ under the
Fourier transform. Note that, for $s>1/2$, $L^{2,s}(\mathbb{R})\subset X$ and
$H^{s}(\mathbb{R})$ consists of continuous functions. If we set%
\[
H_{1}^{s}(\mathbb{R}):=\left\{ r\in H^{s}(\mathbb{R})\cap\widehat
{X_{\mathbb{R}}}:\left\Vert r\right\Vert _{\infty}<1\right\} ,
\]
one can prove the following refined mapping property.
\begin{theorem}
\label{thm:except}For any $s>1/2$, the restrictions $\mathcal{S}_{\pm}%
:L^{2,s}(\mathbb{R})\cap X_{\mathbb{R}}\rightarrow H_{1}^{s}(\mathbb{R})$ are
onto, invertible, locally bi-Lipschitz continuous maps.
\end{theorem}
We will not give the details of the proof but rather concentrate on the more
challenging case where $q\in\mathcal{Q}_{>}$. To formulate our main theorem we
first recall some results from \cite{HMP:2010}.
If $q\in\mathcal{Q}_{>}$, the reflection coefficients $r_{\pm}$ belong to
$\widehat{X_{\mathbb{R}}}$ , but $r_{\pm}(0)=-1$ and $\left\vert
r(k)\right\vert <1$ for $k\neq0$. For smooth, compactly supported generic
potentials, one has $r_{\pm}(k)=-1+\mathcal{O}(k^{2})$ as $k\rightarrow0$
(see, for example, \cite{DT:1979}, \S 2, Theorem 1, Part V and Remark 9); in
general, as shown in \cite{HMP:2010}, one has the weaker condition that the
functions%
\[
\frac{1-\left\vert r_{\pm}(k)\right\vert ^{2}}{k^{2}}%
\]
belong to $\widehat{X_{\mathbb{R}}}$ and do not vanish at $k=0$. The direct
scattering maps in the Riccati variables are given by%
\[
\mathcal{S}_{\pm}:\left( \left. u_{-}\right\vert _{(-\infty,0)},\left.
u_{+}\right\vert _{(0,\infty)},v(0)\right) \mapsto r_{\pm}.
\]
For $r\in\widehat{X_{\mathbb{R}}}$, we shall write%
\begin{equation}
\widetilde{r}(k)=\frac{1-\left\vert r(k)\right\vert ^{2}}{k^{2}}%
\label{eq:tilder-def}%
\end{equation}
and denote%
\[
\mathcal{R}_{>}:=\left\{ r\in\widehat{X_{\mathbb{R}}}:r(0)=-1,~\left\vert
r(k)\right\vert <1\text{ if }k\neq0\text{, }~\widetilde{r}\in\widehat
{X_{\mathbb{R}}},~\widetilde{r}(0)\neq0\right\} .
\]
The space $\mathcal{R}_{>}$ is a metric space when equipped with the metric
\begin{equation}
d\left( r_{1},r_{2}\right) =\left\Vert r_{1}-r_{2}\right\Vert _{\widehat{X}%
}+\left\Vert \widetilde{r}_{1}-\widetilde{r}_{2}\right\Vert _{\widehat{X}%
}.\label{eq:dx}%
\end{equation}
In \cite{HMP:2010}, it was shown that the maps $\mathcal{S}_{\pm}$ are locally
bi-Lipschitz continuous onto maps from $X^{-}\times X^{+}\times(0,\infty)$
onto $\mathcal{R}_{>}$ equipped with the metric (\ref{eq:dx}).
We will prove a finer mapping property, analogous to Theorem \ref{thm:except},
for the scattering map on generic potentials. We set%
\[
\mathcal{R}_{s}=\left\{ r\in\mathcal{R}_{>}\cap H^{s}(\mathbb{R}%
):\widetilde{r}\in H^{s}(\mathbb{R})\right\}
\]
and equip $\mathcal{R}_{s}$ with the metric%
\[
d_{s}\left( r_{1},r_{2}\right) =\left\Vert r_{1}-r_{2}\right\Vert
_{H^{s}(\mathbb{R})}+\left\Vert \widetilde{r}_{1}-\widetilde{r}_{2}\right\Vert
_{H^{s}(\mathbb{R})}.
\]
\begin{theorem}
\label{thm:main}For any $s>1/2$, the direct scattering maps $\mathcal{S}_{\pm
}$ are invertible, locally bi-Lipschitz continuous maps from $L^{2,s}%
(\mathbb{R}^{-})\times L^{2,s}(\mathbb{R}^{+})\times(0,\infty)$ onto the space
$\mathcal{R}_{s}$.
\end{theorem}
Fourier-type mapping properties of the map $q\mapsto r$ have been studied by
many authors, including Cohen \cite{Cohen:1982}, Deift and Trubowitz
\cite{DT:1979}, and Faddeev \cite{Faddeev:1964}. These authors impose weighted
$L^{1}$ assumptions on $q$ and obtain regularity results for $r$ in terms of
$\infty$-norms of $r$ and its derivatives. Kappeler and Trubowitz
\cite{KT:1986}, \cite{KT:1988} studied Sobolev space mapping properties of the
scattering map, defined as follows. Let $s(k)=2ikr(k)/t(k)$, where $r$ is the
reflection coefficient and $t$ is the transmission coefficient, and introduce
the weighted Sobolev spaces
\begin{align*}
H_{n,\alpha} & =\left\{ f\in L^{2}:x^{\beta}\partial_{x}^{j}f\in
L^{2}\text{, }0\leq j\leq n,~0\leq\beta\leq\alpha\right\} ,\\
H_{n,\alpha}^{\#} & =\left\{ f\in H_{n,\alpha}:x^{\beta}\partial_{x}%
^{n+1}f\in L^{2},~1\leq\beta\leq\alpha\right\} .
\end{align*}
Kappeler and Trubowitz show that the map $q\mapsto s$ takes potentials $q\in
H_{N,N}$ without bound states to scattering functions $s$ belonging to
$H_{N-1,N}^{\#}$ for $N\geq3$. They extend their results to potentials with
finitely many bound states in \cite{KT:1988}. They also prove analyticity and
investigate the differential of the scattering map.
Our results are similar to those of Kappeler and Trubowitz in that we study
$L^{2}$-based Sobolev spaces, which leads to a more symmetrical formulation of
the mapping properties. In our case, we examine the scattering map in the
Riccati variables (\ref{eq:vbl.riccati}) and so treat potentials which are
more singular than the class treated by Kappeler and Trubowitz. In a
subsequent paper \cite{HMP:2011}, we will extend the methods developed here to
consider mapping properties between weighted fractional Sobolev spaces which
preserve the KdV flow.
This paper is organized as follows. In section 2, we first review the
connection between Jost solutions to the Schr\"{o}dinger and
ZS--AKNS\ equations. In section 3 we obtain estimates on the direct scattering
map using a Fourier representation for the Jost solutions derived in
\cite{FHMP:2009}. In section 4, we use the representation formulas of
\cite{HMP:2010}, derived from Gelfand--Levitan--Marchenko equation for the
ZS--AKNS\ system, to analyze the inverse scattering map. Finally, in section
5, we give the proof of the main theorem.
\textbf{Acknowledgements}. The research of RH and YM was partially supported
by Deutsche Forschungsgemeinschaft under project 436 UKR 113/84. RH\ was
supported in part by NSF grant DMS-0408419, and PP\ was supported in part by
NSF grants DMS-0408419 and DMS-0710477.
\section{Schr\"{o}dinger Scattering and the ZS--AKNS System}
\label{sec:pre}
In this section, we recall how the Jost solutions and reflection coefficients
for a Schr\"{o}dinger operator with Miura potential may be computed by solving
the associated ZS--AKNS\ equations with potentials $u_{+}$ and $u_{-}$. We
assume throughout that $u_{\pm}\in L^{2}(\mathbb{R})\cap L^{1}(\mathbb{R}%
^{\pm})$ are real-valued.
First, we recall the connection between the Schr\"{o}dinger equation with a
Miura potential and the ZS--AKNS\ system. If $u\in L_{\mathrm{loc}}%
^{2}(\mathbb{R})$ and $q=u^{\prime}+u^{2}$ then the Schr\"{o}dinger equation
\begin{equation}
-y^{\prime\prime}+qy=k^{2}y\label{eq:se}%
\end{equation}
is equivalent to the system%
\begin{equation}
\frac{d}{dx}\left(
\begin{array}
[c]{c}%
y\\
y^{\left[ 1\right] }%
\end{array}
\right) =\left(
\begin{array}
[c]{cc}%
u & 1\\
-k^{2} & -u
\end{array}
\right) \left(
\begin{array}
[c]{c}%
y\\
y^{\left[ 1\right] }%
\end{array}
\right) \label{eq:se.sys}%
\end{equation}
where $y^{\left[ 1\right] }:=y^{\prime}-uy$ is the quasi-derivative of $y$.
Note that $y$ and $y^{\left[ 1\right] }$ are absolutely continuous, and the
initial value problem for (\ref{eq:se.sys}) has a unique solution. For a given
choice of $u$ and solutions $g$ and $h$ of (\ref{eq:se}), the Wronskian%
\begin{equation}
\left[ f,g\right] =g(x)h^{\left[ 1\right] }(x)-g^{\left[ 1\right]
}(x)h(x)\label{eq:Wronski}%
\end{equation}
is independent of $x$.
The Jost solutions $f_{\pm}(x,k)$ satisfy (\ref{eq:se}) with respective
asymptotic conditions%
\begin{equation}
\lim_{x\rightarrow\pm\infty}\left(
\begin{array}
[c]{c}%
f_{\pm}(x)-e^{\pm ikx}\\
f_{\pm}^{\left[ 1\right] }(x)\mp ike^{\pm ikx}%
\end{array}
\right) =0\label{eq:fpm.asy}%
\end{equation}
where%
\[
f_{\pm}^{\left[ 1\right] }:=f_{\pm}-u_{\pm}f_{\pm}.
\]
If $\left[ ~\cdot~,~\cdot~\right] _{\pm}$ denotes the Wronskian
(\ref{eq:Wronski}) with $u=u_{\pm}$, it follows from the asymptotics
(\ref{eq:fpm.asy}) that
\[
-\left[ f_{+}(x,k),f_{+}(x,-k)\right] _{+}=\left[ f_{-}(x,k),f_{-}%
(x,-k)\right] _{-}=2ik.
\]
Thus, for $k\neq0$, there are coefficients $a(k)$ and $b(k)$ so that%
\[
f_{+}(x,k)=a(k)f_{-}(x,-k)+b(k)f_{-}(x,k).
\]
By standard arguments,%
\begin{equation}
\left\vert a(k)\right\vert ^{2}-\left\vert b(k)\right\vert ^{2}%
=1,\label{eq:ab1}%
\end{equation}
and the reality conditions%
\begin{equation}
a(-k)=\overline{a(k)},%
\qquad
b(-k)=\overline{b(k)}\label{eq:abr}%
\end{equation}
hold. Moreover,%
\begin{equation}
a(k)=\frac{\left[ f_{+}(x,k),f_{-}(x,k)\right] _{-}}{\left[ f_{-}%
(x,-k),f_{-}(x,k)\right] _{-}}\label{eq:a.wronski}%
\end{equation}
and%
\begin{equation}
b(k)=\frac{\left[ f_{+}(x,k),f_{-}(x,-k)\right] _{-}}{\left[ f_{-}%
(x,k),f_{-}(x,-k)\right] _{-}}.\label{eq:b.wronski}%
\end{equation}
The reflection coefficients $r_{\pm}$ are given by%
\begin{align}
r_{-}(k) & =b(k)/a(k),\label{eq:r-}\\
r_{+}(k) & =-b(-k)/a(k),\label{eq:r+}%
\end{align}
so that $\left\vert r_{+}(k)\right\vert =\left\vert r_{-}(k)\right\vert $.
\ The transmission coefficient is given by $t(k)=1/a(k)$, and the involution%
\begin{equation}
r(k)\mapsto-\frac{t(k)}{t(-k)}r(-k)\label{eq:inv}%
\end{equation}
maps $r_{-}$ to $r_{+}$ and vice versa.
To compute the Jost solutions $f_{\pm}$ we exploit the following connection
between the Schr\"{o}dinger equation with potential $q=u^{\prime}+u^{2}$ and
the ZS--AKNS\ system%
\begin{equation}
\frac{d}{dx}\Psi=ik\sigma_{3}\Psi+Q(x)\Psi\label{eq:ZS-AKNS}%
\end{equation}
with potential%
\begin{equation}
Q(x)=\left(
\begin{array}
[c]{cc}%
0 & u(x)\\
u(x) & 0
\end{array}
\right) .\label{eq:Qu}%
\end{equation}
If $\Psi=(\psi_{1},\psi_{2})^{T}$ is a vector-valued solution of
(\ref{eq:ZS-AKNS}) with potential (\ref{eq:Qu}), then
\[
\left(
\begin{array}
[c]{c}%
\psi_{1}+\psi_{2}\\
ik(\psi_{1}-\psi_{2})
\end{array}
\right)
\]
solves the system (\ref{eq:se.sys}). In particular, if $\Psi_{+}$ and
$\Psi_{-}$ are the unique matrix-valued solutions of the respective problems%
\begin{gather*}
\frac{d}{dx}\Psi^{\pm}=ik\sigma_{3}\Psi^{\pm}+\left(
\begin{array}
[c]{cc}%
0 & u_{\pm}(x)\\
u_{\pm}(x) & 0
\end{array}
\right) \Psi^{\pm},\\
\lim_{x\rightarrow\pm\infty}\left\vert \Psi^{\pm}(x)-e^{ixk\sigma_{3}%
}\right\vert =0,
\end{gather*}
then the formulas%
\begin{align}
f_{+}(x,k) & =\psi_{11}^{+}(x,k)+\psi_{21}^{+}(x,k)\label{eq:f+}\\
f_{+}^{\left[ 1\right] }(x,k) & =ik\left( \psi_{11}^{+}(x,k)-\psi_{21}%
^{+}(x,k)\right) \label{eq:f+[1]}\\
f_{-}(x,k) & =\overline{\psi_{11}^{-}(x,k)}+\overline{\psi_{21}^{-}%
(x,k)}\label{eq:f-}\\
f_{-}^{\left[ 1\right] }(x,k) & =-ik\left( \overline{\psi_{11}^{-}%
(x,k)}-\overline{\psi_{21}^{-}(x,k)}\right) \label{eq:f-[1]}%
\end{align}
hold, where the bar denotes complex conjugation. A short computation with
(\ref{eq:a.wronski})--(\ref{eq:b.wronski}) leads to the formulas%
\begin{align}
a(k) & =\left\vert
\begin{array}
[c]{cc}%
\psi_{11}^{+}(x,k) & \overline{\psi_{21}^{-}(x,k)}\\
& \\
\psi_{21}^{+}(x,k) & \overline{\psi_{11}^{-}(x,k)}%
\end{array}
\right\vert -\frac{v(x)}{2ik}f_{+}(x,k)f_{-}(x,k),\label{eq:a}\\
& \nonumber\\
b(k) & =\left\vert
\begin{array}
[c]{cc}%
\psi_{11}^{+}(x,k) & \overline{\psi_{11}^{+}(x,-k)}\\
& \\
-\psi_{21}^{+}(x,k) & -\overline{\psi_{21}^{-}(x,-k)}%
\end{array}
\right\vert +\frac{v(x)}{2ik}f_{+}(x,k)f_{-}(x,k),\label{eq:b}%
\end{align}
where, for a $2\times2$ matrix $A$, $\left\vert A\right\vert $ denotes the determinant.
These two formulas lie at the heart of our analysis for the direct problem.
They show explicitly the singularity at $k=0$ that occurs when $u_{+}\neq
u_{-}$; the singularity is always nonzero in this case since $v$ is strictly
nonzero and $f_{\pm}(x,0)$ are positive solutions of the zero-energy
Schr\"{o}dinger equation.
To study the scattering map via the formulas (\ref{eq:a})--(\ref{eq:b}), we
will use integral representations for the solutions $\Psi^{\pm}$. \ These
integral representations give $\Psi^{\pm}$ as Fourier transforms of functions
given by explicit multilinear series in $u_{\pm}$. Let~$\Psi^{\pm}%
(x,k)=\exp(ixk\sigma_{3})N^{\pm}(x,k)$ and denote by~$n_{ij}^{\pm}$ the
entries of~$N^{\pm}$. In order to compute the Jost solutions from
(\ref{eq:f+})--(\ref{eq:f-[1]}), it suffices to study~$n_{11}^{\pm}$ and
$n_{21}^{\pm}$. We will describe only the integral representations for
$n_{11}^{+}$ and $n_{21}^{+}$ and their properties since those of $n_{11}^{-}$
and $n_{21}^{-}$ are very similar.
From \cite{FHMP:2009}, section 3.1, equations (3.14)\ and (3.15) and
following, we have%
\begin{align*}
n_{11}^{+}(x,k)-1 & =\int_{0}^{\infty}A(x,\zeta)e^{i\zeta k}~d\zeta,\\
n_{21}^{+}(x,k) & =\int_{x}^{\infty}B(x,\zeta)e^{i\zeta k}~d\zeta.
\end{align*}
Here $A$ and $B$ have multilinear expansions of the form%
\[
A(x,\zeta)=\sum_{n=1}^{\infty}A_{n}(x,\zeta), \qquad B(x,\zeta)=\sum
_{n=1}^{\infty}B_{n}(x,\zeta)
\]
with%
\[
A_{n}(x,\zeta)=\int_{\Omega_{2n}(\zeta)}u_{+}(y_{1})\ldots u_{+}%
(y_{2n})~dS_{2n}
\]
and%
\[
B_{n}(x,\zeta)=\int_{\Omega_{2n-1}(\zeta)}u_{+}(y_{1})\ldots u_{+}%
(y_{2n})~dS_{2n-1},
\]
where, for $\zeta\in\mathbb{R}$, $\Omega_{n}(\zeta)$ is the set of all
$\mathrm{y}=\left( y_{1},\ldots,y_{n}\right) $ in $\mathbb{R}^{n}$ with
$x\leq y_{1}\leq\ldots\leq y_{n}$ and
\begin{equation}
\sum_{j=0}^{n-1}(-1)^{j}y_{n-j}=\zeta, \label{eq:zeta}%
\end{equation}
while $dS_{n}$ is surface measure on the hyperplane (\ref{eq:zeta}).
For each fixed $x$ we have
\begin{align}
\left\Vert n_{11}^{+}(x,~\cdot~)-1\right\Vert _{H^{s}(\mathbb{R})} &
\leq\left\Vert A(x,~\cdot~)\right\Vert _{L^{2,s}(\mathbb{R})}%
,\label{eq:n11+.to.A}\\
\left\Vert n_{21}^{+}(x,~\cdot~)\right\Vert _{H^{s}(\mathbb{R})} &
\leq\left\Vert B(x,~\cdot~)\right\Vert _{L^{2,s}(\mathbb{R})}%
.\label{eq:n21+.to.B}%
\end{align}
Thus, to estimate the $H^{s}$-norms of $n_{11}^{+}$ and $n_{21}^{+}$ as
functions of $k$, it suffices to obtain summable estimates on $\left\Vert
A_{n}(x,~\cdot~)\right\Vert _{L^{2,s}(\mathbb{R})}$ and $\left\Vert
B_{n}(x,~\cdot~)\right\Vert _{L^{2,s}(\mathbb{R})}$.
To do this, we first note the identity
\[
\left\Vert \psi\right\Vert _{L^{2,s}(\mathbb{R})}=\sup\left\{ \left\vert
\int\varphi(\zeta)(1+\left\vert \zeta\right\vert )^{s}\psi(\zeta
)~d\zeta\right\vert :\left\Vert \varphi\right\Vert _{L^{2}}=1\right\} .
\]
Next, setting ${\mathrm{y}}:=(y_{1},\dots,y_{n})$, ${\mathrm{d}}{\mathrm{y}%
}:=dy_{1}\cdots dy_{n}$, $U({\mathrm{y}}):=u_{+}(y_{1})\cdots u_{+}(y_{n})$,
and defining
\[
\zeta_{n}({\mathrm{y}}):=\sum_{j=0}^{n-1}(-1)^{j}y_{n-j},
\]
we find that
\begin{gather}
\int_{0}^{\infty}\varphi(\zeta)\int_{\Omega_{2n}(\zeta)}U({\mathrm{y}}%
)dS_{2n}d\zeta=\int_{x\leq y_{1}\leq\dots\leq y_{2n}}U({\mathrm{y}}%
)\varphi(\zeta_{2n}({\mathrm{y}})){\mathrm{d}}{\mathrm{y}},\label{eq:8.1}\\
\int_{x}^{\infty}\varphi(\zeta)\int_{\Omega_{2n-1}(\zeta)}U({\mathrm{y}%
})dS_{2n-1}d\zeta=\int_{x\leq y_{1}\leq\dots\leq y_{2n-1}}U({\mathrm{y}%
})\varphi(\zeta_{2n-1}({\mathrm{y}})){\mathrm{d}}{\mathrm{y}}.\label{eq:8.2}%
\end{gather}
Observe that for ${\mathrm{y}}$ obeying $0\leq x\leq y_{1}\leq\cdots\leq
y_{n}$ the estimate $|\zeta_{n}({\mathrm{y}})|\leq y_{n}$ holds. We then get
from the integral representation for $A_{n}$ and \eqref{eq:8.1} that, for any
$\varphi\in L^{2}({\mathbb{R}})$,
\begin{align*}
& \Bigl|\int_{0}^{\infty}(1+\zeta)^{s}\varphi(\zeta)A_{n}(x,\zeta
)\,d\zeta\Bigr|\\
& \leq\int_{x\leq y_{1}\leq\dots\leq y_{2n-1}}|u_{+}(y_{1})\cdots
u_{+}(y_{2n-1})|\int_{x}^{\infty}(1+y_{2n})^{s}|\varphi(\zeta_{2n}%
({\mathrm{y}}))||u_{+}(y_{2n})|\,{\mathrm{d}}{\mathrm{y}}\\
& \leq\frac{\Vert u_{+}\Vert_{L^{1}(\mathbb{R}^{+})}^{2n-1}}{(2n-1)!}\Vert
u_{+}\Vert_{L^{2,s}({\mathbb{R}}^{+})}\Vert\varphi\Vert_{L^{2}({\mathbb{R}})}.
\end{align*}
Therefore
\[
\Vert A_{n}(x,\,\cdot\,)\Vert_{L^{2,s}({\mathbb{R}}^{+})}\leq\frac{\Vert
u_{+}\Vert_{L^{1}(\mathbb{R}^{+})}^{2n-1}}{(2n-1)!}\Vert u_{+}\Vert
_{L^{2,s}({\mathbb{R}}^{+})},
\]
and similar estimates give
\[
\Vert B_{n}(x,\,\cdot\,)\Vert_{L^{2,s}({\mathbb{R}}^{+})}\leq\frac{\Vert
u_{+}\Vert_{L^{1}(\mathbb{R}^{+})}^{2n-2}}{(2n-2)!}\Vert u_{+}\Vert
_{L^{2,s}({\mathbb{R}}^{+})}.
\]
Since $A_{n}(x,\,\cdot\,)$ and $B_{n}(x,\,\cdot\,)$ are multilinear functions
of~$u_{+}$ and the series for $A(x,\,\cdot\,)$ and $B(x,\,\cdot\,)$ converge
absolutely in~$L^{2,s}({\mathbb{R}})$, standard arguments show that, for every
fixed $x\geq0$, $A(x,\,\cdot\,)$ and $B(x,\,\cdot\,)$ depend analytically in
$L^{2,s}({\mathbb{R}})$ on~$u_{+}\in X^{+}$. Hence:
\begin{proposition}
\label{prop:n+.Hs}Assume that $s>1/2$ and that $u_{+}\in L^{2,s}({\mathbb{R}%
}^{+})$. Then $n_{11}^{+}(x,\,\cdot\,)-1$ and $n_{21}^{+}(x,\,\cdot\,)$ belong
to $H^{s}(\mathbb{R})$ for each fixed~$x\geq0$, depend analytically therein
on~$u_{+} \in X^{+}$, and the estimates
\[
\sup_{x\geq0} \bigl(\Vert n_{11}^{+}(x,\,\cdot\,)-1 \Vert_{H^{s}(\mathbb{R})}
+ \Vert n_{21}^{+}(x,\,\cdot\,)\Vert_{H^{s}(\mathbb{R})}\bigr)
\leq\Vert u_{+}\Vert_{L^{2,s}({\mathbb{R}}^{+})} \exp\{\Vert u_{+}\Vert
_{L^{1}({\mathbb{R}}^{+})}\}
\]
hold.
\end{proposition}
A similar analysis, based on the integral representations for $n_{11}^{-}$ and
$n_{21}^{-}$, shows:
\begin{proposition}
\label{prop:n-.Hs}Assume that $s>1/2$ and that $u_{-}\in L^{2,s}({\mathbb{R}%
}^{-})$. Then $n_{11}^{-}(x,\,\cdot\,)-1$ and $n_{21}^{-}(x,\,\cdot\,)$ belong
to $H^{s}(\mathbb{R})$ for each fixed~$x\leq0$, depend analytically therein
on~$u_{-} \in X^{-}$, and the estimates
\[
\sup_{x\leq0}\bigl(\Vert n_{11}^{-}(x,\,\cdot\,)-1\Vert_{H^{s}(\mathbb{R})}
+\Vert n_{21}^{-}(x,\,\cdot\,)\Vert_{H^{s}(\mathbb{R})}\bigr)
\leq\Vert u_{-}\Vert_{L^{2,s}({\mathbb{R}}^{-})} \exp\{\Vert u_{-}\Vert
_{L^{1}({\mathbb{R}}^{-})}\}
\]
hold.
\end{proposition}
\section{The Direct Problem}
\label{sec:direct}
We now consider the mappings $\left( u_{-},u_{+},v(0)\right) \mapsto r_{\pm
}$. In order to study the mapping properties we introduce the auxiliary
functions%
\begin{align}
\widetilde{a}(k) & =\frac{k}{k+i}a(k),\label{eq:tildea}\\
\widetilde{b}(k) & =\frac{k}{k+i}b(k),\label{eq:tildeb}\\
\widetilde{r}(k) & =\frac{1-\left\vert r_{\pm}(k)\right\vert ^{2}}{k^{2}},
\label{eq:tilder}%
\end{align}
and note the relations%
\begin{equation}
r_{-}(k)=\frac{\widetilde{b}(k)}{\widetilde{a}(k)}, \qquad r_{+}(k)=\frac
{i-k}{i+k}\frac{\widetilde{b}(-k)}{\widetilde{a}(k)} \label{eq:rpm}%
\end{equation}
and%
\begin{equation}
\widetilde{r}(k)=\frac{1}{k^{2}+1}\frac{1}{\left\vert \widetilde
{a}(k)\right\vert ^{2}}. \label{eq:tilder.rep}%
\end{equation}
\begin{proposition}
\label{prop:direct}Suppose that $u_{\pm}\in L^{2,s}(\mathbb{R}^{\pm})$ for
some $s>1/2$. Then $r_{\pm}\in H^{s}(\mathbb{R})$ and $\widetilde{r}\in
H^{s}(\mathbb{R})$ with $\widetilde{r}(0)\neq0$, and the maps
\begin{align*}
L^{2,s}(\mathbb{R}^{+})\times L^{2,s}(\mathbb{R}^{-})\times\mathbb{R}^{+} &
\rightarrow H^{s}(\mathbb{R})^{3}\\
\left( u_{+},u_{-},v(0)\right) & \mapsto(r_{-},r_{+},\widetilde{r})
\end{align*}
are locally Lipschitz continuous.
\end{proposition}
\begin{proof}
From the representation formulae (\ref{eq:a}) and (\ref{eq:b}) evaluated at
$x=0$ we have%
\[
\widetilde{a}(k)=\frac{k}{k+i}\left\vert
\begin{array}
[c]{cc}%
n_{11}^{+}(0,k) & \overline{n_{21}^{-}(0,k)}\\
& \\
n_{21}^{+}(0,k) & \overline{n_{11}^{-}(0,k)}%
\end{array}
\right\vert -\frac{1}{k+i}\frac{v(0)}{2i}f_{+}(0,k)f_{-}(0,k)
\]
and
\[
\widetilde{b}(k)=\frac{k}{k+i}\left\vert
\begin{array}
[c]{cc}%
n_{11}^{+}(0,k) & \overline{n_{11}^{+}(0,-k)}\\
& \\
-n_{21}^{+}(0,k) & -\overline{n_{21}^{-}(0,-k)}%
\end{array}
\right\vert +\frac{1}{k+i}\frac{v(0)}{2i}f_{+}(0,k)f_{-}(0,k).
\]
In view of Propositions~\ref{prop:n+.Hs} and \ref{prop:n-.Hs} the functions
$n_{ij}^{\pm}(0,\,\cdot\,)$ and $f_{\pm}(0,\,\cdot\,)$ belong to the Banach
algebra $1\dotplus H^{s}({\mathbb{R}})$ (see Appendix~\ref{sec:app}) and
depend locally Lipschitz continuously therein on the Riccati variables
$(u_{+},u_{-},v(0))$; thus the same is true of $\widetilde{a}$ and
$\widetilde{b}$. Moreover, the function $\widetilde{a}$ is an invertible
element of $1\dotplus H^{s}({\mathbb{R}})$. Indeed, by
Lemma~\ref{lem:spectrum} it suffices to show that $\inf|\widetilde{a}|>0$. We
observe that $\widetilde{a}(0)=v(0)f_{+}(0,k)f_{-}(0,k)\neq0$ while
$|\widetilde{a}(k)|>0$ for all nonzero real~$k$ due to~\eqref{eq:ab1}.
Representation~\eqref{eq:a.wronski} for $a$ along with the asymptotic behavior
of the Jost solutions imply that $|a(k)|\rightarrow1$ as $k\rightarrow
\pm\infty$, so that $|\widetilde{a}(k)|\rightarrow1$ as $k\rightarrow\pm
\infty$ as well. Recalling that $\widetilde{a}$ is a continuous function, we
conclude that $\widetilde{a}$ is an invertible element of the Banach
algebra~$1\dotplus H^{s}({\mathbb{R}})$. Clearly, the same conclusion holds
for all $\widetilde{a}$ in a neighborhood of the given one.
We now use~\eqref{eq:rpm} to conclude that the reflection coefficients
$r^{\pm}$ belong to $H^{s}({\mathbb{R}})$ and depend therein locally Lipschitz
continuously on the Riccati variables. Since $|\widetilde{a}(k)|^{2}$ is an
invertible element of $1\dotplus H^{s}({\mathbb{R}})$,
relation~\eqref{eq:tilder.rep} yields the inclusion $\widetilde{r}\in
H^{s}({\mathbb{R}})$, and the continuous dependence follows by the same
arguments as above. Finally, \eqref{eq:tilder.rep} and $\widetilde{a}(0)\neq0$
yield $\widetilde{r}(0)\neq0$, and the proof is complete.
\end{proof}
Finally, we note the following variant of Proposition~3.3 of~\cite{HMP:2010},
which concerns continuity of the involution (\ref{eq:inv}) between reflection
coefficients. For a given $r\in\mathcal{R}_{s}$ with $\widetilde{r}$
of~\eqref{eq:tilder-def}, we define%
\[
t(z)=\frac{z}{z+i}\exp\left( \frac{1}{2\pi i}\int_{\mathbb{R}}\log\left[
\left( s^{2}+1\right) \widetilde{r}(s)\right] \frac{ds}{s-z}\right)
\]
for $\operatorname{Im}(z)>0$, and by the boundary value for real $z=k$.
\begin{proposition}
The mapping
\[
\mathcal{I}_{s}:r\mapsto-\frac{t(k)}{t(-k)}r(-k)
\]
is a continuous involution from $\mathcal{R}_{s}$ to itself.
\end{proposition}
We omit the proof, since it is completely analogous to that of Proposition~3.3
in~\cite{HMP:2010}, except that the Banach algebra $1\dotplus\widehat{X}$
there is replaced with the Banach algebra~$1\dotplus H^{s}({\mathbb{R}})$.
\section{The Inverse Problem}
\label{sec:inverse}
In this section, we assume given a function $r\in\mathcal{R}_{s}$ and set
$r^{\#}=\mathcal{I}_{s}r$. It follows from \cite{HMP:2010} that there exists a
unique distribution $q\in H^{-1}(\mathbb{R})$ with Riccati representatives
$u=u_{+}\in L^{2}(\mathbb{R})\cap L^{1}(\mathbb{R}^{+})$ and $u^{\#}=u_{-}\in
L^{2}(\mathbb{R})\cap L^{1}(\mathbb{R}^{-})$ so that the corresponding
Schr\"{o}dinger operator has $r$ and $r^{\#}$ as its right and left reflection
coefficients, respectively. We wish to show that the Riccati representatives
$u$ and $u^{\#}$ reconstructed from $r$ and $r^{\#}$ belong respectively to
$L^{2,s}(\mathbb{R}^{+})$ and $L^{2,s}(\mathbb{R}^{-})$. To do so, we will
recall the reconstruction formulas for $u$ and $u^{\#}$ derived in
\cite{HMP:2010} from the Gelfand--Levitan--Marchenko equations. Let us define%
\begin{align*}
F(x) & =\frac{1}{\pi}\int_{-\infty}^{\infty}r(k)e^{2ikx}~dk,\\
F^{\#}(x) & =\frac{1}{\pi}\int_{-\infty}^{\infty}r^{\#}(k)e^{-2ikx}dk.
\end{align*}
Note that $F$ and $F^{\#}$ belong to $L^{2,s}(\mathbb{R})$. Setting%
\[
\Omega(x)=\left(
\begin{array}
[c]{cc}%
0 & F(x)\\
F(x) & 0
\end{array}
\right) ,~~\Omega^{\#}(x)=\left(
\begin{array}
[c]{cc}%
0 & F^{\#}(x)\\
F^{\#}(x) & 0
\end{array}
\right) ,
\]
the right and left Gelfand--Levitan--Marchenko equations are respectively
\begin{align*}
\Omega(x+\zeta)+\Gamma(x,\zeta)+\int_{0}^{\infty}\Gamma(x,t)\Omega
(x+\zeta+t)~dt & =0,%
\qquad
\zeta>0,\\
\Omega^{\#}(x+\zeta)+\Gamma^{\#}(x,\zeta)+\int_{0}^{\infty}\Gamma
^{\#}(x,t)\Omega^{\#}(x+\zeta+t)~dt & =0,%
\qquad
\zeta<0,
\end{align*}
for the $2\times2$ matrix-valued kernels $\Gamma$ and $\Gamma^{\#}$. The right
and left Riccati representatives are reconstructed via%
\begin{align*}
u(x) & =-\Gamma_{12}(x,0),\\
u^{\#}(x) & =\Gamma_{12}^{\#}(x,0).
\end{align*}
Let $\gamma(x,\zeta)=\Gamma_{12}(x,\zeta)$ and $\gamma^{\#}(x,\zeta
)=\Gamma_{12}^{\#}(x,\zeta)$. Let $T_{F}$ and $T_{F^{\#}}$ be the integral
operators (depending parametrically on $x$)%
\begin{align*}
\left( T_{F}\psi\right) (\zeta) & =\int_{0}^{\infty}F(x+\zeta
+t)\psi(t)~dt,\\
\left( T_{F^{\#}}\psi\right) (\zeta) & =\int_{-\infty}^{0}F^{\#}%
(x+\zeta+t)\psi(t)~dt.
\end{align*}
Then, as vectors in $L^{2}(\mathbb{R}^{+})$ (resp. in $L^{2}(\mathbb{R}^{-})$)
for each fixed $x$,
\begin{align*}
\left( I-T_{F}^{2}\right) \gamma(x,~\cdot~) & =-F(x+~\cdot~),\\
\left( I-T_{F^{\#}}^{2}\right) \gamma^{\#}(x,~\cdot~) & =-F^{\#}%
(x+~\cdot~).
\end{align*}
As shown in the proof of \cite{HMP:2010}, Proposition 4.2, the operator
$\left( I-T_{F}^{2}\right) ^{-1}$ is bounded from $L^{2}(\mathbb{R}^{+})$ to
itself (resp. $\left( I-T_{F^{\#}}^{2}\right) ^{-1}$ is bounded from
$L^{2}(\mathbb{R}^{-})$ to itself). From these equations and the
reconstruction formulas, it is not difficult to see that
\begin{align*}
u(x) & =F(x)-G(x),\\
u^{\#}(x) & =-F^{\#}(x)+G^{\#}(x),
\end{align*}
where%
\begin{align*}
G(x) & =\int_{0}^{\infty}F(x+t)H(x,t)~dt,\\
G^{\#}(x) & =\int_{-\infty}^{0}F^{\#}(x+t)H^{\#}(x,t)~dt
\end{align*}
and
\begin{align*}
H(x,~\cdot~) & =\left( I-T_{F}^{2}\right) ^{-1}\left( (T_{F}%
F)(x+~\cdot~)\right) ,\\
H^{\#}(x,~\cdot~) & =\left( I-T_{F^{\#}}^{2}\right) ^{-1}\left(
(T_{F^{\#}}F^{\#})(x+~\cdot~)\right) .
\end{align*}
We are interested in estimating the behavior of $G$ as $x\rightarrow+\infty$
(resp. of $G^{\#}$ as $x\rightarrow-\infty$). It suffices to consider
$x>x_{0}$ (resp. $x<-x_{0}$) for sufficiently large $x_{0}$. Choosing $x_{0}$
so large that
\[
\int_{x_{0}}^{\infty}\left\vert F(s)\right\vert ~ds<1/2,~\int_{-\infty}%
^{x_{0}}\left\vert F^{\#}(s)\right\vert ~ds<1/2,
\]
we have $\left\Vert T_{F}\right\Vert _{L^{p}\rightarrow L^{p}}<1$ for $p=1,2$,
and similarly for $T_{F^{\#}}$. Note that we can make such a choice of fixed
$x_{0}$ in a small neighborhood of a given $F\in L^{2,s}(\mathbb{R})$ since
$L^{2,s}(\mathbb{R})\subset L^{1}(\mathbb{R})$ for $s>1/2$. We can then obtain
convergent multilinear expansions for $G$ and $G^{\#}$ valid respectively for
$x>x_{0}$ and $x<-x_{0}$. These multilinear expansions can be estimated, much
as in the previous section, to obtain the required weighted estimates. We will
give the analysis for $G$ since the analysis for $G^{\#}$ is very similar.
For $x>x_{0}$ we have the expansion%
\[
H(x,~\cdot~)=\sum_{j=0}^{\infty}\left( T_{F}^{2j+1}\left[ F(x+~\cdot
~)\right] \right) (~\cdot~)
\]
convergent in $L^{2}(\mathbb{R}^{+})$. From this expansion and the Cauchy--Schwarz
inequality it follows that
\[
G(x)=\sum_{n=1}^{\infty}G_{n}(x)
\]
in $L^{\infty}(x_{0},\infty)$, where%
\[
G_{n}(x)=\int_{\mathbb{R}_{+}^{2n}}F(x+t_{1})F(x+t_{1}+t_{2})\ldots
F(x+t_{2n-1}+t_{2n})F(x+t_{2n})~\mathrm{dt}%
\]
and $\mathrm{dt}:=dt_{1}\ldots dt_{2n}$. We will show that, for $x_{0}>0$,
\begin{equation}
\int_{x_{0}}^{\infty}\left( 1+x\right)^{2s}
\left\vert G_{n}(x)\right\vert^{2}~dx
\leq\left\Vert F\right\Vert _{L^{1}(x_{0},\infty)}^{4n}
\int_{x_{0}}^{\infty}\left( 1+x\right)^{2s}
\left\vert F(x)\right\vert ^{2}~dx,
\label{eq:Gn.bound}%
\end{equation}
from which it follows that $\int_{0}^{\infty}\left(1+x\right)^{2s}
\left\vert G(x)\right\vert^{2}~dx<\infty$. \ Let $f(x):=\left\vert
F(x)\right\vert $ and $\widetilde{f}(x):=\left( 1+x\right) ^{2s}f^{2}(x)$.
Since $x\leq x+t_{1}$ in the range of integration for $G_{n}$, it follows from
the Cauchy--Schwarz inequality that%
\[
\int_{x_{0}}^{\infty}\left( 1+x\right) ^{2s}\left\vert G_{n}(x)\right\vert
^{2}~dx\leq\int_{x_{0}}^{\infty}I_{n}(x)J_{n}(x)~dx,
\]
where%
\begin{equation}
I_{n}(x):=\int_{\mathbb{R}_{+}^{2n}}\widetilde{f}(x+t_{1})f(x+t_{1}%
+t_{2})\ldots f(x+t_{2n-1}+t_{2n})f(x+t_{2n})~\mathrm{dt}\label{eq:In}%
\end{equation}
and%
\begin{equation}
J_{n}(x):=\int_{\mathbb{R}_{+}^{2n}}f(x+t_{1}+t_{2})\ldots f(x+t_{2n-1}%
+t_{2n})f(x+t_{2n})~\mathrm{dt}.\label{eq:Jn}%
\end{equation}
Clearly,
\begin{equation}
J_{n}(x)\leq\left\Vert f\right\Vert _{L^{1}(x_{0},\infty)}^{2n}%
\label{eq:Jn.bound}%
\end{equation}
for $x\geq x_{0}$. In (\ref{eq:In}), set $y_{2k-1}=x+t_{2k-1}$ and
$y_{2k}=t_{2k}$ for $1\leq k\leq n$; then
\begin{multline*}
\int_{x_{0}}^{\infty}I_{n}(x)~dx=\\
\int_{x_{0}}^{\infty}\int_{0}^{\infty}\ldots\int_{0}^{\infty}\int_{x_{0}%
}^{\infty}\widetilde{f}(y_{1})f(y_{1}+y_{2})\ldots f(y_{2n-1}+y_{2n}%
)f(x+y_{2n})~\mathrm{dy}~dx
\end{multline*}
where $\mathrm{dy}:=dy_{1}\ldots dy_{2n}$. It follows easily that
\begin{equation}
\int_{x_{0}}^{\infty}I_{n}(x)~dx\leq\left( \int_{x_{0}}^{\infty}\widetilde
{f}(x)~dx\right) \left( \int_{x_{0}}^{\infty}f(x)~dx\right) ^{2n}%
.\label{eq:In.bound}%
\end{equation}
Combining (\ref{eq:In.bound}) and (\ref{eq:Jn.bound}) gives (\ref{eq:Gn.bound}).
Together with a similar analysis for $u^{\#}$ and $G^{\#}$, the above
arguments yield:
\begin{proposition}
\label{prop:inverse}Suppose that $r\in\mathcal{R}_{s}$ for $s>1/2$. Then $u\in
L^{2,s}(\mathbb{R}^{+})$ and $u^{\#}\in L^{2,s}(\mathbb{R}^{-})$, and the maps
$r\mapsto u$ and $r^{\#}\mapsto u^{\#}$ are locally Lipschitz continuous
respectively as maps $\mathcal{R}_{s}\rightarrow L^{2,s}(\mathbb{R}^{+})$ and
$\mathcal{R}_{s}\rightarrow L^{2,s}(\mathbb{R}^{-})$.
\end{proposition}
\section{Proof of the Main Theorem}
\label{sec:proof}
We now give the proof of Theorem \ref{thm:main}. Proposition \ref{prop:direct}
shows that $\mathcal{S}_{\pm}$ have range contained in $\mathcal{R}_{s}$ and
that $\mathcal{S}_{\pm}$ are locally Lipschitz continuous maps from
$L^{2,s}(\mathbb{R}^{-})\times L^{2,s}(\mathbb{R}^{+})\times(0,\infty)$ into
the space $\mathcal{R}_{s}$. On the other hand, given a reflection coefficient
$r\in\mathcal{R}_{s}$, Proposition \ref{prop:inverse} shows that the Riccati
representatives reconstructed from $r$ and $r^{\#}$ satisfy $u\in
L^{2,s}(\mathbb{R}^{+})$ and $u^{\#}\in L^{2,s}(\mathbb{R}^{-})$ and are
locally Lipschitz continuous as respective functions of $r$ and $r^{\#}$. It
follows from the analysis of section 4 in \cite{HMP:2010} that $u$ and
$u^{\#}$ are the unique right- and left-hand Riccati representatives of a
real-valued distribution $q\in H^{-1}(\mathbb{R})$ having reflection
coefficients $r$ and $r^{\#}$. This shows that $\mathcal{S}_{\pm}$ are onto
$\mathcal{R}_{s}$ and completes the proof of Theorem \ref{thm:main}.
|
\section{Introduction}
The purpose of this paper is to prove Manin's conjecture
about points of bounded height for a family of Ch\^atelet surfaces
over ${\mathbf Q}$. These surfaces have been considered by F.~Ch\^atelet
in \cite{chatelet:points} and \cite{chatelet:points2},
by V.~A.~Iskovskikh~\cite{iskovskih:hasse},
by D.~Coray and M.~A. Tsfasman \cite{coraytsfasman:delpezzo},
and by {J.-L.} Colliot-Th\'el\`ene, J.-J. Sansuc, and P. Swinnerton-Dyer
in \cite{ctssd:chatelet1} and \cite{ctssd:chatelet2}, among others.
The surfaces considered here are smooth proper models
of the affine surfaces given in ${\mathbf A}^3_{\mathbf Q}$ by an equation
of the form
\begin{equation*}
Y^2+Z^2=X(a_3X+b_3)(a_4X+b_4),
\end{equation*}
for suitable $a_3$, $b_3$, $a_4$, $b_4\in{\mathbf Z}$.
It is important to note that the surfaces we consider
do not satisfy weak approximation,
the lack of which
is explained by the Brauer-Manin obstruction, as described in
\cite{ctssd:chatelet1} and \cite{ctssd:chatelet2}.
Up to now, the only cases for which Manin's principle was
proven despite weak approximation not holding
were obtained using harmonic analysis and required the
action of an algebraic group on the variety with
an open orbit. The method used in this paper is completely
different. Following ideas of P.~Salberger \cite{salberger:tamagawa},
we use versal torsors introduced by
Colliot-Th\'el\`ene and Sansuc in \cite{cts:predescente2},
\cite{cts:descente1}, and \cite{cts:descente2}
to estimate the number of rational points of bounded
height on the surface.
This paper is organised as follows: in section~\ref{section.geometry},
we recall some facts about the geometry of the surfaces. In
section~\ref{section.result}, we define the height and state
our main result. Section~\ref{section.torsors} contains
the description of the versal torsors we use. In
section~\ref{section.jumpingup}, we describe the lifting of rational
points to the versal torsors. This lifting reduces the initial
problem to the estimation of some arithmetic sums denoted by
${\mathcal U}(T)$. The following sections contain
the key analytical tools used in the proof.
In section~\ref{section.bound} we give a uniform upper bound
for ${\mathcal U}(T)$ and in section~\ref{section.estimate}
an asymptotic formula for it.
The last section is devoted to an interpretation of the leading
constant.
Let us fix some notation for the remainder of this text.
\begin{notaconv}
If $k$ is a field, we denote by $\overline k$ an algebraic closure
of $k$. For any variety $X$ over $k$ and any $k$-algebra $A$, we denote
by $X_A$ the product $X\times_{\Spec(k)}\Spec(A)$ and by $X(A)$
the set $\Hom_{\Spec(k)}(\Spec(A),X)$. We also put $\overline X=X_{\overline k}$.
The cohomological Brauer group of $X$ is defined as
$\Br(X)=H^2_\textinmath{\'et}(X,\mathbf G_m)$, where $\mathbf G_m$ denotes
the multiplicative group.
The projective space of dimension $n$
over $A$ is denoted by ${\mathbf P}^n_A$ and the
affine space by ${\mathbf A}^n_A$. For any $(x_0,\dots,x_n)\in k^{n+1}\setminus\{0\}$
we denote by $(x_0:\dots:x_n)$ its image in ${\mathbf P}^n(k)$.
\end{notaconv}
\section{A family of Ch\^atelet surfaces}
\label{section.geometry}
Let us fix $a_1$, $a_2$, $a_3$, $a_4$, $b_1$, $b_2$,
$b_3, b_4\in{\mathbf Z}$ such that
\[\Delta_{i,j}=\left\vert\begin{matrix}
a_i&a_j\\
b_i&b_j
\end{matrix}\right\vert\neq 0\]
for any $i,j\in{\{1,2,3,4\}}$ with $i\neq j$.
We then consider the linear forms $L_i$ defined
by $L_i(U,V)=a_iU+b_iV$ for $i\in{\{1,2,3,4\}}$
and define the hypersurface $S_1$ of ${\mathbf P}^2_{\mathbf Q}\times{\mathbf A}^1_{\mathbf Q}$
given by the equation
\begin{equation*
X^2+Y^2=T^2\prod_{i=1}^4L_i(U,1)
\end{equation*}
and the hypersurface $S_2$ given by the equation
\begin{equation*
{X'}^2+{Y'}^2={T'}^2\prod_{i=1}^4L_i(1,V).
\end{equation*}
Let $U_1$ be the open subset of $S_1$ defined by $U\neq 0$
and $U_2$ be the open subset of $S_2$ defined by $V\neq 0$.
The map $\Phi:U_1\to U_2$ which maps $((X:Y:T),U)$
onto $((X:Y:U^2T),1/U)$ is an isomorphism and we define $S$
as the surface obtained by glueing $S_1$ to $S_2$ using the isomorphism
$\Phi$. The surface $S$ is a smooth projective surface and is
a particular case of a Ch\^atelet surface. The geometry of such surfaces
has been described by J.-L. Colliot-Th\'el\`ene, J.-J. Sansuc
and P. Swinnerton-Dyer in \cite[\S7]{ctssd:chatelet2}. For
the sake of completeness, let us recall part of this description which will
be useful for the description of versal torsors.
The maps $S_1\to{\mathbf P}^1_{\mathbf Q}$ (resp. $S_2\to{\mathbf P}^1_{\mathbf Q}$)
which maps $((X:Y:T),U)$ onto $(U:1)$
(resp. $((X':Y':T'),V)$ onto $(1:V)$) glue together
to give a conic fibration $\pi:S\to{\mathbf P}^1_{\mathbf Q}$
with four degenerate fibres over the points given by
$P_i=(-b_i:a_i)\in{\mathbf P}^1({\mathbf Q})$ for $i\in{\{1,2,3,4\}}$.
In fact, the glueing of ${\mathbf P}^2_{\mathbf Q}\times{\mathbf A}^1_{\mathbf Q}$ to
${\mathbf P}^2_{\mathbf Q}\times{\mathbf A}^1_{\mathbf Q}$ through the map
\begin{equation}\label{equ.glueing}
((X:Y:T),U)\mapsto ((X:Y:U^2T),1/U)
\end{equation}
gives the projective bundle\footnote{We define
here ${\mathbf P}(\mathcal O^2\oplus\mathcal O(-2))$ as the projective
bundle associated to the sheave of graded commutative algebras
$\underline{\Sym}(\mathcal O^2\oplus\mathcal O(2))$. In other
words the fibre over a point is given by the lines in the
fibre of the vector bundle and not by the hyperplanes.}
${\matheusm P}={\mathbf P}(\mathcal O^2\oplus\mathcal O(-2))$
over ${\mathbf P}^1_{\mathbf Q}$ and $S$ may be seen as a hypersurface
in that bundle.
Over ${\mathbf Q}({\mathsl i})$, if $\xi\in\{-{\mathsl i},{\mathsl i}\}$, the map
${\mathbf A}_{{\mathbf Q}({\mathsl i})}\to {S_1}_{{\mathbf Q}({\mathsl i})}$ given by $u\mapsto ((\xi:1:0),U)$
extends to a section $\sigma_\xi$ of $\pi$. The surface
$S_{{\mathbf Q}({\mathsl i})}$ contains $10$ exceptional curves, that is irreducible
curves with negative self-intersection. Eight of them are given
in $S_{{\mathbf Q}({\mathsl i})}$ by the following equations
\begin{equation*
D_j^\xi:\qquad L_j(\pi(P))=0\quad\text{and}\quad X-\xi Y=0
\end{equation*}
for $\xi\in\{-{\mathsl i},{\mathsl i}\}$ and $j\in{\{1,2,3,4\}}$; the last ones
correspond to the section $\sigma_\xi$ and are given by the\
equations
\begin{equation*
E^\xi:\qquad T=0\quad\text{and}\quad X-\xi Y=0.
\end{equation*}
Here $X$, $Y$ and $T$ are seen as sections of
$\mathcal O_{{\matheusm P}}(1)$.
Let us denote by ${\mathcal G}$ the Galois group of ${\mathbf Q}({\mathsl i})$
over ${\mathbf Q}$ and by $z\mapsto \overline z$ the nontrivial element
in ${\mathcal G}$. Then we have
\[\overline{E^\xi}=E^{\overline\xi}\quad\text{and}\quad
\overline{D_j^\xi}=D_j^{\overline\xi}\]
for $\xi\in\{-{\mathsl i},{\mathsl i}\}$ and $j\in{\{1,2,3,4\}}$. We shall also
write $D_j^+$ (resp. $D_j^-$, $E^+$, $E^-$)
for $D_j^{{\mathsl i}}$ (resp. $D_j^{-{\mathsl i}}$, $E^{{\mathsl i}}$, $E^{-{\mathsl i}}$).
The intersection multiplicities of these divisors are given by
\[(E^\xi,E^\xi)=-2,\quad (D_j^\xi,D_j^\xi)=-1,\quad (D_j^\xi,D_j^{-\xi})=1,
\quad(E^\xi,D_j^\xi)=1,\]
where $\xi\in\{-{\mathsl i},{\mathsl i}\}$, and $j\in{\{1,2,3,4\}}$, all other
intersection multiplicities
being equal to~$0$. These intersections are summarized in
figure~\ref{fig.intersection}.
\begin{figure}[ht]
\[\begin{pspicture}(0,0)(10,6)
\psline(1,1)(9,1)
\rput(0.7,1){$E^-$}
\rput(8.7,1.2){$-2$}
\psline(1,5)(9,5)
\rput(0.7,5){$E^+$}
\rput(8.7,4.8){$-2$}
\psbezier(2,0)(2,1)(2,3)(1,4)
\rput(1.7,0){$D_1^-$}
\rput(2.1,2.1){-1}
\psbezier(4,0)(4,1)(4,3)(3,4)
\rput(3.7,0){$D_2^-$}
\rput(4.1,2.1){-1}
\psbezier(6,0)(6,1)(6,3)(5,4)
\rput(5.7,0){$D_3^-$}
\rput(6.1,2.1){-1}
\psbezier(8,0)(8,1)(8,3)(7,4)
\rput(7.7,0){$D_4^-$}
\rput(8.1,2.1){-1}
\psbezier(2,6)(2,5)(2,3)(1,2)
\rput(1.7,6){$D_1^+$}
\rput(2.1,3.9){-1}
\psbezier(4,6)(4,5)(4,3)(3,2)
\rput(3.7,6){$D_2^+$}
\rput(4.1,3.9){-1}
\psbezier(6,6)(6,5)(6,3)(5,2)
\rput(5.7,6){$D_3^+$}
\rput(6.1,3.9){-1}
\psbezier(8,6)(8,5)(8,3)(7,2)
\rput(7.7,6){$D_4^+$}
\rput(8.1,3.9){-1}
\end{pspicture}\]
\caption{Intersection multiplicities}
\label{fig.intersection}
\end{figure}
The geometric Picard group of $S$, that is $\Pic(\overline S)$,
is isomorphic to $\Pic(S_{{\mathbf Q}({\mathsl i})})$ and is generated by these exceptional
divisors with the relations
\begin{equation}\label{equ.pic.fibre}
[D_j^+]+[D_j^-]=[D_k^+]+[D_k^-]
\end{equation}
for $j,k\in{\{1,2,3,4\}}$ and
\begin{equation}\label{equ.pic.can}
[E^+]+[D_j^+]+[D_k^+]=[E^-]+[D_l^-]+[D_m^-]
\end{equation}
whenever $\{j,k,l,m\}={\{1,2,3,4\}}$.
In particular, a basis of $\Pic(S_{{\mathbf Q}({\mathsl i})})$ is
given by the family
\[([E^+],[D_1^+],[D_2^+],[D_3^+],[D_4^+],[D_1^-])\]
and the rank of the geometric Picard group of $S$ is equal
to $6$.
Using the fact that $\Pic(S)=(\Pic(S_{{\mathbf Q}({\mathsl i})}))^{{\mathcal G}}$ it is
easy to deduce that $\Pic(S)$ has rank $2$.
The class of the anticanonical line bundle is given by
\[\omega_S^{-1}=2E^++\sum_{j=1}^4D_j^+=2E^-+\sum_{j=1}^4D_j^-.\]
Indeed, by the adjunction formula, for any curve $C$ in $S$ of genus
$g$, one has the relation
$[C].([C]+\omega_S)=2g-2$. Therefore if $\xi\in\{-{\mathsl i},{\mathsl i}\}$
and $j\in{\{1,2,3,4\}}$,
\[[D_j^\xi].\omega_S^{-1}=1\quad\text{and}\quad[E^\xi].\omega_S^{-1}=0.\]
It is worthwhile noting
that $\omega_S^{-1}=
\mathcal O_{{\matheusm P}}(1)$.
\begin{lemma}
\label{lemma.sectionsomega}%
Using the trivialisation described by \eqref{equ.glueing},
the $5$-tuple of functions
\[(T,UT,U^2T,X,Y)\]
gives
a basis of $\Gamma(S,\omega_S^{-1})$.
\end{lemma}
\begin{proof}
Let $C$ be a generic divisor in $\vert\omega_S^{-1}\vert$. Then
$C$ is a smooth irreducible curve; let
$g_C$ be its genus. According to the adjunction formula,
we have that $2g_C-2=\omega_S.(\omega_S-\omega_S)=0$. Thus $g_C=1$.
The exact sequence of sheaves
\[0\longrightarrow\mathcal O_S\longrightarrow\omega_S^{-1}
\longrightarrow\omega_S^{-1}\otimes\mathcal O_C\longrightarrow 0\]
gives an exact sequence
\[0\longrightarrow H^0(S,\mathcal O_S)\longrightarrow
H^0(S,\omega_S^{-1})\longrightarrow H^0(C,{\omega_S^{-1}}_{\vert C})
\longrightarrow H^1(S,\mathcal O_S).\]
But $S$ is geometrically rational and $H^1(S,\mathcal O_S)=\{0\}$.
We get that
\[h^0(S,\omega_S^{-1})=1+h^0(C,{\omega_S^{-1}}_{\vert C}).\]
Let $D={\omega_S^{-1}}_{\vert C}$. We have that $\deg(D)=4$ and
$\deg(\omega_C-D)=-4$ since $\omega_C=0$.
Applying Riemann--Roch theorem to $C$, we get that
\[h^0(D)=\deg(D)+2g_C-2=4\]
and $h^0(S,\omega_S^{-1})=5$. Since the sections $T,UT,U^2T,X$ and $Y$
are linearly independent, and extend to a section
of $\mathcal O_{{\matheusm P}}(1)$, we get
a basis of $\Gamma(S,\omega_S^{-1})$.
\end{proof}
\begin{lemma}
\label{lemma.linearomega}%
The linear system $\vert\omega_S^{-1}\vert$ has no base point and the
basis given in lemma~\ref{lemma.sectionsomega} gives a morphism
from $S$ to ${\mathbf P}^4_{\mathbf Q}$, the image of which is the surface $S'$
given by the system of equations
\begin{equation*
\begin{cases}
X_0X_2-X_1^2=0\\
X_3^2+X_4^2=(aX_0+bX_1+cX_2)(a'X_0+b'X_1+c'X_2)
\end{cases}
\end{equation*}
where
\begin{align*}
a&=a_1a_2,&b&=a_1b_2+a_2b_1,&c&=b_1b_2,\\
a'&=a_3a_4,&b'&=a_3b_4+a_4b_3,&c'&=b_3b_4.
\end{align*}
The induced map $\psi:S\to S'$ is the blowing up
of the conjugate singular points of $S'$ given by
$P^\xi=(0:0:0:1:-\xi)$ with $\xi^2=-1$ and $\psi^{-1}(P^\xi)=E^{\xi}$.
\end{lemma}
\begin{proof}
This follows from the fact that the map from $S$ to ${\mathbf P}^4_{\mathbf Q}$
induces the maps
\[((x:y:t),u)\longmapsto (t:ut:u^2t:x:y)\]
from $S_1$ to ${\mathbf P}^4_{\mathbf Q}$ and
\[((x':y':t'),v)\longmapsto (v^2t':vt':t':x':y')\]
from $S_2$ to ${\mathbf P}^4_{\mathbf Q}$.
\end{proof}
\begin{rema}
The surface $S'$ is an Iskovskikh surface
\cite{coraytsfasman:delpezzo};
it is a singular Del Pezzo surface of degree $4$ with a singularity
of type $2A_1$ and $\psi:S\to S'$ is a minimal resolution
of singularities for $S'$.
\end{rema}
We finish this section by a brief reminder of the description
of the Brauer group of $S$.
\begin{lemma}
\label{lem:brauer}%
The cokernel of the morphism from the Brauer group of
${\mathbf Q}$ to the Brauer group of $S$ is isomorphic to the Klein group
$({\mathbf Z}/2{\mathbf Z})^2$ and the image of the natural injective map
\[\Br(S)/\Br({\mathbf Q})\longrightarrow \Br({\mathbf Q}(S))/\Br({\mathbf Q})\]
is generated by the elements $(-1,L_j(U,V)/L_k(U,V))$
for $j,k\in{\{1,2,3,4\}}$.
\end{lemma}
\begin{proof}
By \cite[lemma 6.3]{sansuc:brauer} and the fact that $\Pic(S)$
coincides with $\Pic(S_{{\mathbf Q}({\mathsl i})})^{\mathcal G}$, there is an exact sequence
\[0\longrightarrow \Br({\mathbf Q})\longrightarrow\ker(\Br(S)\longrightarrow
\Br(\overline S))\longrightarrow H^1(\Gal(\overline{\mathbf Q}/{\mathbf Q}),
\Pic(\overline S))\longrightarrow 0.\]
Since $\overline S$ is rational and the Brauer group is a
birational invariant of smooth projective varieties, we get
that the cokernel of the morphism $\Br({\mathbf Q})\to\Br(S)$ is isomorphic
to the cohomology group
$H^1(\Gal(\overline {\mathbf Q}/{\mathbf Q}),\Pic(\overline S))$. But
the group $H^1(\Gal(\overline {\mathbf Q}/{\mathbf Q}({\mathsl i})),\Pic(\overline S))$
is trivial and we are reduced to computing the group
$H^1({\mathcal G},\Pic(S_{{\mathbf Q}({\mathsl i})}))$.
Since ${\mathcal G}$ is cyclic of order $2$, this cohomology
group coincides with the homology of the complex
{\CDat
\[\Pic(S_{{\mathbf Q}({\mathsl i})})@>\Id-\sigma>>
\Pic(S_{{\mathbf Q}({\mathsl i})})@>\Id+\sigma>>\Pic(S_{{\mathbf Q}({\mathsl i})})\]}%
where $\sigma$ denotes the complex conjugation.
By the description of the action of $\sigma$, the ${\mathbf Z}$-module
$\ker(\Id+\sigma)$ has a basis given by
\[([D_1^+]-[D_2^+],[D_2^+]-[D_3^+],[D_3^+]-[D_4^+],
[D_1^+]-[D_1^-]).\]
On the other hand, $\im(\Id-\sigma)$ is generated by
\begin{gather*}
[D_1^+]-[D_1^-],\qquad 2[D_2^+]-[D_1^+]-[D_1^-],\\
2[D_3^+]-[D_1^+]-[D_1^-]\quad
\text{and}\quad
2[D_4^+]-[D_1^+]-[D_1^-].
\end{gather*}
Thus the quotient is isomorphic to $({\mathbf Z}/2{\mathbf Z})^2$ and generated
by the classes of elements of the form $[D_j^+]-[D_k^+]$
with $j,k\in{\{1,2,3,4\}}$.
It remains to describe the images of the classes in the Brauer
group of the function field ${\mathbf Q}(S)$. But the isomorphism
\[H^1(\Gal(\overline{\mathbf Q}/{\mathbf Q}),\Pic(\overline S))\longrightarrow
\Br(S)/\Br({\mathbf Q})\]
may be described as follows:
let us consider the exact sequence of $\Gal(\overline {\mathbf Q}/{\mathbf Q})$-modules:
{\CDat
\[0\longrightarrow\overline{\mathbf Q}^*\longrightarrow\overline {\mathbf Q}(S)^*@>\divi>>
\Div(\overline S)\longrightarrow\Pic(\overline S)\longrightarrow 0\]}%
which yields two short exact sequences:
\[0\longrightarrow\overline{\mathbf Q}^*\longrightarrow\overline {\mathbf Q}(S)^*\longrightarrow
\overline{\mathbf Q}(S)^*/\overline{\mathbf Q}^*\longrightarrow 0\]
and
\[0\longrightarrow\overline {\mathbf Q}(S)^*/\overline{\mathbf Q}^*\longrightarrow
\Div(\overline S)\longrightarrow\Pic(\overline S)\longrightarrow 0.\]
Taking the corresponding cohomology long exact sequences
we get exact sequences
\[0\longrightarrow H^1(\Gal(\overline {\mathbf Q}/{\mathbf Q}),\Pic(\overline S))
\buildrel\partial\over\longrightarrow H^2(\Gal(\overline{\mathbf Q}/{\mathbf Q}),
\overline{\mathbf Q}(S)^*/\overline{\mathbf Q}^*)\]
and
\[0\longrightarrow\Br({\mathbf Q})\longrightarrow \Br({\mathbf Q}(S))
\longrightarrow H^2(\Gal(\overline{\mathbf Q}/{\mathbf Q}),\overline{\mathbf Q}(S)^*/\overline{\mathbf Q}^*)
\longrightarrow 0\]
and using the natural injection $\Br(S)\to\Br({\mathbf Q}(S))$
we get an isomorphism from the image of $\partial$
to $\coker(\Br({\mathbf Q})\to\Br(S))$. But if $D$ is a divisor
on $S$ such that its class $[D]$ belongs to $\ker(1+\sigma)$ and represents
$\alpha\in H^1(\Gal(\overline {\mathbf Q}/{\mathbf Q}),\Pic(\overline S))$
then
\[(1+\sigma)D\in\ker(\Div(\overline S)\to\Pic(\overline S))
\cap\Div(S).\]
Therefore $(1+\sigma)D=\divi(f)$ for a function $f$ in ${\mathbf Q}(S)^*$
and $\partial(\alpha)$ coincides with the image of $(-1,f)$.
In our particular case, we get that
\[(1+\sigma)(D_j^+-D_k^+)=D_j^++D_j^--D_k^+-D_k^-
=\divi(L_j(U,V)/L_k(U,V))\]
which concludes the proof.
\end{proof}
\section{Points of bounded height}
\label{section.result}
Over $\overline{\mathbf Q}$ or even ${\mathbf Q}({\mathsl i})$, the only geometrical
invariant of $S$ is the cross-ratio
\[\alpha=\frac{\left\vert
\begin{matrix}
a_3&a_1\\
b_3&b_1
\end{matrix}
\right\vert\left/\left\vert
\begin{matrix}
a_3&a_2\\
b_3&b_2
\end{matrix}
\right\vert\right.}{\left\vert
\begin{matrix}
a_4&a_1\\
b_4&b_1
\end{matrix}
\right\vert\left/\left\vert
\begin{matrix}
a_4&a_2\\
b_4&b_2
\end{matrix}
\right\vert\right.}\quad\in{\mathbf Q}.\]
Indeed the automorphisms of ${\mathbf P}^1_{\mathbf Q}$
sending the points $P_1$, $P_2$, $P_3$
onto $\infty=(0:1)$, $0=(1:0)$ and $1=(1:1)$
lifts to an isomorphism from $S$ to the Ch\^atelet
surface with an equation of the form
\[X^2+Y^2=\beta U(U-1)(U-\alpha)T^2\]
where $\beta\in{\mathbf Q}$. Over ${\mathbf Q}({\mathsl i})$ we may further reduce
to the case where $\beta=1$.
In particular, without any loss of generality, we may assume
that
\begin{equation}\label{equ.pthreefour}
a_1=b_2=1\qquad\text{and}\qquad a_2=b_1=0.
\end{equation}
\begin{hypo
From now on we assume the relations \eqref{equ.pthreefour},
that we have $\gcd(a_3,b_3)=\gcd(a_4,b_4)=1$, and that
$a_3b_3a_4b_4(a_3b_4-a_4b_3)\neq 0$.
\end{hypo}
\begin{nota}
\label{nota.height}%
Let $C=\sqrt{\prod_{j=1}^4(\vert a_j\vert+\vert b_j\vert)}$.
We equip the projective space ${\mathbf P}^4_{\mathbf Q}$ with the exponential height
$H_4:{\mathbf P}^4({\mathbf Q})\to{\mathbf R}$
defined by
\begin{equation*
H_4(x_0:x_1:x_2:x_3:x_4)=
\max\left(\vert x_0\vert,\vert x_1\vert,\vert x_2\vert,
\frac{\vert x_3\vert}C,\frac{\vert x_4\vert}C\right)
\end{equation*}
if $x_0,\dots,x_4$ are coprime integers. Using the morphism
$\psi:S\to S'$, we get a height $H=H_4\circ\psi$
which is associated to the anticanonical line bundle $\omega_S^{-1}$.
\par
We denote by $\Val({\mathbf Q})$ the set of places of ${\mathbf Q}$.
For any $v\in\Val({\mathbf Q})$, ${\mathbf Q}_v$ is the corresponding
completion of ${\mathbf Q}$.
As explained in \cite[\S2]{peyre:fano}, such a height
enables us to define a Tamagawa measure ${\boldsymbol\omega}_H$ on the adelic space
$S({\boldsymbol A}_{\mathbf Q})=\prod_{v\in\Val({\mathbf Q})}S({\mathbf Q}_v)$.
We also consider the constant $\alpha(S)$ defined
in \cite[definition 2.4]{peyre:fano} which is equal to $1$
in our particular case and, following Batyrev and
Tschinkel~\cite{batyrevtschinkel:toric},
we also put
\[\beta(S)=\sharp\bigl(\coker(\Br({\mathbf Q})\to\Br(S))\bigr)=4,\]
by lemma \ref{lem:brauer}. We then set
\[C_H(S)=\alpha(S)\beta(S){\boldsymbol\omega}_H(S({\boldsymbol A}_{\mathbf Q})^{\Br})\]
where $S({\boldsymbol A}_{\mathbf Q})^{\Br}$ is
the set of points in the adelic space for which the Brauer-Manin obstruction
to weak approximation is trivial.
\par
We are interested in the asymptotic behaviour
of the number of points of bounded height in $S({\mathbf Q})$,
that is by the number
\[N_{S,H}(B)=\sharp\{\,P\in S({\mathbf Q}),\ H(P)\leq B\,\}\]
for $B\in{\mathbf R}$ with $B>1$.
\end{nota}
\begin{figure}[ht]
\drawing{iskovskih3D.eps}
\caption{Obstruction to weak approximation}\label{fig.zolidessin}
\end{figure}
As an illustration of our problem we have drawn in
figure~\ref{fig.zolidessin}
the set of points
\[\{\,P\in S({\mathbf Q}),\ H(P)\leq 2000\,\}\]
for the surface $S$ obtained with $a_2=b_1=0$,
$a_1=b_2=a_3=b_3=a_4=1$,
and $b_4=-1$.
The colour of a rational point $P=((y:z:t),u)$ is
black if $u/2^{v_2(u)}\equiv 1\mod 4$, white otherwise.
The fact that all black points are on one of the real
connected components of $S({\mathbf R})$ may be explained by
the Brauer-Manin obstruction to weak approximation.
\par
We can now state the main result of this paper.
\begin{theo}\label{theo.main}
For any Ch\^atelet surface as above,
we have the asymptotic formula
\begin{equation}\tag{F}
N_{S,H}(B)= C_H(S)B\log(B) +O\big(B \log (B)^{0.972}\big).
\end{equation}
\end{theo}
\begin{rems}
(i) One may note that, as $S({\mathbf Q})$ is dense in $S({\boldsymbol A}_{\mathbf Q})^{\Br}$
by \cite[theorem B]{ctssd:chatelet1},
this formula is compatible
with the empirical formula (F) described in
\cite[formule empirique 5.1]{peyre:lille}
which is a refinement of a conjecture of Batyrev and
Manin~\cite{batyrevmanin:hauteur}.
(ii) Over ${\mathbf R}$, the image of $S({\mathbf R})$ on ${\mathbf P}^1({\mathbf R})$ is
the union of two intervals defined by the conditions
$\prod_{j=1}^4L_j(U,V)>0$.
Therefore we may choose $j,k\in{\{1,2,3,4\}}$ such that $j\neq k$
and the sign of $L_j(U,V)L_k(U,V)$ is not constant on $S({\mathbf R})$.
The evaluation of the corresponding element $(-1,L_j(U,V)/L_k(U,V))\in\Br(S)$
(see lemma~\ref{lem:brauer}) is not constant on $S({\mathbf R})$. Therefore
in all the cases we consider,
\[S({\boldsymbol A}_{\mathbf Q})^{\Br}\neq S({\boldsymbol A}_{\mathbf Q}).\
\end{rems}
\section{Description of versal torsors}
\label{section.torsors}
Versal torsors were first
introduced by J.-L. Colliot-Th\'el\`ene and J.-J. Sansuc
in \cite{cts:predescente2}, \cite{cts:descente1}
and \cite{cts:descente2} as a tool to prove that the Brauer--Manin
obstruction to the Hasse principle and weak approximation
is the only one. In their setting, it is sufficient
to construct a variety which is birational
over the ground field to the versal torsors.
Such a construction for Ch\^atelet surfaces has
been carried out in \cite[\S7]{ctssd:chatelet2}.
Our purpose, however, is slightly different:
we want to parametrise the points of $S({\mathbf Q})$ using versal
torsors. Therefore we shall make the description of
\cite[\S7]{ctssd:chatelet2} slightly more precise
in the particular case we are considering and construct the
versal torsors with rational points as constructible subsets
of an affine space of dimension ten. Our construction is also
akin to the constructions based upon Cox rings.
We shall first introduce an intermediate versal torsor
which corresponds to the Picard group of $S$ over ${\mathbf Q}$,
that is to the maximal split quotient of $T_{\NS}$.
This intermediate torsor is easy to describe and shall
be useful in the parametrisation of the rational points.
\begin{defi}
Let ${\mathcal T}_\textinmath{spl}$ be the subscheme of ${\mathbf A}^5_{\mathbf Z}=\Spec({\mathbf Z}[X,Y,T,U,V])$
defined by the equation
\begin{equation}
\label{equ.torsiz}%
X^2+Y^2=T^2 \prod_{j=1}^4L_j(U,V)
\end{equation}
and the conditions
\[(X,Y,T)\neq 0\quad\text{and}\quad(U,V)\neq 0.\]
The split algebraic torus $T_{\textinmath{spl}}={\mathbf G}_{m,\ZZ}^2$ acts on ${\mathcal T}_\textinmath{spl}$
via the morphism of tori
\[(\lambda,\mu)\mapsto(\lambda,\lambda,\mu^{-2}\lambda,\mu,\mu)\]
from ${\mathbf G}_{m,\ZZ}^2$ to ${\mathbf G}_{m,\ZZ}^5$ and the natural action of ${\mathbf G}_{m,\ZZ}^5$
on ${\mathbf A}^5_{\mathbf Z}$. Let ${\matheusm T}_\textinmath{spl}$ be the variety ${\mathcal T}_{\textinmath{spl},{\mathbf Q}}$.
We have an obvious morphism $\pi_\textinmath{spl}$ from ${\matheusm T}_\textinmath{spl}$ to
$S$ which may be described as follows: for any extension ${\mathbf K}$ of ${\mathbf Q}$
and any point $(x,y,t,u,v)$ of ${\matheusm T}_\textinmath{spl}({\mathbf K})$,
if $v\neq 0$, then the point ${((x:y:tv^2),u/v)}$ belongs to ${S_1({\mathbf K})\subset
S({\mathbf K})}$. If $u\neq 0$ then the point
${((x:y:tu^2),v/u)}$ belongs to ${S_2({\mathbf K})\subset S({\mathbf K})}$
and the points obtained in $S({\mathbf K})$ coincide if $uv\neq0$.
The morphism $\pi_\textinmath{spl}$ makes of ${\matheusm T}_\textinmath{spl}$ a ${\mathbf G}_m^2$-torsor
over $S$.
\end{defi}
We now turn to the construction of the versal torsors.
\begin{nota}
We denote by ${\boldsymbol\Delta}$ the set of exceptional divisors
in $S_{{\mathbf Q}({\mathsl i})}$ and consider it as a ${\mathcal G}$-set. We then
consider the affine space ${\mathbf A}_{\boldsymbol\Delta}$ of dimension $10$ over ${\mathbf Q}$
defined by
\[{\mathbf A}_{\boldsymbol\Delta}=\Spec\left(({\mathbf Q}({\mathsl i})[Z_\delta,\delta\in{\boldsymbol\Delta}])^{\mathcal G}\right)\]
where ${Z_\delta,\delta\in\Delta}$ are ten variables.
We also consider the algebraic torus
\[T_\Delta=\Spec\left(({\mathbf Q}({\mathsl i})[Z_\delta,Z^{-1}_\delta,
\delta\in{\boldsymbol\Delta}])^{\mathcal G}\right).\]
We shall also write $Z_k^\varepsilon$ (resp. $Z_0^\varepsilon$)
for $Z_{D_k^\varepsilon}$ (resp. $Z_{E^\varepsilon}$).
Let ${\boldsymbol\Delta}_{\mathbf Q}$ be the set of ${\mathcal G}$-orbits in ${\boldsymbol\Delta}$.
We put $E=\{E^+,E^-\}$ and $D_j=\{D_j^+,D_j^-\}$ for $j\in{\{1,2,3,4\}}$.
Then ${\boldsymbol\Delta}_{\mathbf Q}=\{E,D_1,D_2,D_3,D_4\}$.
For $\delta\in{\boldsymbol\Delta}_{\mathbf Q}$, we may also write $\delta=\{\delta^+,\delta^-\}$
and we put
\[X_\delta=\frac 12(Z_{\delta^+}+Z_{\delta^-})\qquad\text{and}
\qquad Y_\delta=\frac 1{2{\mathsl i}}(Z_{\delta^+}-Z_{\delta^-}).\]
Then
\[({\mathbf Q}({\mathsl i})[Z_\delta,\delta\in{\boldsymbol\Delta}])^{\mathcal G}=
{\mathbf Q}[X_\delta,Y_\delta,\delta\in{\boldsymbol\Delta}_{\mathbf Q}].\]
\end{nota}
We now wish to construct for each isomorphism class of versal torsor
over $S$ with a rational point a representative
of this class in ${\mathbf A}_{\boldsymbol\Delta}$.
It follows from \cite[proposition~2]{cts:descente1} that the set
of isomorphism classes of such torsors is finite.
We first introduce a finite set which will be used to
parametrise this set of torsors.
\begin{nota}
Let ${\matheusm S}$ be the set of primes $p$ such that $p\mid\prod_{1\leq j<k\leq 4}
\Delta_{j,k}$\footnote{Over ${\mathbf Z}/2{\mathbf Z}$,
one of the $\Delta_{j,k}$ has to be zero,
and so $2 \in{\matheusm S}$.}. For any $j$ in ${\{1,2,3,4\}}$, we put
\[{\matheusm S}_j=\{\,p\in{\matheusm S},\ p\equiv 3\mod 4\quad
\text{and}\quad p\mid\prod_{k\neq j}
\Delta_{j,k}\,\}\]
and
\[\Sigma_j=\biggl\{(-1)^{\varepsilon_{-1}}\prod_{p\in{\matheusm S}_j}p^{\varepsilon_p},
(\varepsilon_{-1},(\varepsilon_p)_{p\in{\matheusm S}_j})
\in\{0,1\}\times\{0,1\}^{{\matheusm S}_j}\biggr\}.\]
Finally, we define $\Sigma$ to be the set of ${\boldsymbol{m}}=(m_j)_{1\leq j\leq 4}
\in \prod_{j=1}^4\Sigma_j$ such that the four integers are
relatively prime,
$m_1$ is positive and $\prod_{j=1}^4m_j$
is a square. For any ${\boldsymbol{m}}\in\Sigma$, we denote by $\alpha_{\boldsymbol{m}}$
the positive square root of $\prod_{j=1}^4m_j$.
Let ${\boldsymbol{m}}$ belong to $\Sigma$. We denote by ${\matheusm T}_{\boldsymbol{m}}$
the constructible subset of ${\mathbf A}_{\boldsymbol\Delta}$ defined by the equations
\begin{equation}\label{equ.tors.closed}
\Delta_{j,k}m_lZ_l^+Z_l^-+\Delta_{k,l}m_jZ_j^+Z_j^-
+\Delta_{l,j}m_kZ_k^+Z_k^-=0
\end{equation}
if $1\leq j<k<l\leq 4$ and the inequalities
\begin{equation}\label{equ.tors.open}
(Z_{\delta_1},Z_{\delta_2})\neq(0,0)
\end{equation}
whenever $\delta_1\cap\delta_2=\emptyset$.
Note that these conditions are invariant under the action
of the Galois group ${\mathcal G}$. Thus ${\matheusm T}_{\boldsymbol{m}}$ is defined
over ${\mathbf Q}$.
We then define a morphism $\pi_{\boldsymbol{m}}:{\matheusm T}_{\boldsymbol{m}}\to S$.
In order to do this, it is enough to define a morphism
$\widehat \pi_{\boldsymbol{m}}:{\matheusm T}_{\boldsymbol{m}}\to{\matheusm T}_\textinmath{spl}$ which is done
as follows: for any extension ${\mathbf K}$ of ${\mathbf Q}$ and any
${\boldsymbol{z}}=(z_\delta)_{\delta\in\Delta}$ in ${\matheusm T}_{\boldsymbol{m}}({\mathbf K})$, the
conditions~\eqref{equ.tors.closed} and \eqref{equ.tors.open}
ensure that there exists a pair $(u,v)\in {\mathbf K}^2\setminus\{0\}$
such that
\begin{equation}\label{equ.tors.uv}
L_j(u,v)=m_jz_j^+z_j^-
\end{equation}
for $j\in{\{1,2,3,4\}}$. Let $(x,y,t)\in {\mathbf K}^3\setminus\{0\}$ be given
by the conditions
\begin{equation}\label{equ.tors.xyz}
\begin{cases}
x+{\mathsl i} y=\alpha_{\boldsymbol{m}} (z_0^+)^2\prod_{j=1}^4z_j^+,\\
x-{\mathsl i} y=\alpha_{\boldsymbol{m}} (z_0^-)^2\prod_{j=1}^4z_j^-,\\
t=z_0^+z_0^-.
\end{cases}
\end{equation}
Then we have the relation
\[x^2+y^2=t^2\prod_{j=1}^4L_j(u,v).\]
and $(x,y,t,u,v)$ belongs to ${\matheusm T}_\textinmath{spl}({\mathbf K})$.
\par
It remains to describe the action of the torus
$T_{\NS}$ associated to the ${\mathcal G}$-lattice $\Pic(\overline S)$
on ${\matheusm T}_{\boldsymbol{m}}$. The algebraic torus
$T_{\boldsymbol\Delta}$ corresponds to the ${\mathcal G}$-lattice ${\mathbf Z}^{\boldsymbol\Delta}$ and $T_{\boldsymbol\Delta}$
acts by multiplication of the coordinates on ${\mathbf A}_{\boldsymbol\Delta}$.
The natural surjective morphism of ${\mathcal G}$-lattices
\[-\pr:{\mathbf Z}^{\boldsymbol\Delta}\longrightarrow \Pic(\overline S)\]
induces an embedding of the algebraic torus $T_{\NS}$ on
$T_{\boldsymbol\Delta}$.\footnote{There is some question of convention
in the definition of versal torsors which leads us to use
the opposite of the projection map.}
\par
The description of the kernel of the morphism $\pr$ (see~\eqref{equ.pic.fibre}
and~\eqref{equ.pic.can}) give the following
equations for $T_{\NS}$:
\begin{equation}\label{equ.torus.fibre}
Z_j^+Z_j^-=Z_k^+Z_k^-
\end{equation}
for $j$, $k\in{\{1,2,3,4\}}$ and
\begin{equation}\label{equ.torus.can}
Z_0^+Z_j^+Z_k^+=Z_0^-Z_l^-Z_m^-
\end{equation}
if $\{j,k,l,m\}={\{1,2,3,4\}}$. The equations~\eqref{equ.tors.closed}
are invariant under the action of $T_{\NS}$ thanks to~\eqref{equ.torus.fibre}
as are the inequalities~\eqref{equ.tors.open}. Therefore
the action of $T_{\NS}$ on ${\mathbf A}_{\boldsymbol\Delta}$ induces a natural action
of $T_{\NS}$ on ${\matheusm T}_{\boldsymbol{m}}$. This description of $T_{\NS}$ also
implies that $\pi_{\boldsymbol{m}}$ is invariant under the action
of $T_{\NS}$ on ${\matheusm T}_{\boldsymbol{m}}$. Indeed let ${\mathbf K}$ be an extension of ${\mathbf Q}$,
let $t$ belong to $T_{\NS}({\mathbf K})$ and ${\boldsymbol{z}}$ to ${\matheusm T}_{\boldsymbol{m}}({\mathbf K})$.
We put ${\boldsymbol{z}}'=t{\boldsymbol{z}}$. It follows from \eqref{equ.tors.uv}
and \eqref{equ.torus.fibre} that ${\boldsymbol{z}}$ and ${\boldsymbol{z}}'$ define
the same point $(u:v)\in{\mathbf P}^1({\mathbf K})$ and from \eqref{equ.tors.xyz},
\eqref{equ.torus.fibre} and \eqref{equ.torus.can} that ${\boldsymbol{z}}$
and ${\boldsymbol{z}}'$ give the same point $(x:y:tv^2)$ (resp.
$(x:y:tu^2)$ in ${\mathbf P}^2({\mathbf K})$).
\end{nota}
\begin{prop}
\label{prop.torsor}%
For any ${\boldsymbol{m}}\in\Sigma$,
the variety ${\matheusm T}_{\boldsymbol{m}}$ equipped with the map
$\pi_{\boldsymbol{m}}:{\matheusm T}_{\boldsymbol{m}}\to S$ and the above action of $T_{\NS}$
is a versal torsor above $S$.
\end{prop}
\begin{proof}
First of all, we may note that for any extension $K$ of
${\mathbf Q}$, if $R\in{\matheusm T}_{\boldsymbol{m}}(K)$ then $\pi_{\boldsymbol{m}}^{-1}(\pi_{\boldsymbol{m}}(R))$
coincides with the orbit of $R$ under the action of $T_{\NS}$.
Indeed if $R'\in{\matheusm T}_{\boldsymbol{m}}(K)$ satisfies $\pi_{\boldsymbol{m}}(R')=\pi_{\boldsymbol{m}}(R)$,
then there exists a unique ${\boldsymbol{z}}\in T_{\boldsymbol\Delta}({\mathbf K})$ such that $R'={\boldsymbol{z}} R$.
Let us write ${\boldsymbol{z}}=(z_\delta)_{\delta\in{\boldsymbol\Delta}}$.
Using~\eqref{equ.tors.uv} and~\eqref{equ.tors.xyz}
and the description of the action of ${\mathbf G}_m(K)$ on ${\mathcal T}_\textinmath{spl}$,
we get that $z_i^+z_i^-=z_j^+z_j^-$ if $1\leq i<j\leq 4$ and
\[z_0^+z_0^-(z_k^+z_k^-)^2=(z_0^+)^2\prod_{j=1}^4z_j^+=(z_0^-)^2
\prod_{j=1}^4z_j^-.\]
for $k\in{\{1,2,3,4\}}$. We deduce from these equations that ${\boldsymbol{z}}\inT_{\NS}(K)$.
\par
It is enough to prove the result over $\overline{\mathbf Q}$.
By choosing square roots $\alpha_j$ of $m_j$ such that
$\prod_{j=1}^4\alpha_j=\alpha_{\boldsymbol{m}}$, and using a change of
variable of the form ${Z_j^\varepsilon}'=\alpha_jZ_j^\varepsilon$
for $\varepsilon\in\{+1,-1\}$ and $j\in{\{1,2,3,4\}}$ we may
assume that ${\boldsymbol{m}}=(1,1,1,1)$.
Note that for any $\delta$ in ${\boldsymbol\Delta}$, the variety
$\pi_{\boldsymbol{m}}^{-1}(E_{\boldsymbol\Delta})$ is the subvariety of ${\matheusm T}_{\boldsymbol{m}}$
defined by $Z_\delta=0$. If $\varepsilon\in\{+1,-1\}$,
we consider the open subset
\[U_\varepsilon=S-E^{\varepsilon}-\bigcup_{j=1}^4E_j^\varepsilon\]
of $S$ and for $j\in{\{1,2,3,4\}}$, we put
\[U_j=S-E^+-E^--\bigcup_{k\neq j}(E_k^+\cup E_k^-).\]
The open subsets $U_1,U_2,U_3,U_4,U_+$ and $U_-$ form
an open covering of $S$. If $\varepsilon \in\{+1,-1\}$,
we may consider that $X+\varepsilon {\mathsl i} Y=1$ on $U_\varepsilon$
and we define a section $s^1_\varepsilon$ (resp. $s^2_\varepsilon$)
of $\pi_{\boldsymbol 1}$ over $U_\varepsilon\cap S_1$ (resp. $U_\varepsilon\cap S_2$)
by $Z_0^\varepsilon=Z_1^\varepsilon=Z_2^\varepsilon=Z_3^\varepsilon=Z_4^\varepsilon=1$,
$Z_0^{-\varepsilon}=t$ and $Z_j^{-\varepsilon}=L_j(U,1)$
(resp. $Z_j^{-\varepsilon}=L_j(1,V)$) for $j\in{\{1,2,3,4\}}$.
Similarly, for $j\in{\{1,2,3,4\}}$, fix $k,l,m$ so that $\{j,k,l,m\}=
{\{1,2,3,4\}}$. On $U_j$, we may consider that $L_k(U,V)=1$
and $T=1$. We may then define a section $s_j$ of $\pi_{\boldsymbol 1}$
over $U_j$ by
$Z_k^+=Z_k^-=Z_0^+=Z_0^-=Z_l^+=Z_m^+=1$
and
\[Z_l^-=L_l(U,V),\quad Z_m^-=L_m(U,V),
\quad Z_j^+=\frac{X+{\mathsl i} Y}{\prod_{r\neq j}Z_r^+}\quad
\text{and}\quad Z_j^-=\frac{X-{\mathsl i} Y}{\prod_{r\neq j}Z_r^+}.\]
The conditions~\eqref{equ.tors.open} ensures that, for any point
$P\in{\matheusm T}_{\boldsymbol 1}(\overline{\mathbf Q})$,
the stabilizer of $P$ in $T_{\NS}(\overline{\mathbf Q})$
is trivial. Using the action of $T_{\NS}$ on ${\matheusm T}_{\boldsymbol 1}$
we then get an equivariant isomorphism from $T_{\NS}\times U$ to
$\pi_{\boldsymbol 1}^{-1}(U)$ for each open subset $U$ described above.
This proves that ${\matheusm T}_{\boldsymbol{m}}$ is a $T_{\NS}$-torsor over~$S$.
\par
It remains to prove that the endomorphism of $\Pic(\overline S)$
defined by this torsor is the identity map. Let us first recall
how this endomorphism may be defined. If $L$ is a line bundle
over~$\overline S$, then the class of~$L$ defines a morphism
of Galois lattices ${\mathbf Z}\to\Pic(\overline S)$ and therefore
a morphism of algebraic tori $\phi_L:T_{\NS}\to{\mathbf G}_m$ and an action
of $T_{\NS}$ on ${\mathbf G}_m$. The restricted product ${\matheusm T}\times^{T_{\NS}}{\mathbf G}_m$
is a ${\mathbf G}_m$-torsor over $\overline S$ which defines an element
of $\Pic(\overline S)$.
For any $\delta$ in ${\boldsymbol\Delta}$, the function $Z_\delta$ on ${\matheusm T}_{\boldsymbol{m}}$
is invariant under the action of the kernel of the map
${\phi_\delta:T_{\NS}\to{\mathbf G}_m}$ defined by the class of $\delta$
in $\Pic(\overline S)$. Therefore this function defines
an antiequivariant map from ${\matheusm T}_{\boldsymbol{m}}\times^{T_{\NS}}{\mathbf G}_m$ to ${\mathbf A}^1$ which
vanishes with multiplicity one over $\pi_{\boldsymbol{m}}^{-1}(\delta)$.
Thus the endomorphism defined by ${\matheusm T}_{\boldsymbol{m}}$ on $\Pic(\overline S)$
sends the class of $\delta$ to itself for any $\delta\in{\boldsymbol\Delta}$.
This proves that ${\matheusm T}_{\boldsymbol{m}}$ is a versal torsor over $S$.
\end{proof}
To conclude these constructions it remains to prove
that the set of rational points $S({\mathbf Q})$ is the disjoint union
of the sets $\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf Q}))$ where ${\boldsymbol{m}}$ runs over the
set $\Sigma$.
\begin{lemma}
\label{lemma.lift.split}
For any $P\in S({\mathbf Q})$, we have
\[\sharp(\pi_\textinmath{spl}^{-1}(P)\cap{\mathcal T}_\textinmath{spl}({\mathbf Z}))=\sharp{\mathbf G}_m^2({\mathbf Q})_{\textinmath{tors}}=2^2.\]
\end{lemma}
\begin{proof}
Let us start with a point
$P=((x_0:y_0:t_0),u_0)$ in $S_1({\mathbf Q})$. We then have the relation
\[x_0^2+y_0^2=t_0^2\prod_{j=1}^4L_i(u_0,1)\]
We may write $u_0=u/v$ with $u,v\in{\mathbf Z}$ and $\gcd(u,v)=1$. Then
we may find an element $\lambda$ of ${\mathbf Q}$ such that the rational numbers
$x=\lambda x_0$, $y=\lambda y_0$ and $t=\lambda t_0/v^2$ are coprime
integers and we have
\[x^2+y^2=t^2\prod_{j=1}^4L_j(u,v).\]
The same construction works for any point of $S_2({\mathbf Q})$
and if $P$ belongs to $S_1({\mathbf Q})\cap S_2({\mathbf Q})$ the elements
of ${\mathbf Z}^5$ thus obtained coincide up to multiplication
of the first three or the last two coordinates by $-1$.
\end{proof}
\begin{rema}
Note that if we impose conditions like
\[t>0,\quad L_1(u,v)\geq 0\quad\text{and}\quad\prod_{j=2}^4L_j(u,v)\geq 0,\]
the lifting of $P$ is unique.
\end{rema}
\begin{prop}
\label{prop.uniquem}%
Let $P$ belong to $S({\mathbf Q})$.
Then there exists a unique ${\boldsymbol{m}}$ in $\Sigma$ such that
$P$ belongs to $\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf Q}))$.
\end{prop}
\begin{proof}
Let $Q=(x,y,t,u,v)\in{\mathcal T}_\textinmath{spl}({\mathbf Z})$ be such that $\pi_\textinmath{spl}(Q)=P$.
Without loss of generality we may assume that $Q=(x,y,t,u,v)\in{\mathbf Z}^5$
is such that
\begin{equation}\label{equ.lifting.first}
\begin{cases}
x^2+y^2=t^2\prod_{j=1}^4L_j(u,v),\\
\gcd(x,y,t)=1,\ \gcd(u,v)=1,\\
t>0,\ L_1(u,v)\geq 0,\ \text{and}\ \prod_{j=2}^4L_j(u,v)\geq 0.\\
\end{cases}
\end{equation}
The fact that $t^2\prod_{j=1}^4L_j(u,v)$
is the sum of two squares implies that
\begin{equation}\label{equ.lifting.pos}
\prod_{j=1}^4L_j(u,v)\geq 0
\end{equation}
and, if $\prod_{j=1}^4L_j(u,v)\neq 0$,
for any prime $p$ congruent to $3$ modulo $4$
\begin{equation}\label{equ.lifting.cong}
\sum_{j=1}^4v_p(L_j(u,v))\equiv 0\ \text{mod}\ 2.
\end{equation}
Let $j$ belong to ${\{1,2,3,4\}}$.
If $L_j(u,v)\neq0 $, we denote by $\epsilon_j\in\{-1,+1\}$
the sign of $L_j(u,v)$
and by $\Sigma_j(Q)$ the set of prime numbers $p$
which are congruent to $3$ modulo $4$ and such that
$v_p(L_j(u,v))$ is odd. We then put
\[m_j=\epsilon_j\times\prod_{p\in\Sigma_j(Q)}p.\]
If $L_j(u,v)=0$ we define $m_j$ as the only integer in $\Sigma_j$
such that $\prod_{k=1}^4m_k$ is a square. By construction,
we have $m_j\mid L_j(u,v)$
and the quotient $L_j(u,v)/m_j$ is the sum of two squares.
\par
Let us now check that ${\boldsymbol{m}}=(m_1,m_2,m_3,m_4)$ belongs to $\Sigma$.
According to~\eqref{equ.lifting.cong}, if a prime number belongs to
$\Sigma_j(Q)$ for some $j\in{\{1,2,3,4\}}$, then there exists
$k\in{\{1,2,3,4\}}$ with $k\neq j$
such that $p\in\Sigma_k(Q)$.
In particular, $p$ divides both $L_j(u,v)$ and $L_k(u,v)$
as well as
\[\Delta_{j,k}u=b_kL_j(u,v)-b_jL_k(u,v)\]
and $\Delta_{j,k}v$. Since $\gcd(u,v)=1$, we get that $p\mid\Delta_{j,k}$.
This proves that ${\boldsymbol{m}}\in\prod_{j=1}^4\Sigma_j$. But
combining~\eqref{equ.lifting.pos}, \eqref{equ.lifting.cong} and
the definition of ${\boldsymbol{m}}$ we get that $\prod_{j=1}^4m_j$ is a square.
If $d$ divides all the $m_j$, it divides $\gcd_{1\leq j<k\leq 4}(\Delta_{j,k})$
which is equal to $1$ since $\Delta_{1,2}=1$ under the
condition~\eqref{equ.pthreefour}. Finally $m_1>0$ since
$L_1(u,v)>0$ or $\prod_{j=2}^4L_j(u,v)>0$. Thus, ${\boldsymbol{m}}$ belongs to $\Sigma$.
\par
We now wish to prove that $Q$ belongs to $\hat\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf Q}))$.
By construction of ${\boldsymbol{m}}$, for any $j$ in ${\{1,2,3,4\}}$, the integer
$L_j(u,v)/m_j$ is the sum of two squares. Moreover if $p$
is a prime number, congruent to $3$ modulo $4$, then $p$
generates a prime ideal of ${\mathbf Z}[{\mathsl i}]$. From the
relations~\eqref{equ.lifting.first}, if $p\mid t$, then
$p\mid (x+{\mathsl i} y)(x-{\mathsl i} y)$. In that case we have $p\mid x$ and
$p\mid y$, which contradicts the fact that $\gcd(x,y,t)=1$.
As $t>0$, we get that $t$ may also be written as the sum
of two squares.
\par
If $\prod_{j=1}^4L_j(u,v)\neq 0$, we choose for $j\in\{1,2,3\}$
an element $z^+_j\in{\mathbf Z}[{\mathsl i}]$ such that $L_j(u,v)/m_j=z_j^+\overline{z_j^+}$
and an element $z_0^+\in{\mathbf Z}[{\mathsl i}]$ such that $t=z_0^+\overline{z_0^+}$.
Then we get the relation
\[L_4(u,v)/m_4=\left(\frac{x+{\mathsl i} y}{\alpha_{\boldsymbol{m}} (z_0^+)^2\prod_{j=1}^3z_j^+}
\right)\overline{\left(\frac{x+{\mathsl i} y}{\alpha_{\boldsymbol{m}} (z_0^+)^2\prod_{j=1}^3z_j^+}
\right)}\]
and we put $z_4^+=(x+{\mathsl i} y)/(\alpha_{\boldsymbol{m}} (z_0^+)^2\prod_{j=1}^3z_j^+)\in{\mathbf Q}[{\mathsl i}]$.
If $\prod_{j=1}^4L_j(u,v)=0$, we choose $z_1^+,z_2^+,z_3^+,z^+$
as above and $z_4^+\in{\mathbf Z}[{\mathsl i}]$ such that $L_4(u,v)/m_4=z_4^+\overline{z_4^+}$.
In both cases, we put $z_j^-=\overline{z_j^+}$ for $j\in{\{1,2,3,4\}}$
and $z_0^-=\overline{z_0^+}$.
\par
The family so constructed satisfy the
relations~\eqref{equ.tors.xyz} and~\eqref{equ.lifting.first},
from which it follows that the corresponding family
$(z_\delta)_{\delta\in{\boldsymbol\Delta}}$ is a solution to the
systems~\eqref{equ.tors.closed} and~\eqref{equ.tors.open}.
Thus we obtain a point $R$ in ${\matheusm T}_{\boldsymbol{m}}({\mathbf Q})$ such that
$\pi_{\boldsymbol{m}}(R)=P$.
\par
Let ${\boldsymbol{m}}'$ belong to $\Sigma$ and assume that the point $P$ belongs to
the set $\pi_{{\boldsymbol{m}}'}({\matheusm T}_{{\boldsymbol{m}}'}({\mathbf Q}))$ as well. Then
by~\eqref{equ.lifting.first}, we have for any prime number $p$
\[v_p(m'_j)-v_p(m'_k)=v_p(L_j(u,v))-v_p(L_k(u,v))=v_p(m_j)-v_p(m_k)\]
for any $j,k$ in ${\{1,2,3,4\}}$ such that $L_j(u,v)L_k(u,v)\neq 0$.
Similarly, denoting by $\sgn(m)$ the sign of an integer $m$,
we have
\[\sgn(m'_j)/\sgn(m'_k)=\sgn(m_j)/\sgn(m_k).\]
These relations between ${\boldsymbol{m}}$ and ${\boldsymbol{m}}'$ remain valid if
$L_j(u,v)L_k(u,v)=0$ since the products $\prod_{j=1}^4m_j$
and $\prod_{j=1}^4m_j'$ are squares.
But, by definition of $\Sigma$, we have
\[m'_1>0\quad\text{and}\quad\min_{1\leq j\leq 4}v_p(m'_j)=0\]
for any prime number $p$, and similarly for ${\boldsymbol{m}}$.
We obtain that ${\boldsymbol{m}}={\boldsymbol{m}}'$.
\end{proof}
\section{Jumping up}
\label{section.jumpingup}%
Having constructed the needed versal torsors explicitly,
we now wish to lift our initial counting problem
to these torsors. In order to do this, we shall define
an adelic domain $\mathcal D_{\boldsymbol{m}}$ in the adelic space
${\matheusm T}_{\boldsymbol{m}}({\boldsymbol A}_{\mathbf Q})$ so that for any $P\in\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf Q}))$
the cardinality of $\pi_{\boldsymbol{m}}^{-1}(P)\cap\mathcal D_{\boldsymbol{m}}$ is
$\cardT_{\NS}({\mathbf Q})_\textinmath{tors}$.
\subsection{Idelic preliminaries}
We first need to gather a few facts about the adelic
space $T_{\NS}({\boldsymbol A}_{\mathbf Q})$.
\begin{nota}
We consider the affine space
\[{\mathbf A}_{{\boldsymbol\Delta},{\mathbf Z}}=
\Spec({\mathbf Z}[X_{\delta},Y_\delta,\delta\in{\boldsymbol\Delta}_{\mathbf Q}]).\]
Let~$A$ be a commutative ring. The group ${\mathcal G}$ acts
on the ring
\[\prod_{\delta\in{\boldsymbol\Delta}}A\otimes_{{\mathbf Z}}{\mathbf Z}[{\mathsl i}]\]
and we may identify the $A$-points of ${\mathbf A}_{\boldsymbol\Delta}$ with
the elements of the invariant ring
\[A_{\boldsymbol\Delta}=\Bigl(\prod_{\delta\in{\boldsymbol\Delta}}A\otimes_{{\mathbf Z}}{\mathbf Z}[{\mathsl i}]\Bigr)^{\mathcal G}.\]
\par
Let ${\mathcal P}$ be the set of prime numbers.
Let $p\in{\mathcal P}$. We put ${{\matheusm S_p}=\Spec({\mathbf Q}_p\otimes_{\mathbf Z}\ZZ[{\mathsl i}])}$
which we may identify with the set of places of ${\mathbf Q}[{\mathsl i}]$ above
$p$. If $\boldsymbol a=(a_{\mathfrak p})_{\mathfrak p\in{\matheusm S_p}}$
and $\boldsymbol b=(b_{\mathfrak p})_{\mathfrak p\in{\matheusm S_p}}$
belong to ${\mathbf Z}^{\matheusm S_p}$, we write
$\boldsymbol a\geq \boldsymbol b$ if $a_{\mathfrak p}\geq b_{\mathfrak p}$
for $\mathfrak p\in{\matheusm S_p}$
and $\min(\boldsymbol a,\boldsymbol b)
=(\min(a_{\mathfrak p},b_{\mathfrak p}))_{\mathfrak p\in{\matheusm S_p}}$.
The valuations induce a map
\[\widehat v_p:{\mathbf Q}_p\otimes_{\mathbf Z}\ZZ[{\mathsl i}]
\longrightarrow({\mathbf Z}\cup\{+\infty\})^{{\matheusm S_p}}.\]
Thus we get a natural map
\[({\mathbf Q}_p\otimes_{\mathbf Z}\ZZ[{\mathsl i}])^{\boldsymbol\Delta}\longrightarrow
({\mathbf Z}\cup\{+\infty\})^{{\matheusm S_p}\times{\boldsymbol\Delta}}.\]
The action of ${\mathcal G}$ on ${\matheusm S_p}$ and ${\boldsymbol\Delta}$ induces
an action of ${\mathcal G}$ on the set on the right-hand side
so that the above map is ${\mathcal G}$ equivariant.
Denoting by $\overline\Gamma_p$ the set of invariants
in $({\mathbf Z}\cup\{+\infty\})^{{\matheusm S_p}\times{\boldsymbol\Delta}}$ and by $\Gamma_p$
its intersection with ${\mathbf Z}^{{\matheusm S_p}\times{\boldsymbol\Delta}}$, we get
a map
\[\log_p:{\mathbf A}_{\boldsymbol\Delta}({\mathbf Q}_p)\longrightarrow\overline\Gamma_p\]
whose restriction to $T_{\boldsymbol\Delta}({\mathbf Q}_p)$ is a morphism from
this group to the group $\Gamma_p$ and $\log_p$ is compatible with
the action of $T_{\boldsymbol\Delta}({\mathbf Q}_p)$ on the left and the action
of $\Gamma_p$ on the right.
We denote by $\Xi_p$ the set of elements $(r_{\mathfrak p,\delta})$
of $\Gamma_p$ such that
$r_{\mathfrak p,\delta}\geq 0$ for any $\mathfrak p\in{\matheusm S_p}$ and
any $\delta\in{\boldsymbol\Delta}$.
\par
If $T$ is an algebraic torus over ${\mathbf Q}$ which splits over ${\mathbf Q}({\mathsl i})$,
then $X^*(T)$ denotes the group of characters of $T$ over ${\mathbf Q}({\mathsl i})$
and $X_*(T)=\Hom(X^*(T),{\mathbf Z})$ its dual, that is the group
of cocharacters of $T$.
We denote by $\langle\cdot{,}\cdot\rangle$ the natural pairing $X^*(T)\times X_*(T)\to{\mathbf Z}$.
For any place $v$ of ${\mathbf Q}$, we denote
by $X_*(T)_v$ the group of cocharacters of $T$ over ${\mathbf Q}_v$,
which may be described as $X_*(T)^{\Gal(\overline{\mathbf Q}_v/{\mathbf Q}_v)}$.
We also consider the groups $X_*(T)_{\mathbf Q}=X_*(T)^{\mathcal G}$
and $X^*(T)_{\mathbf Q}=X^*(T)^{\mathcal G}$.
The group $\Gamma_p$ may then be seen as the group $X_*(T_{\boldsymbol\Delta})_p$.
The restriction of $\log_p$ from $T_{\boldsymbol\Delta}({\mathbf Q}_p)$ to $\Gamma_p$
is then the natural morphism defined in \cite[\S2.1]{ono:tori}.
For any $({\boldsymbol{r}}_\delta)_{\delta\in{\boldsymbol\Delta}}\in\Gamma_p$,
we put ${\boldsymbol{r}}_j^{\pm}=
{\boldsymbol{r}}_{D_j^{\pm}}$ for $j\in{\{1,2,3,4\}}$
and ${\boldsymbol{r}}_0^{\pm}={\boldsymbol{r}}_{E^\pm}$.
The group $X_*(T_{\NS})_p$ is then the subgroup of $\Gamma_p$
given by the equations
\begin{align*}
{\boldsymbol{r}}_j^++{\boldsymbol{r}}_j^-&={\boldsymbol{r}}_l^++{\boldsymbol{r}}_l^-\\
\noalign{\noindent for $1\leq j<l\leq 4$ and}
{\boldsymbol{r}}_0^++{\boldsymbol{r}}_j^++{\boldsymbol{r}}^+_l&={\boldsymbol{r}}_0^-+{\boldsymbol{r}}^-_m+{\boldsymbol{r}}^-_n
\end{align*}
if $\{j,l,m,n\}={\{1,2,3,4\}}$.
\end{nota}
\begin{rema}
If $p\equiv 3\mod 4$ or $p=2$ then there exists a unique
element $\mathfrak p$ in ${\matheusm S_p}$. Thus $\Gamma_p$ is canonically isomorphic
to ${\mathbf Z}^{{\boldsymbol\Delta}_{\mathbf Q}}$. If $p\equiv 1\mod 4$, then choosing
an element $\mathfrak p\in{\matheusm S_p}$, we get an isomorphism
from ${\mathbf Z}^{\boldsymbol\Delta}$ to $\Gamma_p$.
\end{rema}
\begin{lemma}
\label{lemma.idelic}%
For any prime $p$ the morphism $\log_p$ induces
an isomorphism from the quotient $T_{\NS}({\mathbf Q}_p)/T_{\NS}({\mathbf Z}_p)$
to $X_*(T_{\NS})_p$ and there is an exact sequence
\[1\longrightarrow T_{\NS}({\mathbf Q})_{\textinmath{tors}}\longrightarrowT_{\NS}({\mathbf Q})
\longrightarrow\bigoplus_{p\in{\mathcal P}}X_*(T_{\NS})_p\longrightarrow 0.\]
\end{lemma}
\begin{proof}
By \cite[p.~449]{draxl:tori}, the kernel of the map $\log_p$
from $T_{\NS}({\mathbf Q}_p)$ to $X_*(T_{\NS})_p$
coincides with $T_{\NS}({\mathbf Z}_p)$ for any prime $p$.
Let us prove that the map $\bigoplus_p\log_p$ from $T_{\NS}({\mathbf Q})$ to
$\bigoplus_pX_*(T_{\NS})_p$ is surjective.
We first assume that $p\neq 2$. If $p\equiv 1\mod 4$
we choose an element $\varpi\in{\mathbf Z}[i]$ such that $p=\varpi\overline\varpi$
and identify ${\matheusm S_p}$ with $\{\varpi,\overline\varpi\}$.
If ${\boldsymbol{r}}\in\Gamma_p$, we then define
\begin{equation*
\exp_\varpi({\boldsymbol{r}})=(\varpi^{r_{\varpi,\delta}}
\overline\varpi^{r_{\overline\varpi,\delta}})_{\delta\in{\boldsymbol\Delta}}.
\end{equation*}
If $p\equiv 3\mod 4$, then we put $\varpi=p$ and for ${\boldsymbol{r}}\in\Gamma_p$,
we define $\exp_\varpi({\boldsymbol{r}})$ to be $(\varpi^{r_{p,\delta}})_{\delta\in{\boldsymbol\Delta}}$.
By construction, $\exp_\varpi$ is a morphism from
$\Gamma_p$ to $T_{\boldsymbol\Delta}({\mathbf Q})$ and satisfies
$\log_p\circ\exp_\varpi=\Id_{\Gamma_p}$
and $\log_\ell\circ\exp_\varpi=0$ for
any prime $\ell\neq p$. Moreover we have
\begin{equation}
\label{equ.good.split}
\chi(\exp_\varpi({\boldsymbol{r}}))=p^{\langle\chi,{\boldsymbol{r}}\rangle}
\end{equation}
for any $\chi\in X^*(T_{\boldsymbol\Delta})_{\mathbf Q}$ and any ${\boldsymbol{r}}\in\Gamma_p$.
Therefore, if ${\boldsymbol{r}}$ belongs to $X_*(T_{\NS})_p$,
then $\exp_\varpi({\boldsymbol{r}})$ belongs to $T_{\NS}({\mathbf Q})$. It remains
to prove a similar result for $p=2$, although there is no
morphism which satisfies~\eqref{equ.good.split}.
Let ${\boldsymbol{r}}$ belong to $X_*(T_{\NS})_2$.
Let us write $r_j=r_j^+=r_j^-$
for~$j$ in ${\{0,\dots,4\}}$. Since ${\boldsymbol{r}}$ belong to $X_*(T_{\NS})_2$,
we have $r_1=r_2=r_3=r_4$.
We put $z^+_j=(1+{\mathsl i})^{r_j}$ for $j\in\{0,1,2,3\}$
and $z^+_4=(-{\mathsl i})^{r_0+2r_1}(1+{\mathsl i})^{r_0}$ and $z_j^-=\overline z_j^+$
for $j\in{\{0,\dots,4\}}$. Then $\log_2({\boldsymbol{z}})={\boldsymbol{r}}$ and ${\boldsymbol{z}}$ satisfies
equation~\eqref{equ.torus.fibre}. Moreover
if $\{j,k,l,m\}={\{1,2,3,4\}}$ one has
\[z_0^+z_j^+z_k^+/(z_0^-z_l^-z_m^-)=
\frac{(1+{\mathsl i})^{r_0+2r_1}}{(1-{\mathsl i})^{r_0+2r_1}}(-{\mathsl i})^{r_0+2r_1}=1\]
which proves that ${\boldsymbol{z}}$ satisfies~\eqref{equ.torus.can}.
\par
If ${\boldsymbol{z}}$ belongs to the kernel of the map $\bigoplus_p\log_p$
then its coordinates are invertible elements in ${\mathbf Z}[{\mathsl i}]$.
Thus ${\boldsymbol{z}}$ is a torsion element of $T_{\NS}({\mathbf Q})$.
\end{proof}
\subsection{Local domains}
To construct $\mathcal D_{\boldsymbol{m}}$, for any prime $p$
and any ${\boldsymbol{m}}\in\Sigma$
we shall define a fundamental domain in ${\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$
under the action of $T_{\NS}({\mathbf Q}_p)$ modulo $T_{\NS}({\mathbf Z}_p)$.
In other words, we want to construct an open domain
$\mathcal D_{{\boldsymbol{m}},p}\subset{\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$ such that
\begin{conditions}
\item The open set $\mathcal D_{{\boldsymbol{m}},p}$ is stable
under the action of $T_{\NS}({\mathbf Z}_p)$;
\item For any $t$ in $T_{\NS}({\mathbf Q}_p)\setminusT_{\NS}({\mathbf Z}_p)$,
one has $t.\mathcal D_{{\boldsymbol{m}},p}\cap\mathcal D_{{\boldsymbol{m}},p}=\emptyset$;
\item For any~$x$ in ${\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$, there exists
an element~$t$ in $T_{\NS}({\mathbf Q}_p)$ such that~$x$ belongs to
$t.\mathcal D_{{\boldsymbol{m}},p}$.
\end{conditions}
\begin{lemma}
\label{lemma.lift.inter}%
For any prime number $p$, the domain ${\mathcal T}_\textinmath{spl}({\mathbf Z}_p)$ is
a fundamental domain in ${\mathcal T}_\textinmath{spl}({\mathbf Q}_p)$ under the action
of $T_{\textinmath{spl}}({\mathbf Q}_p)$ modulo $T_{\textinmath{spl}}({\mathbf Z}_p)$.
\end{lemma}
\begin{proof}
As in the proof of lemma~\ref{lemma.lift.split}, if $P$ belongs
to $S({\mathbf Q}_p)$, there exists a point $Q=(x,y,t,u,v)\in{\mathcal T}_\textinmath{spl}({\mathbf Q}_p)$
such that $\pi_\textinmath{spl}(Q)=P$ and
\begin{equation*}
\min(v_p(x),v_p(y),v_p(t))=\min(v_p(u),v_p(v))=0.
\end{equation*}
The last condition is equivalent to $Q\in{\mathcal T}_\textinmath{spl}({\mathbf Z}_p)$. The lemma then
follows from the facts that the action of $T_{\textinmath{spl}}({\mathbf Q}_p)$
on ${\mathcal T}_\textinmath{spl}({\mathbf Q}_p)$ is given by
\[((\lambda,\mu),(x,y,t,u,v))\mapsto
(\lambda x,\lambda y,\mu^{-2}\lambda t,\mu u,\mu v)\]
and that the $T_{\textinmath{spl}}({\mathbf Q}_p)$-orbits are the fibers of the
projection ${\pi_{\textinmath{spl}}:{\mathcal T}_\textinmath{spl}({\mathbf Q}_p)\to S({\mathbf Q}_p)}$.
\end{proof}
\begin{nota}
Let ${\boldsymbol{n}}=(n_1,n_2,n_3,n_4)$ belong to $({\mathbf Z}\setminus\{0\})^4$. We then define
${\mathcal Y}_{\boldsymbol{n}}$ as the subscheme of ${\mathbf A}_{{\boldsymbol\Delta},{\mathbf Z}}$ given by the
equations
\begin{equation}
\label{equ.torcz}
\Delta_{j,k}n_l(X^2_l+Y^2_l)+\Delta_{k,l}n_j(X_j^2+Y_j^2)
+\Delta_{l,j}n_k(X_k^2+Y_k^2)=0
\end{equation}
if $1\leq j<k<l\leq 4$.
The scheme ${\mathcal T}_{\boldsymbol{n}}$ is the open subset of ${\mathcal Y}_{\boldsymbol{n}}$
given by the conditions~\eqref{equ.tors.open},
where we put $Z_{\delta^+}=X_\delta+{\mathsl i} Y_\delta$ and
$Z_{\delta^-}=X_\delta-{\mathsl i} Y_\delta$ for $\delta\in{\boldsymbol\Delta}_{\mathbf Q}$.
\end{nota}
\begin{rems}
(i) Let ${\boldsymbol{m}}$ be an element of $\Sigma$.
The scheme ${\mathcal T}_{\boldsymbol{m}}$ is a model of ${\matheusm T}_{\boldsymbol{m}}$
over $\Spec({\mathbf Z})$.
\par
(ii)
The variety ${\mathcal Y}_{{\boldsymbol{m}},{\mathbf Q}}$ corresponds to the
restricted product of the versal torsor
by the affine toric variety associated
to the opposite of the effective cone which
has been introduced in \cite[prop.~4.2.2]{peyre:torseurs}.
\par
(iii)
We may note that an element $Q\in{\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$
belongs to ${\mathcal Y}_{\boldsymbol{m}}({\mathbf Z}_p)$ if and only if $\log_p(Q)$
belongs to $\Xi_p$.
\par
(iv
The equations~\eqref{equ.torcz} define an intersection of two quadrics in
${\mathbf P}_{\mathbf Q}^7$,
upon which we will ultimately need to count integral points
of bounded height. As shown by Cook in \cite{cook:quadratic},
the Hardy--Littlewood circle method can be adapted to handle
intersections of diagonal quadrics in at least $9$ variables
provided that the associated singular locus is empty.
Here we will need to deal with an intersection
of diagonal quadrics in only $8$ variables.
For this we will call upon the alternative approach
based on the geometry of numbers in~\cite{bretechebrowning:4linear}.
\end{rems}
\begin{lemma}
\label{lemma.orbit}%
Two elements of ${\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$ belong to the same orbit
under the action of $T_{\NS}({\mathbf Z}_p)$ if and only if they have the
same image by $\pi_{\boldsymbol{m}}$ and $\log_p$.
\end{lemma}
\begin{proof}
According to proposition~\ref{prop.torsor},
two elements of ${\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$ belong to the same orbit
under the action of $T_{\NS}({\mathbf Q}_p)$ if and
only if their image by $\pi_{\boldsymbol{m}}$ coincide.
On the other hand, $T_{\NS}({\mathbf Z}_p)=T_{\NS}({\mathbf Q}_p)\cap T_{\boldsymbol\Delta}({\mathbf Z}_p)$
is the set of elements of ${\mathbf A}_{\boldsymbol\Delta}({\mathbf Q}_p)$ which
are sent to the origin of $\Gamma_p$ by $\log_p$.
Therefore if two elements of ${\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$
belong to the same orbit for $T_{\NS}({\mathbf Z}_p)$ their image
in $\overline\Gamma_p$ coincides. Conversely, let
$x$ and $y$ be elements of ${\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$ which have
the same image by $\pi_{\boldsymbol{m}}$ and $\log_p$. Then there
exists an element $t\inT_{\NS}({\mathbf Q}_p)$ such that $y=tx$.
Since $\log_p(x)=\log_p(y)$, if a coordinate $z_\delta$
of $x$ is different from $0$, the corresponding component
of $\log_p(t)$ is $0$. Taking into account the
conditions~\eqref{equ.tors.open} and the equations~\eqref{equ.torus.fibre}
and~\eqref{equ.torus.can} which define $T_{\NS}$, this implies
that $\log_p(t)$ is the unit element and thus $t\inT_{\NS}({\mathbf Z}_p)$.
\end{proof}
\begin{rema}
\label{rema.lifting}%
The idea behind the construction of $\mathcal D_{{\boldsymbol{m}},p}$
is first to consider
the intersection
\[\widehat\pi_m^{-1}({\mathcal T}_\textinmath{spl}({\mathbf Z}_p))\cap{\mathcal Y}_{\boldsymbol{m}}({\mathbf Z}_p),\]
which is stable under the action of $T_{\NS}({\mathbf Z}_p)$.
For all primes $p$ for which there is good reduction,
this intersection coincides with ${\mathcal T}_{\boldsymbol{m}}({\mathbf Z}_p)$.
More generally, if $p$ is good or if $p\not\equiv 1\mod 4$,
this intersection satisfies the conditions (i) to (iii)
and yields the wanted domain.
On the other hand, if $p$ is a prime dividing one of the
$\Delta_{j,k}$ and such that $p\equiv 1\mod 4$, then
for any $Q\in{\mathcal T}_\textinmath{spl}({\mathbf Z}_p)\cap\widehat\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p))$
the intersection
\[\widehat\pi_{\boldsymbol{m}}^{-1}(Q)\cap{\mathcal Y}_{\boldsymbol{m}}({\mathbf Z}_p)\]
is the union of a finite number of $T_{\NS}({\mathbf Z}_p)$-orbits.
We then select a total order on $\Gamma_p$ and choose
the minimal element in the image of the last intersection
by $\phi_p$. In that way, we construct the wanted domain.
\par
To better understand the construction,
let us first describe the conditions satisfied by $\log_p(R)$ for
a lifting $R$ of a point $Q\in{\matheusm T}_\textinmath{spl}({\mathbf Q}_p)$.
Let $R=(z_\delta)_{\delta\in{\boldsymbol\Delta}}\in{\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$ and let
$Q=(x,y,t,u,v)=\widehat\pi_{\boldsymbol{m}}(R)$.
Let us denote by $({\boldsymbol{r}}_\delta)_{\delta\in{\boldsymbol\Delta}}\in
\overline\Gamma_p$ the image of $R$ by $\log_p$.
We also put $\boldsymbol n_j=\widehat v_p(L_j(u,v)/m_j)$ for $j\in{\{1,2,3,4\}}$,
$\boldsymbol n_0=\widehat v_p(t)$
and $\boldsymbol n^{\pm}=\widehat v_p((x\pm{\mathsl i} y)/\alpha_{\boldsymbol{m}})$.
Then we have the relations
\begin{align}
\label{equ.rela.ni}%
\boldsymbol n_i&={\boldsymbol{r}}_i^++
{\boldsymbol{r}}_i^-\\
\noalign{\noindent for $j\in{\{0,\dots,4\}}$, and}
\label{equ.rela.no}%
\boldsymbol n^{\pm}&=2
{\boldsymbol{r}}_0^{\pm}+\sum_{j=1}^4{\boldsymbol{r}}_j^{\pm}.
\end{align}
\end{rema}
\begin{lemma}
\label{lemma.coeff.inter}
Let $p$ be a prime number and let ${\boldsymbol{m}}$ belong to $\Sigma$.
Let $Q$ belong to the intersection
${\mathcal T}_\textinmath{spl}({\mathbf Z}_p)\cap\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p))$
and let $(\boldsymbol n_j)_{j\in{\{0,\dots,4\}}}$ and
$\boldsymbol n^+,\boldsymbol n^-$
be the corresponding elements of ${\mathbf Z}^{\matheusm S_p}$
defined in remark~\ref{rema.lifting}.
\begin{assertions}
\item One has $\boldsymbol n_j\geq 0$ for $j\in{\{0,\dots,4\}}$,
$\boldsymbol n^+\geq 0$ and $\boldsymbol n^-\geq 0$.
\item If $p\not\in{\matheusm S}$, then $\min(\boldsymbol n_i,
\boldsymbol n_j)=0$
if $1\leq i<j\leq 4$.
\item If $p\not\equiv 1\mod 4$, then $\boldsymbol n_0=0$.
\item One has $\min(\boldsymbol n_0,\boldsymbol n^+,\boldsymbol n^-)=0$.
\item There exists a solution in $\Xi_p$ to the
equations~\eqref{equ.rela.ni}
and \eqref{equ.rela.no}.
\item The number of such solutions is finite.
\item There exists a unique solution to these equations
in $\Xi_p$ if $p\not\in{\matheusm S}$ or if $p\not\equiv 1\mod 4$.
\end{assertions}
\end{lemma}
\begin{proof}
We write ${\boldsymbol{m}}=(m_1,\dots,m_4)$ and $Q=(x,y,t,u,v)$.
As $Q$ belongs to the set $\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p))$,
one has that $p|m_i$ if and only if $p\equiv 3\mod 4$ and
$v_p(L_i(u,v))$ is odd.
If these conditions are verified, $v_p(\alpha_{\boldsymbol{m}})=1$
and $\alpha_{\boldsymbol{m}}|L_i(u,v)$. Similarly, using the equation~\eqref{equ.torsiz},
we have that $\alpha_{\boldsymbol{m}}|x\pm{\mathsl i} y$ and this concludes the proof
of a).
\par
We now assume that $p \not\in{\matheusm S}$. Let $i,j$ be such that
$1\leq i<j\leq 4$. Thus $p$ does not divide
$\Delta_{i,j}$. This implies that $\min(v_p(L_i(u,v)),v_p(L_j(u,v)))=0$
and so $\min(\boldsymbol n_i,\boldsymbol n_j)=0$.
\par
We now prove assertion c). If $p|t$ then by equation~\eqref{equ.torsiz},
it follows that $p^2|x^2+y^2$. If we assume that $p=2$ or $p\equiv 3\mod 4$
this implies that $p|x$ and $p|y$ which contradicts
the fact that $\min(v_p(x),v_p(y),v_p(t))=0$.
\par
Let $\mathfrak p\in{\matheusm S_p}$. If $\mathfrak p$ divides $x+{\mathsl i} y$, $x-{\mathsl i} y$
and $t$, then $p$ divides $x$, $y$ and $t$. This proves assertion d).
\par
Since $Q$ belongs to
$\pi_{\boldsymbol{m}}({\matheusm T}({\mathbf Q}_p))$,
the equations~\eqref{equ.rela.ni}
and \eqref{equ.rela.no} have a solution in $\Gamma_p$. If $p\equiv 3\mod 4$
or $p=2$, then the integers $r_j^{\pm}\in{\mathbf Z}$ are
such that $r_j^+=r_j^-$ for $j\in{\{0,\dots,4\}}$.
Therefore the equations~\eqref{equ.rela.ni}
have a unique solution in $\Gamma_p$. By a) the coordinates
of this solution are positive. If $p\equiv 1\mod 4$, then
by choosing an element $\mathfrak p\in {\matheusm S_p}$ we are reduced
to solving the equations
\begin{align*}
n_i&= r_i^++
r_i^-\\
\noalign{\noindent for $j\in{\{0,\dots,4\}}$, and}
n^{\pm}&=2 r_0^{\pm}+\sum_{j=1}^4 r_j^{\pm}.
\end{align*}
in ${\mathbf Z}^{\boldsymbol\Delta}$, where $n_j\geq 0$ for $j\in{\{0,\dots,4\}}$,
$n^+\geq 0$ and $n^-\geq 0$. Since we have the
relation $2n_0+\sum_{j=1}^4n_j=n^++n^-$, we may write
$n^+=2a_0^++\sum_{j=1}^4a_j^+$ where $0\leq a_j^+\leq n_j$
for $j\in{\{0,\dots,4\}}$. Then we put $a_j^-=n_j-a_j^+$
for $j\in{\{0,\dots,4\}}$ to get a solution with nonnegative coordinates.
\par
The assertion f) follows from the fact that there is only
a finite number of nonnegative
integral solutions to an equation of the form
$n=k^++k^-$.
\par
If $p\equiv 3\mod 4$ or $p=2$ we have already seen that
the solution to the system of equations is unique.
If $p\not\in{\matheusm S}$ and $p\equiv 1\mod 4$,
then it follows from the assertions
b) and d) that $r_j^{\pm}=\min(n_j,n^{\pm})$,
which implies that the solution is unique.
\end{proof}
\begin{lemma}
\label{lemma.most.places}%
If $p$ is a prime number such that $p\equiv 1\mod 4$
or $p\not\in{\matheusm S}$,
then for ${\boldsymbol{m}}\in\Sigma$, the set
${\mathcal Y}_{\boldsymbol{m}}({\mathbf Z}_p)\cap\widehat\pi_{\boldsymbol{m}}^{-1}(
{\mathcal T}_\textinmath{spl}({\mathbf Z}_p))$
satisfies the conditions (i) to (iii) and defines
a fundamental domain in ${\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$ under the action
of $T_{\NS}({\mathbf Z}_p)$.
\end{lemma}
\begin{proof}
To prove the lemma it is sufficient to prove that
the intersection of any nonempty
fiber of $\pi_{\boldsymbol{m}}$ with ${\mathcal T}_{\boldsymbol{m}}({\mathbf Z}_p)$ is not empty and is an orbit
under the action of $T_{\NS}({\mathbf Z}_p)$.
Let $P$ belong to the set $\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p))$.
By lemma~\ref{lemma.lift.inter} we may lift $P$ to a point
$Q$ which belongs to ${\mathcal T}_\textinmath{spl}({\mathbf Z}_p)$.
According to lemma~\ref{lemma.coeff.inter}, e), we may find
an element ${\boldsymbol{r}}\in\Xi_p$ which is a solution to the
equations~\eqref{equ.rela.ni}
and \eqref{equ.rela.no}. Let $R'$ be any lifting of $P$ to
${\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$ and let ${\boldsymbol{r}}'=\log_p(R)$. The difference
${\boldsymbol{r}}'-{\boldsymbol{r}}$ belongs to $X_*(T_{\NS})_p$. According to
lemma~\ref{lemma.idelic}, there exists $t\inT_{\NS}({\mathbf Q}_p)$ such
that $\log_p(t)={\boldsymbol{r}}-{\boldsymbol{r}}'$. Then the point $R=t.R'\in{\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$
satisfies ${\log_p(R)={\boldsymbol{r}}}$
and $R$ belongs to
${\mathcal Y}_{\boldsymbol{m}}({\mathbf Z}_p)\cap\widehat\pi_{\boldsymbol{m}}^{-1}(
{\mathcal T}_\textinmath{spl}({\mathbf Z}_p))$.
\par
It remains to prove that if two element $R$ and $R'$
of ${\mathcal T}_{\boldsymbol{m}}({\mathbf Z}_p)$ are
in the same fibre for $\pi_{\boldsymbol{m}}$ then they belong to the same
orbit under the action of $T_{\NS}({\mathbf Z}_p)$. Their images in
${\mathcal T}_\textinmath{spl}({\mathbf Q}_p)$ belong to ${\mathcal T}_\textinmath{spl}({\mathbf Z}_p)$ and therefore
are contained in the same orbit for the action of
$T_{\textinmath{spl}}({\mathbf Z}_p)$, which means that
the equations described in remark~\ref{rema.lifting}
for $\log_p(R)$ and $\log_p(R')$
are exactly the same. We then apply assertion g) of
lemma~\ref{lemma.coeff.inter} and
lemma~\ref{lemma.orbit}.
\end{proof}
\begin{lemma}
If the prime number $p$ does not belong to ${\matheusm S}$,
then for ${\boldsymbol{m}}\in\Sigma$, we have
\[{\mathcal T}_{\boldsymbol{m}}({\mathbf Z}_p)={\mathcal Y}_{\boldsymbol{m}}({\mathbf Z}_p)\cap\widehat\pi_{\boldsymbol{m}}^{-1}(
{\mathcal T}_\textinmath{spl}({\mathbf Z}_p)).\]
\end{lemma}
\begin{proof}
We keep the notation used in the proof of the previous lemma.
Using lemma~\ref{lemma.coeff.inter}, b)
and d), and the positivity of the coefficients in ${\boldsymbol{r}}$, we get that
$\min({\boldsymbol{r}}_{\delta_1},{\boldsymbol{r}}_{\delta_2})=0$ whenever $\delta_1\cap\delta_2=\emptyset$,
which means that $R$ belongs to ${\mathcal T}_{\boldsymbol{m}}({\mathbf Z}_p)$.
\end{proof}
\begin{defi}
Let ${\boldsymbol{m}}$ belong to $\Sigma$.
If $p\not\in{\matheusm S}$, we put $\mathcal D_{{\boldsymbol{m}},p}={\mathcal T}_{\boldsymbol{m}}({\mathbf Z}_p)$.
If $p\in{\matheusm S}$ and $p\not\equiv 1\mod 4$, we put
\[\mathcal D_{{\boldsymbol{m}},p}={\mathcal Y}_{\boldsymbol{m}}({\mathbf Z}_p)\cap\widehat\pi_{\boldsymbol{m}}^{-1}(
{\mathcal T}_\textinmath{spl}({\mathbf Z}_p)).\]
\end{defi}
It remains to define the domain for the primes $p\in{\matheusm S}$
such that $p\equiv 1\mod 4$.
\begin{nota}
\label{nota.badp.one}
We put ${\matheusm S}'=\{\,p\in{\matheusm S},\ p\equiv 1\mod 4\,\}$. For any
$p\in{\matheusm S}'$ we fix in the remainder of this text
a decomposition $p=\mathfrak \varpi_p\overline{\varpi_p}$
for an irreducible element $\varpi_p\in{\mathbf Z}[i]$.
We may then write ${\matheusm S_p}=\{\varpi_p,\overline{\varpi_p}\}$.
The group $\Gamma_p$ is isomorphic to ${\mathbf Z}^{\boldsymbol\Delta}$ through the
map $\phi_p$ which applies a family
$(r_{\mathfrak p,\delta})_{(\mathfrak p,\delta)\in{\matheusm S_p}\times{\boldsymbol\Delta}}$
onto the family $(r_{\varpi_p,\delta})_{\delta\in{\boldsymbol\Delta}}$.
Let $j\neq k$ be two elements of ${\{1,2,3,4\}}$ such that $p|\Delta_{j,k}$.
We then define
${\boldsymbol{f}}_{j,k}=(f_\delta)_{\delta\in\Delta}\in {\mathbf Z}^{{\boldsymbol\Delta}}$ by
\[f_\delta=\begin{cases}
1\text{ if $\delta\in\{D_j^-,D_k^+\}$},\\
0\text{ otherwise.}
\end{cases}\]
We put ${\boldsymbol{e}}_{j,k}=\phi_p^{-1}({\boldsymbol{f}}_{j,k})$ and consider the set
\begin{equation}
\label{equ.def.lambdap}
\Lambda_p=\Xi_p\setminus\bigcup_{\{(j,k)\in{\{1,2,3,4\}}\mid j<k
\text{ and }p\mid\Delta_{j,k}\}}
{\boldsymbol{e}}_{j,k}+\Xi_p.
\end{equation}
\end{nota}
\begin{defi}
Let ${\boldsymbol{m}}$ belong to $\Sigma$.
If $p\in{\matheusm S}$ and $p\equiv 1\mod 4$, then we define $\mathcal D_{{\boldsymbol{m}},p}$
to be the set of $R\in\widehat\pi_{\boldsymbol{m}}^{-1}({\mathcal T}_\textinmath{spl}({\mathbf Z}_p))$ such that
$\log_p(R)\in\Lambda_p$.
\end{defi}
\begin{rema}
In particular, one has $\mathcal D_{{\boldsymbol{m}},p}\subset
{\mathcal Y}_{\boldsymbol{m}}({\mathbf Z}_p)$ for any prime number $p$.
\end{rema}
\begin{lemma}
\label{lemma.bad.places}%
If $p\in{\matheusm S}$ and $p\equiv 1\mod 4$,
then for ${\boldsymbol{m}}\in\Sigma$, the set
$\mathcal D_{{\boldsymbol{m}},p}$
satisfies the conditions (i) to (iii) and defines
a fundamental domain in ${\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$ under the action
of $T_{\NS}({\mathbf Z}_p)$.
\end{lemma}
\begin{proof}
According to lemma~\ref{lemma.orbit} and lemma~\ref{lemma.coeff.inter} e),
we have only to prove that for any
$Q\in{\mathcal T}_\textinmath{spl}({\mathbf Z}_p)\cap\widehat\pi_{\boldsymbol{m}}({\matheusm T}_p)$, there
exist a unique solution of the equations~\eqref{equ.rela.ni}
and \eqref{equ.rela.no} which belongs to $\Lambda_p$.
Among the solutions in $\Xi_p$, there is a unique solution such that
if ${\boldsymbol{s}}=\phi_p({\boldsymbol{r}})$, the quadruple $(s_1^+,s_2^+,s_3^+,s_4^+)$
is maximal for the lexicographic order. It remains to prove that the solution
satisfies this last condition if and only if ${\boldsymbol{r}}$ belongs to $\Lambda_p$.
Let ${\boldsymbol{r}}$ be the solution for which the above quadruple is maximal
and $\widetilde{\boldsymbol{r}}$ be any solution in $\Xi_p$
and $\widetilde{\boldsymbol{s}}=\phi_p(\widetilde{\boldsymbol{r}})$.
If ${\boldsymbol{r}}\neq\widetilde{\boldsymbol{r}}$, then we consider the smallest $j\in{\{1,2,3,4\}}$
such that $s_j^+>\widetilde s_j^{\,+}$. With the notation
of remark~\ref{rema.lifting}, this implies that
${\boldsymbol{n}}_j\neq 0$, ${\boldsymbol{n}}^+\neq 0$ and ${\boldsymbol{n}}^-\neq 0$. Therefore
${\boldsymbol{n}}_0=0$ and there exists $k>j$ such that $s_k^+<\widetilde s_k^{\,+}$.
Since $s_j^-<\widetilde s_j^{\,-}$, we may conclude that
$\widetilde{\boldsymbol{r}}\in {\boldsymbol{e}}_{j,k}+\Xi_p$.
Moreover $p\mid\Delta_{j,k}$.
Conversely if $\widetilde{\boldsymbol{r}}$
belongs to ${\boldsymbol{e}}_{j,k}+\Xi_p$,
for some $j,k\in{\{1,2,3,4\}}$ such that $j<k$,
then $\widetilde{\boldsymbol{r}}-{\boldsymbol{e}}_{j,k}+{\boldsymbol{e}}_{k,j}$
is another solution to system of equations which gives
a bigger quadruple for the lexicographic order.
\end{proof}
\Subsection{Adelic domains and lifting of the points}
\begin{defi}
Let ${\boldsymbol{m}}\in\Sigma$.
We define the open subset $\mathcal{D}_{\boldsymbol{m}}$ of ${\matheusm T}_{\boldsymbol{m}}({\boldsymbol A}_{\mathbf Q})$
as the product ${\matheusm T}_{\boldsymbol{m}}({\mathbf R})\times\prod_{p\in{\mathcal P}}\mathcal D_{{\boldsymbol{m}},p}$.
\end{defi}
\begin{prop}
The set $\mathcal D_{\boldsymbol{m}}$ is a fundamental domain
in ${\matheusm T}_{\boldsymbol{m}}({\boldsymbol A}_{\mathbf Q})$ under the action of
$T_{\NS}({\mathbf Q})$ modulo $T_{\NS}({\mathbf Q})_\textinmath{tors}$. In other words
\begin{conditions}
\item The open set $\mathcal D_{\boldsymbol{m}}$ is stable
under the action of $T_{\NS}({\mathbf Q})_\textinmath{tors}$;
\item For any $t$ in $T_{\NS}({\mathbf Q})\setminusT_{\NS}({\mathbf Q})_\textinmath{tors}$,
one has $t.\mathcal D_{\boldsymbol{m}}\cap\mathcal D_{\boldsymbol{m}}=\emptyset$;
\item For any~$x$ in ${\matheusm T}_{\boldsymbol{m}}({\boldsymbol A}_{\mathbf Q})$, there exists
an element~$t$ in $T_{\NS}({\mathbf Q})$ such that~$x$ belongs to
$t.\mathcal D_{\boldsymbol{m}}$.
\end{conditions}
\end{prop}
\begin{proof}
The assertion (i) follows from the fact that $\mathcal D_{{\boldsymbol{m}},p}$
is stable under $T_{\NS}({\mathbf Z}_p)$ for any prime number $p$.
If $t$ belongs to $T_{\NS}({\mathbf Q})\setminusT_{\NS}({\mathbf Q})_\textinmath{tors}$,
then, by lemma~\ref{lemma.idelic}, there exists a prime number $p$
such that $\log_p(t)\neq 0$. Thus
$t.\mathcal D_{{\boldsymbol{m}},p}\cap\mathcal D_{{\boldsymbol{m}},p}=\emptyset$, which proves (ii).
Let~$x$ belong to ${\matheusm T}_{\boldsymbol{m}}({\boldsymbol A}_{\mathbf Q})$. For any prime number~$p$,
there exists an element $t_p\inT_{\NS}({\mathbf Q}_p)$ such that
$t_p.x\in\mathcal D_{{\boldsymbol{m}},p}$. By lemma~\ref{lemma.idelic},
there exists an element $t\inT_{\NS}({\mathbf Q})$ such that
$\log_p(t)=\log_p(t_p)$ for any prime number $p$
and $t.x\in\mathcal D_{\boldsymbol{m}}$.
\end{proof}
\begin{cor}
\label{cor.snow}%
Let $P$ belong to $S({\mathbf Q})$ and let ${\boldsymbol{m}}$ be the unique
element of $\Sigma$ such that $P\in\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf Q}))$.
Then
\[\sharp(\pi_{\boldsymbol{m}}^{-1}(P)\cap\mathcal D_{\boldsymbol{m}})=\cardT_{\NS}({\mathbf Q})_\textinmath{tors}=2^{8}.\]
\end{cor}
\begin{proof}
This corollary follows from the last proposition and the
fact that $\pi_{\boldsymbol{m}}^{-1}(x)$ is an orbit under the action of
$T_{\NS}({\mathbf Q})$.
\end{proof}
Let us now lift the heights to the versal torsors.
\begin{defi}
\label{defi.localheight}%
As in notation~\ref{nota.height} we put $C=\sqrt{\prod_{j=1}^4
|a_j|+|b_j|}$.
Let $w$ be a place of ${\mathbf Q}$. We define a function $H_w$
on ${\mathbf Q}_w^5$ by
\begin{equation*}
H_w(x,y,t,u,v)=
\begin{cases}
\max(\frac{|x|_w}C,\frac{|y|_w}C,\max(|u|_w,|v|_w)^2|t|_w)&
\text{if $w=\infty$,}\\
\max(|x|_w,|y|_w,\max(|u|_w,|v|_w)^2|t|_w)&\text{otherwise,}
\end{cases}
\end{equation*}
for any $(x,y,t,u,v)\in{\mathbf Q}_w^5$. If ${\boldsymbol{m}}\in\Sigma$,
we shall also denote
by $H_w:{\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_w)\to{\mathbf R}$ the composite function
$H_w\circ\widehat\pi_{\boldsymbol{m}}$. We then define $H:{\matheusm T}_{\boldsymbol{m}}({\boldsymbol A}_{\mathbf Q})\to{\mathbf R}$
by $H=\prod_{w\in\Val({\mathbf Q})}H_w$.
\end{defi}
\begin{rems}
(i) The line bundle $\omega_S^{-1}$ defines a character
$\chi_\omega$ on the torus $T_{\textinmath{spl}}={\mathbf G}_{m,\QQ}^2$ simply given by
${(\lambda,\mu)\mapsto\lambda}$ and we have
the relation
\begin{equation}
\label{equ.localheight}
H_w(t.R)=|\chi_\omega(t)|_wH_w(R)
\end{equation}
for any $t\inT_{\textinmath{spl}}({\mathbf Q}_w)$ and any $R\inT_{\textinmath{spl}}({\mathbf Q}_w)$.
A similar assertion is true on ${\matheusm T}_{\boldsymbol{m}}$ for ${\boldsymbol{m}}\in\Sigma$.
\par
(ii)
As a point $Q=(x:y:t:u:v)$ in ${\mathcal T}_\textinmath{spl}({\mathbf R})$
satisfies the equations~\eqref{equ.torsiz},
we have that
\[\max(\vert x\vert,\vert y\vert)^2\leq
\prod_{j=1}^4(|a_j|+|b_j|)
\max(\vert u\vert,\vert v\vert)^4\vert t\vert^2.\]
and it follows that
\[H_\infty(Q)=\max(\vert u\vert,\vert v\vert)^2\vert t\vert.\]
\end{rems}
\begin{prop}
\label{prop.liftingheight}%
Let ${\boldsymbol{m}}\in\Sigma$.
For any $R\in{\matheusm T}_{\boldsymbol{m}}({\mathbf Q})$, one has
\[H(\pi_{\boldsymbol{m}}(R))=H(R).\]
\end{prop}
\begin{proof}
We may define a map $\widehat\psi:{\mathbf Q}^5\to{\mathbf Q}^5$ by
$(x,y,t,u,v)\mapsto(v^2t:uvt:u^2t:x:y)$. The
restriction of the map~$\widehat\psi$ from
${\mathcal T}_\textinmath{spl}$ to ${\mathbf A}_{\mathbf Q}^5\setminus\{0\}$
is a lifting of the map $\psi:S\to S'$.
On $S'$ the height $H_4$ is given by
\[H_4(x_0:\cdots:x_4)=
\max\left(\vert x_0\vert_\infty,\vert x_1\vert_\infty,\vert x_2\vert_\infty,
\frac{\vert x_3\vert_\infty}C,\frac{\vert x_4\vert_\infty}C\right)
\times\prod_{p\in{\mathcal P}}\max_{0\leq j\leq 4}(|x_j|_p)\]
for any $(x_0,\dots,x_4)\in{\mathbf Q}^5$. This formula implies the statement
of the lemma.
\end{proof}
\begin{cor}
For any real number $B$, we have
\[N(B)=\frac 1{\cardT_{\NS}({\mathbf Q})_\textinmath{tors}}
\sum_{{\boldsymbol{m}}\in\Sigma}\sharp\{\,R\in{\matheusm T}_{\boldsymbol{m}}({\mathbf Q})\cap\mathcal D_{\boldsymbol{m}}
,\ H(R)\leq B\,\}\]
\end{cor}
\begin{proof}
This corollary follows from propositions~\ref{prop.uniquem},
\ref{prop.torsor},
and~\ref{prop.liftingheight} and corollary \ref{cor.snow}.
\end{proof}
\begin{rema}
For any prime number~$p$ and any ${\boldsymbol{m}}\in\Sigma$, we have
$\mathcal D_{{\boldsymbol{m}},p}\subset\widehat\pi_{\boldsymbol{m}}^{-1}({\mathcal T}_\textinmath{spl}({\mathbf Z}_p))$.
Therefore, for any $R=(R_w)_{w\in\Val({\mathbf Q})}$ belonging
to $\mathcal D_{\boldsymbol{m}}$, we have $H(R)=H_\infty(R_\infty)$.
\end{rema}
\begin{nota}
For any real number~$B$, and any ${\boldsymbol{m}}\in\Sigma$,
we denote by $\mathcal D_{{\boldsymbol{m}},\infty}(B)$ the set
of $R\in{\matheusm T}_{\boldsymbol{m}}({\mathbf R})$ such that the point
$Q=(x,y,t,u,v)=\widehat\pi_{\boldsymbol{m}}(R)$
satisfies the conditions
\begin{equation}
\label{equ.domain.infty}
H_\infty(Q)\leq B\quad\text{and}\quad H_\infty(Q)\geq\max(|u|,|v|)^2\geq 1.
\end{equation}
We define $\mathcal D_{\boldsymbol{m}}(B)$ as the product
$\mathcal D_{{\boldsymbol{m}},\infty}(B)\times\prod_{p\in{\mathcal P}}\mathcal D_{{\boldsymbol{m}},p}$.
\end{nota}
\begin{rema}
Let $F$ be a fiber of the morphism $\pi:S\to{\mathbf P}^1_{\mathbf Q}$.
Then the Picard group of $S$ is a free ${\mathbf Z}$-module
with a basis given by the pair $([F],[\omega_S^{-1}])$.
According to the formula~\eqref{equ.localheight},
the function $H_\infty$ corresponds to $[\omega_S^{-1}]$.
In a similar way the map applying $(x,y,t,u,v)$ to $\max(|u|,|v|)$
corresponds to $[F]$. On the other hand, the cone of effective
divisors in $\Pic(S)$ is the cone generated by $[F]$ and
$[E^+]+[E^-]=[\omega_S^{-1}]-2[F]$. But, by the preceding remark, the function
\[Q=(x,y,t,u,v)\longmapsto\frac{H_\infty(Q)}{\max(|u|,|v|)^2}\]
corresponds to $[E^+]+[E^-]$. Thus the lower bounds
imposed in the definition of $\mathcal D_{{\boldsymbol{m}},\infty}(B)$
corresponds to the condition (3.9)
of~\cite[p.~268]{peyre:cercle}.
\par
These lower bounds are automatically
satisfied by any point~$R$ in $\mathcal D_{\boldsymbol{m}}\cap{\matheusm T}_{\boldsymbol{m}}({\mathbf Q})$.
Indeed $Q=\widehat\pi_{\boldsymbol{m}}(R)$ belongs to ${\mathcal T}_\textinmath{spl}({\mathbf Z})$ and
writing $Q=(x,y,t,u,v)$ we get that $\max(|u|,|v|)\geq 1$.
Since $(x,y,t)\neq 0$, by equation~\eqref{equ.torsiz},
we also have that $t\neq 0$ and therefore $|t|\geq 1$ which yields
the second inequality.
\end{rema}
\begin{cor}
For any real number $B$, we have
\[N(B)=\frac 1{\cardT_{\NS}({\mathbf Q})_\textinmath{tors}}
\sum_{{\boldsymbol{m}}\in\Sigma}\sharp({\matheusm T}_{\boldsymbol{m}}({\mathbf Q})\cap\mathcal D_{\boldsymbol{m}}(B)).\]
\end{cor}
\begin{proof}
This follows from the last remark and the preceding corollary.
\end{proof}
\subsection{Moebius inversion formula and change of variables}
As is usual with these type of problems, we now wish
to use a Moebius inversion formula to replace the
primality conditions by divisibility conditions.
In fact we shall perform three inversions corresponding
to the various primality conditions.
\par
We shall simultaneously parametrize the sets thus introduced
to reduce our problem to the study of a series which
may be handled with techniques of analytic number theory.
\subsubsection{First inversion}
The first inversion corresponds to the conditions
imposed at the places $p\in{\matheusm S}$ with $p\equiv 1\mod 4$.
\begin{nota}
Let $\mathrm{N}(\mathfrak{a}) =\#({\mathbf Z}[{\mathsl i}]/\mathfrak{a})$ denote the norm of
an ideal $\mathfrak{a}$ of the ring of Gaussian integers ${\mathbf Z}[{\mathsl i}]$.
We define
\begin{equation*
\widehat{\mathfrak{D}}=\{\mathfrak b\subset {\mathbf Z}[{\mathsl i}],\
\mathrm{N}(\mathfrak b)\in \mathfrak{D}\},
\end{equation*}
where
\begin{equation}\label{18-D}
\mathfrak{D}=\{d\in\mathbf{Z}_{>0},\ p\mid d \Rightarrow p\equiv 1 \bmod{4} \}.
\end{equation}
\par
Let $A$ be a commutative ring.
Let ${\boldsymbol{\mathfrak b}}=(\mathfrak b_\delta)_{\delta\in{\boldsymbol\Delta}}$
be a family of ideals of $A\otimes_{\mathbf Z}\ZZ[{\mathsl i}]$ such that
$\mathfrak b_{\overline \delta}=\overline{\mathfrak b_\delta}$ for
any $\delta\in{\boldsymbol\Delta}$. Then $(\prod_{\delta\in{\boldsymbol\Delta}}\mathfrak b_\delta)^{\mathcal G}$
is an ideal of $A_{\boldsymbol\Delta}$ and for any ${\boldsymbol{n}}\in{\mathbf Z}^4$, we define
\[{\mathcal Y}_{\boldsymbol{n}}({\boldsymbol{\mathfrak b}})={\mathcal Y}_{\boldsymbol{n}}(A)\cap
\Bigl(\prod_{\delta\in{\boldsymbol\Delta}}\mathfrak b_\delta\Bigr)^{\mathcal G}.\]
We define $\mathcal I_{\boldsymbol\Delta}(A)$ as the set of such families of ideals.
For any $p$, the map $\log_p$ induces a map from
$\mathcal I_{\boldsymbol\Delta}({\mathbf Z})$ to $\overline\Gamma_p$.
If $\log_2({\boldsymbol{\mathfrak a}})=0$, then we define
\[{\boldsymbol{\lambda}}({\boldsymbol{\mathfrak a}})=\prod_{p\in{\mathcal P}\setminus\{2\}}\exp_{\varpi_p}(\log_p({\boldsymbol{\mathfrak a}})).\]
For any ${\boldsymbol{\mathfrak a}}\in\mathcal I_{\boldsymbol\Delta}({\mathbf Z})$, we also put
$\mathrm{N}({\boldsymbol{\mathfrak a}})=(\mathrm{N}(\mathfrak a^+_j))_{1\leq j\leq 4}\in\mathbf Z_{\geq 0}^4$.
\par
If ${\boldsymbol{\lambda}}=(\lambda_\delta)_{\delta\in{\boldsymbol\Delta}}$
belongs to $T_{\boldsymbol\Delta}({\mathbf Q})\cap {\mathbf Z}_{\boldsymbol\Delta}$,
then we put
$\mathrm{N}({\boldsymbol{\lambda}})=(\lambda^+_j\lambda^-_j)_{1\leq j\leq 4}\in\mathbf{Z}_{>0}^4$
and define a morphism
$m_{\boldsymbol{\lambda}}:{\mathcal Y}_{\mathrm{N}({\boldsymbol{\lambda}}){\boldsymbol{n}}}\to{\mathcal Y}_{\boldsymbol{n}}$ using the action
of the torus $T_{\boldsymbol\Delta}$ on ${\mathbf A}_{\boldsymbol\Delta}$.
For any commutative ring $A$, we may define an element ${\boldsymbol{\lambda}} A_{\boldsymbol\Delta}\in
\mathcal I_\Delta(A)$ by taking the family of ideals
$(\lambda_\delta A)_{\delta\in{\boldsymbol\Delta}}$. If ${\boldsymbol{\mathfrak a}}\in\mathcal I_{\boldsymbol\Delta}({\mathbf Z})$
satisfies $\log_2({\boldsymbol{\mathfrak a}})=0$, then ${\boldsymbol{\mathfrak a}}={\boldsymbol{\lambda}}({\boldsymbol{\mathfrak a}}){\mathbf Z}_{\boldsymbol\Delta}$.
For any
${\boldsymbol{\mathfrak a}}\in\mathcal I_{\boldsymbol\Delta}({\mathbf Z})$, we similarly define
${\boldsymbol{\mathfrak a}} A_{\boldsymbol\Delta}$ as
$(\mathfrak a_\delta A)_{\delta\in{\boldsymbol\Delta}}\in\mathcal I_{\boldsymbol\Delta}(A)$.
\par
Let ${\boldsymbol{m}}\in\Sigma$ and let ${\boldsymbol{\mathfrak a}}=(\mathfrak{a}_j)_{1\leq j\leq 4}\in
\widehat{\mathfrak D}^4$. We may see ${\boldsymbol{\mathfrak a}}$ as an element
of $\mathcal I_{\boldsymbol\Delta}({\mathbf Z})$ by putting $\mathfrak{a}_j^+=\mathfrak{a}_j$ and $\mathfrak{a}_j^-=\overline
\mathfrak{a}_j$ for $j\in{\{1,2,3,4\}}$ and $\mathfrak{a}_0^+=\mathfrak{a}_0^-={\mathbf Z}[{\mathsl i}]$.
Let ${\boldsymbol{n}}={\boldsymbol{m}}\mathrm{N}({\boldsymbol{\mathfrak a}})=(m_j\mathrm{N}(\mathfrak{a}_j))_{1\leq j\leq 4}$.
Recall that $\alpha_{\boldsymbol{m}}$ is the positive square root of
$\prod_{j=1}^4m_j$.
We put
\[\alpha_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}=\alpha_{\boldsymbol{m}}\times\prod_{j=1}^4\lambda({\boldsymbol{\mathfrak a}})_j^+.\]
Note that $\prod_{j=1}^4n_j=N(\alpha_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}})$.
We then define a map $\widehat\pi_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}:{\mathcal Y}_{\boldsymbol{n}}\to{\mathbf A}_{\mathbf Z}^5$
as follows: thanks to equations~\eqref{equ.torcz}
and the fact that, by~\eqref{equ.pthreefour},
the family $(a_j,b_j)_{1\leq j\leq 4}$ generates ${\mathbf Z}^2$,
the system of equations
\begin{equation}
L_j(U,V)=n_j(X_j^2+Y_j^2)
\end{equation}
in the variables $U$ and $V$ has a unique solution
in the ring of functions on ${\mathcal Y}_{\boldsymbol{n}}$.
We also define $T=X_0^2+Y_0^2$ and define $X$ and $Y$ by
the relation
\[X+{\mathsl i} Y=\alpha_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}(X_0+{\mathsl i} Y_0)^2\prod_{j=1}^4(X_j+{\mathsl i} Y_j).\]
The morphism $\widehat\pi_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}$ is then defined
by the family of functions $(X,Y,T,U,V)$.
Since these functions satisfy the relation
\[X^2+Y^2=T^2\prod_{j=1}^4L_j(U,V),\]
the image of $\widehat\pi_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}$ is contained in the
Zariski closure ${\mathcal Y}_\textinmath{spl}$ of ${\mathcal T}_\textinmath{spl}$ in ${\mathbf A}_{\mathbf Z}^5$.
\par
Let ${\boldsymbol{m}}\in\Sigma$ and ${\boldsymbol{\mathfrak a}}\in\widehat{\mathfrak{D}}^4$. For any prime number $p$
we define $\mathcal D_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},p}^1$ as
${\mathcal Y}_{\boldsymbol{n}}({\mathbf Z}_p)\cap\widehat\pi^{-1}_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}({\mathcal T}_\textinmath{spl}({\mathbf Z}_p))$
where ${\boldsymbol{n}}={\boldsymbol{m}}\mathrm{N}({\boldsymbol{\mathfrak a}})$.
For any real number $B$, we also define
$\mathcal D^1_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},\infty}(B)$ as the set of $R\in{\mathcal Y}_{\boldsymbol{n}}({\mathbf R})$
such that $\widehat\pi_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}(R)$ satisfies the
conditions~\eqref{equ.domain.infty}.
We then put $\mathcal D^1_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}(B)=\mathcal D^1_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},\infty}(B)
\times\prod_{p\in{\mathcal P}}\mathcal D^1_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},p}$.
When ${\boldsymbol{\mathfrak a}}_j={\mathbf Z}[{\mathsl i}]$ for $j\in{\{1,2,3,4\}}$, we shall forget ${\boldsymbol{\mathfrak a}}$ in
the notation.
\par
Let ${\matheusm S}'$ be the set of
$p\in{\matheusm S}$ such that $p\equiv 1\mod 4$.
For any $p\in{\matheusm S}'$,
we consider the set $\mathcal E_p$ of subsets $I$
of $\Delta\setminus\{E^+,E^-\}$ such that
\begin{conditions}
\item if $\delta_j^+\in I$ then there exists $k<j$ such that
$\delta_k^-\in I$;
\item if $\delta_k^-\in I$ then there exists $j>k$ such that
$\delta_j^+\in I$;
\item if $\delta_j^+\in I$ and $\delta_k^-\in I$ with $j\neq k$
then $p\mid\Delta_{j,k}$.
\end{conditions}
For any $I\in\mathcal E_p$ we define
${\boldsymbol{f}}_I=(f_\delta)_{\delta\in{\boldsymbol\Delta}}\in{\mathbf Z}^{\boldsymbol\Delta}$
by
\[f_\delta=\begin{cases}
1&\text{if $\delta\in I$,}\\
0&\text{otherwise.}
\end{cases}\]
Using notation~\ref{nota.badp.one}, we then consider
${\boldsymbol{e}}_I=\varphi_p^{-1}({\boldsymbol{f}}_I)$ and
$\Sigma'_p=\bigl\{\,\exp_{\varpi_p}({\boldsymbol{e}}_I),\ I\in
\mathcal E_p\,\bigr\}$.
We define $\Sigma'$ as the subset of $\mathcal I_{\boldsymbol\Delta}({\mathbf Z})$
defined by
\[\Sigma'=\biggl\{\,\biggl(\prod_{p\in{\matheusm S}'}{\boldsymbol{\lambda}}_p
\biggr){\mathbf Z}_{\boldsymbol\Delta},\ ({\boldsymbol{\lambda}}_p)_{p\in{\matheusm S}'}\in
\prod_{p\in{\matheusm S}'}\Sigma'_p\,\biggr\}\]
An element ${\boldsymbol{\mathfrak a}}\in\Sigma'$ is determined
by the quadruple $(\mathfrak{a}_j^+)_{1\leq j\leq 4}$ and we shall
also consider $\Sigma'$ as a subset of $\widehat{\mathfrak{D}}^4$.
For $p\in{\matheusm S}'$ we define a map $\mu_p:\mathcal E_p\to{\mathbf Z}$
by the conditions
\[\mu_p(\emptyset)=1\quad \text{and}\quad
\sum_{J\subset I}\mu_p(J)=0\text{ if $I\neq\emptyset$.}\]
The map $\mu:\Sigma'\to{\mathbf Z}$ is defined by
$\mu({\boldsymbol{\mathfrak a}})=\prod_{p\in{\matheusm S}'}\mu_p(I_p({\boldsymbol{\mathfrak a}}))$.
\par
We shall denote by ${\boldsymbol A}_{f,\infty}$ the ring
${\mathbf R}\times\prod_{p\in{\mathcal P}}{\mathbf Z}_p$.
\end{nota}
\begin{rems}
\label{rems.notas.ideals}
(i) Let ${\boldsymbol{\lambda}}=(\lambda_\delta)_{\delta\in{\boldsymbol\Delta}}\in T_{\boldsymbol\Delta}({\mathbf Q})\cap
{\mathbf Z}_{\boldsymbol\Delta}$.
Let $A$ be a commutative ring.
Then $m_{\boldsymbol{\lambda}}$ is a bijection from the set
${\mathcal Y}_{\mathrm{N}({\boldsymbol{\lambda}}){\boldsymbol{n}}}(A)$ to the set
${\mathcal Y}_{\boldsymbol{n}}({\boldsymbol{\lambda}} A_{\boldsymbol\Delta})$.
\par
(ii) With the same notation, for the ring $A={\mathbf Z}_p$, the set
${\mathcal Y}_{\boldsymbol{n}}(\boldsymbol{\mathfrak d})$ is the inverse
image by $\log_p$ of the set $\log_p({\boldsymbol{\lambda}})+\Xi_p$.
\end{rems}
\begin{lemma}
\label{lemma.firstinversion}%
Let $p\in{\matheusm S}'$. For any subset $K$ of $\Gamma_p$,
we denote by ${\boldsymbol 1}_K$ its characteristic function.
Then
\[{\boldsymbol 1}_{\Lambda_p}=\sum_{I\in\mathcal E_p}\mu_p(I){\boldsymbol 1}_{{\boldsymbol{e}}_I+\Xi_p}.\]
\end{lemma}
\begin{proof}
For any $j,k$ in ${\{1,2,3,4\}}$ such that $j<k$ and $p\mid\Delta_{j,k}$,
we put $I_{j,k}=\{\delta^-_j,\delta^+_k\}$.
Let $K$ be a subset of $\{\,(j,k)\in{\{1,2,3,4\}}^2,\
j<k\text{ and }p\mid\Delta_{j,k}\,\}$.
Let $I=\bigcup_{(j,k)\in K}I_{j,k}$. Then
we have
\[\bigcap_{(j,k)\in K}({\boldsymbol{e}}_{j,k}+\Xi_p)={\boldsymbol{e}}_I+\Xi_p.\]
On the other hand, a subset $I$ of ${\boldsymbol\Delta}$ belongs
to $\mathcal E_p$ if and only if it is the union of subsets
$I_{j,k}$ with $j<k$ and $p\mid\Delta_{j,k}$.
The lemma then follows from equation~\eqref{equ.def.lambdap}
which defines $\Lambda_p$
and the fact that the map $I\mapsto{\boldsymbol{e}}_I+\Xi_p$ reverses the
inclusions.
\end{proof}
\begin{lemma}
\label{lemma.multlambda}
Let ${\boldsymbol{\mathfrak a}}\in\Sigma'$ and let $B$ be a positive real number.
The multiplication by ${\boldsymbol{\lambda}}({\boldsymbol{\mathfrak a}})\in T_{\boldsymbol\Delta}({\mathbf Q})$
maps $\mathcal D^1_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}(B)$ onto
$\mathcal D^1_{{\boldsymbol{m}}}(B)\cap{\mathcal Y}_{\boldsymbol{m}}({\boldsymbol{\mathfrak a}}({\boldsymbol A}_{f,\infty})_{\boldsymbol\Delta})$.
\end{lemma}
\begin{proof}
By remark~\ref{rems.notas.ideals} (i),
the map $m_{{\boldsymbol{\lambda}}({\boldsymbol{\mathfrak a}})}$ is a bijection
from the set ${\mathcal Y}_{\mathrm{N}({\boldsymbol{\mathfrak a}}){\boldsymbol{m}}}({\boldsymbol A}_{f,\infty})$
onto the set ${\mathcal Y}_{\boldsymbol{m}}({\boldsymbol{\mathfrak a}}({\boldsymbol A}_{f,\infty})_{\boldsymbol\Delta})$.
Let us now compare the maps $\widehat\pi_{\boldsymbol{m}}\circ m_{{\boldsymbol{\lambda}}({\boldsymbol{\mathfrak a}})}$
and $\widehat\pi_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}$.
The map $\widehat\pi_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}$ is given by the relations
\[\begin{cases}
L_j(U,V)=\mathrm{N}(\mathfrak a^+_j)m_i(X_j^2+Y_j^2)\text{ for }j\in{\{1,2,3,4\}},\\
T=X_0^2+Y_0^2,\\
X+{\mathsl i} Y=\alpha_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}(X_0+{\mathsl i} Y_0)^2\prod_{j=1}^4(X_j+{\mathsl i} Y_j),
\end{cases}\]
whereas $\widehat\pi_{\boldsymbol{m}}\circ m_{{\boldsymbol{\lambda}}({\boldsymbol{\mathfrak a}})}$ is given by
\[\begin{cases}
L_j(U,V)=\lambda({\boldsymbol{\mathfrak a}})^+_j\lambda({\boldsymbol{\mathfrak a}})^-_j
m_i(X_j^2+Y_j^2)\text{ for }j\in{\{1,2,3,4\}},\\
T=X_0^2+Y_0^2,\\
X+{\mathsl i} Y=\alpha_{{\boldsymbol{m}}}\biggl(\prod_{j=1}^4\lambda({\boldsymbol{\mathfrak a}})_j^+\biggr)
(X_0+{\mathsl i} Y_0)^2\prod_{j=1}^4(X_j+{\mathsl i} Y_j).
\end{cases}\]
Therefore $\widehat\pi_{\boldsymbol{m}}\circ m_{{\boldsymbol{\lambda}}({\boldsymbol{\mathfrak a}})}$
coincides with $\widehat\pi_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}$.
This proves that for any prime number $p$, the
map $m_{\lambda({\boldsymbol{\mathfrak a}})}$ maps $\widehat\pi_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}^{-1}({\mathbf Z}_p)$
onto $\widehat\pi_{\boldsymbol{m}}^{-1}({\mathbf Z}_p)$.
Moreover $m_{{\boldsymbol{\lambda}}({\boldsymbol{\mathfrak a}})}$ sends the set $\mathcal D^1_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},\infty}(B)$
onto $\mathcal D^1_{{\boldsymbol{m}},\infty}(B)$.
\end{proof}
\begin{prop}
For any real number $B$, we have
\[N(B)=\frac 1{\cardT_{\NS}({\mathbf Q})_\textinmath{tors}}
\sum_{{\boldsymbol{m}}\in\Sigma}\sum_{{\boldsymbol{\mathfrak a}}\in\Sigma'}
\mu({\boldsymbol{\mathfrak a}})\sharp({\matheusm T}_{\mathrm{N}({\boldsymbol{\mathfrak a}}){\boldsymbol{m}}}({\mathbf Q})\cap
\mathcal D^1_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}(B)).\]
\end{prop}
\begin{proof}
This follows from lemma~\ref{lemma.firstinversion},
the definition of $\mathcal D_{\boldsymbol{m}}(B)$ and lemma~\ref{lemma.multlambda}.
\end{proof}
\subsubsection{Second inversion}
The inversion we shall now perform corresponds
to the condition $\gcd(x,y,t)=1$.
\begin{nota}
The map $\mu:\widehat{\mathfrak D}\to {\mathbf Z}$ is the multiplicative function
such that
\[\mu(\mathfrak p^k)=\begin{cases}
1&\text{if $k=0$,}\\
-1&\text{if $k=1$,}\\
0&\text{otherwise.}
\end{cases}\]
for any prime ideal $\mathfrak p$ in $\widehat{\mathfrak D}$ and
any integer $k\geq 0$.
\par
Let ${\boldsymbol{m}}\in\Sigma$ and ${\boldsymbol{\mathfrak a}}\in\Sigma'\subset\widehat{\mathfrak{D}}^4$. Let
${\boldsymbol{\mathfrak b}}=(\mathfrak{b}_j)_{j\in{\{1,2,3,4\}}}\in\widehat{\mathfrak{D}}^4$.
We put ${\boldsymbol{n}}=\mathrm{N}({\boldsymbol{\mathfrak a}}{\boldsymbol{\mathfrak b}}){\boldsymbol{m}}$ and
$\mu({\boldsymbol{\mathfrak b}})=\prod_{j=1}^4\mu(\mathfrak{b}_j)$.
Let $B$ be a real number. Let $p$ be a prime number.
If $R$ belongs to ${\mathcal Y}_{{\boldsymbol{n}}}({\mathbf Z}_p)$, we denote
by $X,Y,T,U$ and~$V$ the functions on ${\mathcal Y}_{{\boldsymbol{n}}}$
which define $\widehat\pi_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}{\boldsymbol{\mathfrak b}}}$.
The local domain $\mathcal D^2_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},p}$ is then defined
as follows:
\begin{itemize}
\item
If $p\equiv 3\mod 4$ or $p=2$, then $\mathcal D^2_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},p}$
is the set of $R\in{\mathcal Y}_{{\boldsymbol{n}}}({\mathbf Z}_p)$
such that $T(R)\in{\mathbf Z}_p^*$ and $\min(v_p(U(R)),v_p(V(R)))=0$;
\item
If $p\equiv 1\mod 4$ then $\mathcal D^2_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},p}$
is the set of $R=(z_\delta)_{\delta\in{\boldsymbol\Delta}}\in
{\mathcal Y}_{{\boldsymbol{n}}}({\mathbf Z}_p)$
such that $z_0^-$ belongs to $\bigcap_{j=1}^4\mathfrak{b}_j$, such that
$\min\bigl(v_p(T(R)),v_p\bigl(\prod_{j=1}^4\mathrm{N}({\boldsymbol{\mathfrak a}}_j)\bigr)\bigr)=0$
and such that
$\min(v_p(U(R)),v_p(V(R)))=0$.
\end{itemize}
We also put $\mathcal D^2_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\infty}(B)=\mathcal D^1_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},\infty}(B)$
and
\[\mathcal D^2_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}}}(B)=\mathcal D^2_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\infty}(B)
\times\prod_{p\in{\mathcal P}}\mathcal D^2_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},p}.\]
\end{nota}
\begin{prop}
\label{prop.secondinversion}%
For any real number $B$, we have the relation
\[N(B)=\frac 1{\cardT_{\NS}({\mathbf Q})_\textinmath{tors}}
\sum_{{\boldsymbol{m}}\in\Sigma}\sum_{{\boldsymbol{\mathfrak a}}\in\Sigma'}\sum_{{\boldsymbol{\mathfrak b}}\in\widehat
{\mathfrak D}^4}
\mu({\boldsymbol{\mathfrak a}})\mu({\boldsymbol{\mathfrak b}})
\sharp({\matheusm T}_{\mathrm{N}({\boldsymbol{\mathfrak a}})\mathrm{N}({\boldsymbol{\mathfrak b}}){\boldsymbol{m}}}({\mathbf Q})\cap
\mathcal D^2_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}}}(B)).\]
\end{prop}
\begin{proof}
Let ${\boldsymbol{m}}\in\Sigma$, let ${\boldsymbol{\mathfrak a}}\in\Sigma'$ and let $p$ be a prime
number.
\par
Let us first assume that $p\not\equiv 1\mod 4$.
By lemma~\ref{lemma.coeff.inter} c), we have $v_p(t)=0$
for any $(x,y,t,u,v)\in{\mathcal T}_\textinmath{spl}({\mathbf Z}_p)$. Conversely,
let $R$ belong to ${{\mathcal Y}_{{\boldsymbol{m}}\mathrm{N}({\boldsymbol{\mathfrak a}})}({\mathbf Z}_p)}$. If $v_p(T(R))=0$,
then ${\min(v_p(X(R)),v_p(Y(R)),v_p(T(R)))=0}$.
\par
We now assume that $p\equiv 1\mod 4$.
For any $R=(z_\delta)_{\delta\in{\boldsymbol\Delta}}\in{\mathcal Y}_{{\boldsymbol{m}}\mathrm{N}({\boldsymbol{\mathfrak a}})}({\mathbf Q}_p)$
we have the relations
\[T(R)=z_0^+z_0^-\quad\text{and}\quad
X(R)+{\mathsl i} Y(R)=\alpha_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}(z_0^+)^2\prod_{j=1}^4z_j^+.\]
Note that if $\varpi_p|\alpha_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}$ for any prime $p\equiv 1\mod 4$,
then $p|\alpha_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}}}$.
Therefore we have the relation $\gcd(X(R),Y(R),T(R))=1$ in ${\mathbf Z}_p$
if and only if $R$ satisfies the following two conditions:
\begin{conditions}
\item One has $\min(v_p(T(R)),v_p(\mathrm{N}(\prod_{j=1}^4\mathfrak{a}_j)))=0$;
\item There is no $j\in{\{1,2,3,4\}}$ and no $\varpi\in{\matheusm S_p}$ such that
$z_j^+\in\varpi$ and $z_0^+\in\overline\varpi$.
\end{conditions}
We denote by $\widehat{\boldsymbol{\mathfrak b}}$ the unique element
of $\mathcal I_{\boldsymbol\Delta}({\mathbf Z})$ such that
$\widehat\mathfrak{b}_j^+=\mathfrak{b}_j$ for $j\in{\{1,2,3,4\}}$
and $\widehat\mathfrak{b}_0^-=\bigcap_{j=1}^4\mathfrak{b}_j$.
A classical Moebius inversion yields that the characteristic
function of the set of the elements $R$ in ${\mathcal Y}_{{\boldsymbol{m}}\mathrm{N}({\boldsymbol{\mathfrak a}})}({\mathbf Z}_p)$
which satisfy condition (ii)
is equal to
\[\sum_{{\boldsymbol{\mathfrak b}}\in\widehat{\mathfrak D}^4}
\mu({\boldsymbol{\mathfrak b}}){\boldsymbol 1}_{{\mathcal Y}_{{\boldsymbol{m}}\mathrm{N}({\boldsymbol{\mathfrak a}})}\bigl(\widehat{\boldsymbol{\mathfrak b}}({\mathbf Z}_p)_{\boldsymbol\Delta}\bigr)}.
\]
By remark~\ref{rems.notas.ideals}~(i), the multiplication
map $m_{{\boldsymbol{\lambda}}({\boldsymbol{\mathfrak b}})}$ maps
${\mathcal Y}_{{\boldsymbol{m}}\mathrm{N}({\boldsymbol{\mathfrak a}})}\bigl(\widehat{\boldsymbol{\mathfrak b}}({\mathbf Z}_p)_{\boldsymbol\Delta}\bigr)$
onto the set of $(z_\delta)_{\delta\in{\boldsymbol\Delta}}$ in
${\mathcal Y}_{{\boldsymbol{m}}\mathrm{N}({\boldsymbol{\mathfrak a}}{\boldsymbol{\mathfrak b}})}({\mathbf Z}_p)$
such that $z_0^-$ belongs to $\bigcap_{j=1}^4\mathfrak{b}_j$.
The rest of the proof is similar to the proof of lemma~\ref{lemma.multlambda}.
\end{proof}
\subsubsection{Third inversion}
The last inversion corresponds to
the condition $\gcd(u,v)=1$, in which it will prove
nonetheless useful to retain the fact that $u,v$ cannot both be even.
\begin{nota}
Let ${\boldsymbol{m}}\in\Sigma$ and ${\boldsymbol{\mathfrak a}}\in\Sigma'$.
Let ${\boldsymbol{\mathfrak b}}=(\mathfrak{b}_j)_{j\in{\{1,2,3,4\}}}\in\widehat{\mathfrak{D}}^4$.
We put ${\boldsymbol{n}}=\mathrm{N}({\boldsymbol{\mathfrak a}})\mathrm{N}({\boldsymbol{\mathfrak b}}){\boldsymbol{m}}$.
Let~$\ell$ be an odd integer.
Let~$p$ be a prime number. The local domain $\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell,p}$
is then defined as follows:
\begin{itemize}
\item
If $p=2$, then $\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell,p}$
is the set of $R\in{\mathcal Y}_{{\boldsymbol{n}}}({\mathbf Z}_p)$
such that $T(R)\in{\mathbf Z}_p^*$ and $\min(v_p(U(R)),v_p(V(R)))=0$;
\item
If $p\equiv 3\mod 4$, then $\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell,p}$
is the set of $R\in{\mathcal Y}_{{\boldsymbol{n}}}({\mathbf Z}_p)$
such that $T(R)\in{\mathbf Z}_p^*$ and $\ell$ divides $U(R)$ and $V(R)$.
\item
If $p\equiv 1\mod 4$ then $\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell,p}$
is the set of $R=(z_\delta)_{\delta\in{\boldsymbol\Delta}}\in
{\mathcal Y}_{{\boldsymbol{n}}}({\mathbf Z}_p)$
such that $z_0^-$ belongs to $\bigcap_{j=1}^4\mathfrak{b}_j$, such that
$\min\bigl(v_p(T(R)),v_p\bigl(\prod_{j=1}^4\mathrm{N}({\boldsymbol{\mathfrak a}}_j)\bigr)\bigr)=0$
and $\ell$ divides $U(R)$ and $V(R)$.
\end{itemize}
We define $\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell,\infty}(B)
=\mathcal D^2_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\infty}(B)$
and
\[\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell}(B)
=\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell,\infty}(B)
\times\prod_{p\in{\mathcal P}}\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell,p}.\]
\end{nota}
\begin{prop}
\label{prop.aftermoebius}%
For any positive real number $B$,
we have that $N(B)$ is equal to
\[\frac 1{\cardT_{\NS}({\mathbf Q})_\textinmath{tors}}
\sum_{{\boldsymbol{m}}\in\Sigma}\sum_{{\boldsymbol{\mathfrak a}}\in\Sigma'}\sum_{{\boldsymbol{\mathfrak b}}\in\widehat
{\mathfrak D}^4}
\sum_{\substack{\ell=1\\ 2\nmid \ell}}^\infty
\mu({\boldsymbol{\mathfrak a}})\mu({\boldsymbol{\mathfrak b}})\mu(\ell)
\sharp({\matheusm T}_{\mathrm{N}({\boldsymbol{\mathfrak a}})\mathrm{N}({\boldsymbol{\mathfrak b}}){\boldsymbol{m}}}({\mathbf Q})\cap
\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell}(B)).\]
\end{prop}
\section{Formulation of the counting problem}
We are now ready to begin the analytic part of the proof
of theorem \ref{theo.main}. Let us recall that
the linear forms that we are working with take the shape
\[\begin{array}{ll}
L_1(U,V)=U, &
L_2(U,V)=V, \\
L_3(U,V)=a_3U+b_3V, &
L_4(U,V)=a_4U+b_4V,
\end{array}\]
with integers $a_3,b_3,a_4,b_4$ such that
$\gcd(a_3,b_3)=\gcd(a_4,b_4)=1$ and
\begin{equation}
\label{eq:resultant}
\Delta=a_3b_3a_4b_4(a_3b_4-a_4b_3)\neq 0.
\end{equation}
It is clear that the forms involved are all pairwise non-proportional.
In this section we will further reduce our counting problem
using the familiar multiplicative
arithmetic function
\[r(n)=\sharp\{(x,y)\in {\mathbf Z}^2,\ x^2+y^2=n\}=
4 \sum_{d\mid n }\chi(d),\]
where $\chi$ is the real non-principal character modulo $4$. It is to
this expression that we will be able to direct the full force of analytic
number theory.
In what follows we will allow the implied constant in any estimate to
depend arbitrarily upon the coefficients of the linear forms
involved. Furthermore, we will henceforth
reserve $j$ for an arbitrary index from the set $\{1,2,3,4\}$.
Finally, many of our estimates will involve a small parameter
$\varepsilon>0$ and it will ease notation if we also allow the implied constants to
depend on the choice of $\varepsilon$. We will follow common practice and allow $\varepsilon$ to take
different values at different parts of the argument.
Recall the definitions of $\Sigma, \Sigma'$ from section
\ref{section.torsors} and section
\ref{section.jumpingup} respectively.
In particular we have $m_jN(\mathfrak{a}^+_j)=O(1)$ whenever ${\boldsymbol{m}}\in
\Sigma$ and ${\boldsymbol{\mathfrak a}}\in\Sigma'$.
\begin{prop}\label{prop:6.1}
For $B\geq 1$, we have
\begin{align*}
N(B)&=\frac 1{\cardT_{\NS}({\mathbf Q})_\textinmath{tors}}
\sum_{\substack{{\boldsymbol{m}}\in \Sigma\\ {\boldsymbol{\mathfrak a}}\in \Sigma'}}\mu({\boldsymbol{\mathfrak a}})
\sum_{\substack{\ell=1\\ 2\nmid \ell}}^\infty \mu(\ell)
\sum_{{\boldsymbol{\mathfrak b}} \in \widehat{{\mathfrak{D}}}^4} \mu({\boldsymbol{\mathfrak b}})
\sum_{\substack{t \in {\mathfrak{D}}\\ \gcd(t,\mathrm{N}({\boldsymbol{\mathfrak a}}))=1\\ \mathrm{N}(\bigcap
\mathfrak{b}_j)\mid t}}
r\Big(\frac{t}{\mathrm{N}(\bigcap \mathfrak{b}_j)}\Big)
\mathcal{U}\Big(\frac{B}{t}\Big),
\end{align*}
where
\begin{align*}
\mathcal{U}(T)=
\sum_{\substack{(u,v)\in {\mathbf Z}^2\cap \sqrt{T}\mathcal{R_{\boldsymbol{m}}}\\
\ell\mid u,v\\
2\nmid \gcd(u,v)\\
m_j\mathrm{N}(\mathfrak{a}^+_j\mathfrak{b}_j) \mid L_j(u,v)
}}\prod_{j=1}^4r\Big(\frac{L_j(u,v)}{m_j\mathrm{N}(\mathfrak{a}^+_j\mathfrak{b}_j) } \Big)
\end{align*}
and
\begin{equation}
\label{eq:RR}
\mathcal{R_{\boldsymbol{m}}}=\Big\{(u,v)\in {\mathbf R}^2,\
0<|u|,|v|\leq 1, ~m_jL_j(u,v)>0\text{ for }j\in{\{1,2,3,4\}}\Big\}.
\end{equation}
\end{prop}
\begin{proof}
We apply proposition~\ref{prop.aftermoebius}.
Let ${\boldsymbol{m}}\in\Sigma$, ${\boldsymbol{\mathfrak a}}\in\Sigma'$ and ${\boldsymbol{\mathfrak b}}\in\widehat{\mathcal D}^4$.
We wish to express $\sharp({\matheusm T}_{\mathrm{N}({\boldsymbol{\mathfrak a}})\mathrm{N}({\boldsymbol{\mathfrak b}}){\boldsymbol{m}}}({\mathbf Q})\cap
\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell}(B))$ in terms of the function $r$.
But given $(t,u,v)\in{\mathbf Z}^3$, the number of elements $R$ in that
intersection such that $(T(R),U(R),V(R))=(t,u,v)$
is $0$ if $(t,u,v)$ does not satisfy the conditions
\[\gcd(t,\mathrm{N}({\boldsymbol{\mathfrak a}}))=1,\quad
N(\bigcap\mathfrak{b}_j)|t,\quad
\ell|u,v,\quad 2\nmid t\gcd(u,v
\text{ and }m_j\mathrm{N}(\mathfrak{a}_j^+\mathfrak{b}_j) \mid L_j(u,v)\]
and is equal to
\[r\left(\frac t{\mathrm{N}(\bigcap\mathfrak{b}_j)}\right)
\prod_{j=1}^4r\left(\frac{L_j(u,v)}{m_j\mathrm{N}(\mathfrak{a}_j^+\mathfrak{b}_j)}\right)\]
otherwise.
\end{proof}
Let us set
\begin{equation}
\label{eq:thed}
d_j=m_j\mathrm{N}(\mathfrak{a}^+_j)\mathrm{N}(\mathfrak{b}_j), \quad D_j=\begin{cases}
[d_j,\ell], & \mbox{if $j=1$ or $2$},\\
d_j, & \mbox{if $j=3$ or $4$},
\end{cases}
\end{equation}
where $[d_j,\ell]$ is the least common multiple of $d_j,\ell$.
Then $d_j,D_j$ are odd positive integers such that $d_j\mid D_j$.
We may then write
\begin{equation}\label{eq:UUT}
\mathcal{U}(T)=
\sum_{\substack{(u,v)\in \mathsf{\Gamma}_{\mathbf{D}}\cap \sqrt{T}\mathcal{R_{\boldsymbol{m}}}\\
2\nmid \gcd(u,v)\\
}}\prod_{j=1}^4r\Big(\frac{L_j(u,v)}{d_j} \Big),
\end{equation}
where
\begin{equation}\label{eq:lattice}
\mathsf{\Gamma}_{\mathbf{D}}=\{(u,v)\in{\mathbf Z}^2,\ D_j\mid L_j(u,v)\}.
\end{equation}
Before passing to a detailed analysis of the sum $\mathcal{U}(T)$ and its
effect on the behaviour of the counting function $N(B)$, we will first
corral together some of the technical tools that
will prove useful to us.
\subsection{Geometric series}
\label{geom-series}
Given a vector $\mathbf{n}=(n_1,n_2,n_3,n_4)\in {\mathbf Z}_{\geq 0}^4$, let
$$
m(\mathbf{n})=\max_{i\neq j}\{n_i+n_j\}.
$$
It will be useful to note that
$m(n_1+\lambda,\ldots,n_4+\lambda)=m(\mathbf{n})+2\lambda$, for any $\lambda\in{\mathbf Z}$,
whence in particular $m(\mathbf{n})-2=m(n_1-1,n_2-1,n_3-1,n_4-1)$.
For
$\varepsilon\in\{-1,+1\}$ we will need to calculate the geometric series
\begin{equation}\label{30-S0pm}
S_0^{\varepsilon}(z)=\sum_{\mathbf{n}\in{\mathbf Z}_{\geq 0}^4} \varepsilon^{n_1+n_2+n_3+n_4}
z^{m(\mathbf{n})},
\end{equation}
for $|z|<1$. To do so we will break up the sum according to the values of
$\min\{n_1,n_2\}$ and $\min\{n_3,n_4\}$.
Let $S_{0,0}^\varepsilon(z)$ denote the contribution to $S_0^\varepsilon(z)$ from
$\mathbf{n}$ such that $\min\{n_1,n_2\}=\min\{n_3,n_4\}=0$, and
let $S_{0,1}^\varepsilon(z)$ denote the corresponding contribution from
$\mathbf{n}$ such that $\min\{n_1,n_2\}\geq 1$ and $\min\{n_3,n_4\}=0$.
Now it is rather easy to see that
\begin{equation}\label{31-0,0}
\begin{split}
S_{0,0}^\varepsilon(z)
&=
\Big(\sum_{\min\{n_1,n_2\}=0}(\varepsilon z)^{n_1+n_2}\Big)^2 =
\Big(\frac{1+\varepsilon z}{1-\varepsilon z}\Big)^2.
\end{split}
\end{equation}
since $m(\mathbf{n})=n_1+n_2+n_3+n_4$ in this setting. Next we claim that
\begin{equation}\label{31-1,0}
S_{0,1}^\varepsilon(z)= \frac{ (1+2\varepsilon +2z+\varepsilon z^2)z^2}{(1-\varepsilon
z)^2(1-\varepsilon z^2)}.
\end{equation}
To see this we note that
\begin{align*}
S_{0,1}^\varepsilon(z)&=\Big(2\sum_{n_1,n_2,n_3\geq 1, n_4=0}+ \sum_{n_1,n_2\geq 1,
n_3=n_4=0} \Big) \varepsilon^{n_1+n_2+n_3+n_4}z^{m(\mathbf{n})}.
\end{align*}
Now the second summation is clearly $\big(\sum_{a\geq
1}(\varepsilon z)^a\big)^2=z^2/(1-\varepsilon z)^2$. Similarly, the first summation is
\begin{align*}
&= 2
\sum_{n_1,n_2,n_3\geq 1} (\varepsilon z)^{n_1+n_2+n_3}z^{-\min\{n_j\}}\\
&=2\sum_{k\geq 1}z^{-k} \sum_{\min\{n_j\}=k} (\varepsilon z)^{n_1+n_2+n_3}\\
&=2\sum_{k\geq 1}z^{-k} \Big(\sum_{n_1,n_2,n_3\geq k}
(\varepsilon z)^{n_1+n_2+n_3}- \sum_{n_1,n_2,n_3,\geq k+1}
(\varepsilon z)^{n_1+n_2+n_3}\Big)\\
&=2\sum_{k\geq 1}z^{-k}
\Big(\frac{(\varepsilon z)^{3k}}{(1-\varepsilon z)^3}-\frac{(\varepsilon
z)^{3k+3}}{(1-\varepsilon z)^3}
\Big)
= 2\varepsilon\frac{(1+\varepsilon z+z^2)z^2}{(1-\varepsilon
z)^2(1-\varepsilon z^2)}.
\end{align*}
Combining these two equalities completes the proof of \eqref{31-1,0}.
We may now establish the following result.
\begin{lemma}\label{31-s0-}
Let $|z|<1$. Then we have
$$
S_0^-(z)=\frac{(1-z)^2}{(1+z)^2(1+z^2)}
$$
and
$$
S_0^+(z)=\frac{1+2z+6z^2+2z^3+z^4}{(1-z)^4(1+z)^2}.
$$
\end{lemma}
\begin{proof}
The proof of lemma \ref{31-s0-} is based on the simple observation that
$$
S_0^\varepsilon(z)=S_{0,0}^\varepsilon(z)+2 S_{0,1}^\varepsilon(z) + z^2
S_0^\varepsilon(z),
$$
from which it follows that
$$
S_0^\varepsilon(z)=(1-z^{2})^{-1}\big(S_{0,0}^\varepsilon(z)+2
S_{0,1}^\varepsilon(z)\big).
$$
We complete the proof of the lemma by inserting
\eqref{31-0,0} and \eqref{31-1,0} into this equality.
\end{proof}
\subsection{Geometry of numbers}
It will be useful to collect together some elementary facts concerning the
set $\mathsf{\Gamma}_{\mathbf{D}}$ that was defined in \eqref{eq:lattice}.
For the moment we allow $\mathbf{D}\in \mathbf{Z}_{>0}^4$ to be arbitrary.
It is clear that $\mathsf{\Gamma}_{\mathbf{D}}$
defines a sublattice of ${\mathbf Z}^2$ of rank $2$, since it is closed
under addition and contains the vector $D_1D_2D_3D_4(u,v)$ for any $(u,v)\in {\mathbf Z}^2$.
Let us write
\begin{equation}
\label{eq:def-rho}
\rho(\mathbf{D})=\det \mathsf{\Gamma}_{\mathbf{D}},
\end{equation}
for the determinant.
It follows from the Chinese remainder theorem that there is a multiplicativity property
$$
\rho(g_1 h_1,\ldots,g_4 h_4)=\rho(g_1,\ldots,g_4)\rho(h_1,\ldots,h_4),
$$
whenever $\gcd(g_1 g_2 g_3 g_4,h_1 h_2 h_3 h_4)=1$.
Recall the definition \eqref{eq:resultant} of $\Delta$. Then
\cite[Eqn. (3.12)]{heathbrown:2003} shows that
\begin{equation}
\label{eq:tenn''}
\rho(p^{e_1},\ldots,p^{e_4})=p^{\max_{i< j}\{e_{i}+e_{j}\}},
\end{equation}
for any prime $p\nmid \Delta$. Likewise,
when $p\mid \Delta$ one has
\begin{equation}
\label{eq:tenn-}
\rho(p^{e_1},\ldots,p^{e_4})\asymp p^{\max_{i< j}\{e_{i}+e_{j}\}},
\end{equation}
where the symbol $\asymp$ indicates that the two
quantities involved have the same order of magnitude.
It follows from the properties that we have recorded here that
\begin{equation}
\label{eq:251.1}
\rho(\mathbf{D})\asymp [D_1D_2,D_1D_3,D_1D_4,D_2D_3,D_2D_4,D_3D_4].
\end{equation}
We can also say something about the size of the smallest successive
minimum, $s_1$ say, of
$\mathsf{\Gamma}_{\mathbf{D}}$. Thus we have
\begin{equation}\label{sm}
s_1 \geq \min\{D_1,D_2\}.
\end{equation}
For this we note that
$\mathsf{\Gamma}_{\mathbf{D}}\subseteq \mathsf{\Lambda}=\{(u,v) \in {\mathbf Z}^2,\ ~D_1 \mid u, ~D_2 \mid v\}$.
Now $\mathsf{\Lambda}\subseteq {\mathbf Z}^2$ is a sublattice of rank $2$, with smallest
successive minimum $\min\{D_1,D_2\}$.
The desired inequality is now obvious.
\section{Estimating \texorpdfstring{$\mathcal{U}(T)$}{U(T)}: an upper bound}
\label{section.bound}
Our goal in this section is to provide an upper bound for
$\mathcal{U} (T)$, which is uniform in the
various parameters.
This will allow us to reduce the range of
summation for the various parameters appearing in our expression for $N(B)$.
Our main tool will be previous work of the first two
authors \cite{bretechebrowning:sums},
which is concerned with the average order of
arithmetic functions ranging over the values taken by binary forms.
Throughout this section we continue to adhere to the convention that
all of our implied constants are allowed to depend upon the
coefficients of the forms $L_j$. Recall the expression for $\mathcal{U}(T)$
given in \eqref{eq:UUT}, with $d_j, D_j$ given by \eqref{eq:thed}.
With these in mind we have the following result.
\begin{lemma}\label{lem:UU-upper}
Let $\varepsilon>0$ and let $T\geq 1$.
Then we have
$$
\mathcal{U}(T) \ll (d\ell)^{\varepsilon} \Big(
\frac{T}{[D_1D_2,\ldots, D_3D_4]}+\frac{T^{1/2+\varepsilon}}{\ell}\Big),
$$
where $d=d_1d_2d_3d_4$.
\end{lemma}
\begin{proof}
Since we are only concerned with providing an upper bound for
$\mathcal{U}(T)$, we may drop any of the conditions in the
summation over $(u,v)$ that we care to choose. Thus it follows that
$$
\mathcal{U}(T)\leq
\sum_{(u,v)\in \mathsf{\Gamma}_{\mathbf{D}}\cap (0,\sqrt{T}]^2}
\prod_{j=1}^4
r\Big(\frac{|L_j(u,v)|}{d_{j}}\Big),
$$
where $\mathsf{\Gamma}_{\mathbf{D}}$ is the lattice defined in \eqref{eq:lattice}.
Let $\mathbf{e}_1,\mathbf{e}_2$ be a minimal basis for
$\mathsf{\Gamma}_{\mathbf{D}}$.
This is constructed by taking
$\mathbf{e}_1 \in \mathsf{\Gamma}_{\mathbf{D}}$ to be any non-zero vector for which
$|\mathbf{e}_1|$ is least, and then choosing $\mathbf{e}_2 \in \mathsf{\Gamma}_{\mathbf{D}}$
to be any vector not proportional to $\mathbf{e}_1$, for which $|\mathbf{e}_2|$ is
least. The successive minima of $\mathsf{\Gamma}_{\mathbf{D}}$ are
the numbers $s_i=|\mathbf{e}_i|,$ for $i=1,2$. They satisfy the inequalities
\begin{equation}\label{rog6}
\ell \leq s_1\leq s_2, \quad s_1s_2 \ll \rho(\mathbf{D}) \leq s_1s_2,
\end{equation}
where $\rho$ is defined in \eqref{eq:def-rho} and the lower bound for
$s_1$ follows from \eqref{sm} and the definition \eqref{eq:thed} of $D_1,D_2$.
Write $M_j(X,Y)$ for the linear form obtained from $d_j^{-1}L_j(U,V)$
via the change of variables $(U,V)\mapsto X \mathbf{e}_1+Y\mathbf{e}_2$.
Each $M_j$ has integer coefficients of size $O(\rho(\mathbf{D}))$.
Furthermore, it follows from work of Davenport \cite[lemma 5]{davenport:cubic}
that $x\ll \max\{|u|,|v|\}/s_1$ and $y\ll \max\{|u|,|v|\}/s_2$ whenever one writes
$(u,v)\in\mathsf{\Gamma}_{\mathbf{D}}$ as
$
(u,v)=x \mathbf{e}_1+y\mathbf{e}_2,
$
with $x,y\in {\mathbf Z}$. Let
$$
T_1=s_1^{-1}\sqrt{T}, \quad T_2=s_2^{-1}\sqrt{T},
$$
so that in particular $T_1\geq T_2>0$. Then we may deduce that
$$
\mathcal{U}(T)
\leq
\sum_{x \ll T_1, y\ll T_2}\prod_{j=1}^4r(|M_j(x,y)|).
$$
Suppose that $M_j(X,Y)=a_{j1}X+a_{j2}Y$,
with integer coefficients $a_{ji}=O(\rho(\mathbf{D}))$.
We proceed to introduce a multiplicative function $r_1(n)$, via
$$
r_1(p^\nu)=\left\{
\begin{array}{ll}
1+\chi(p), & \mbox{$\nu=1$ and $p \nmid 6d\ell \prod a_{ji}$,}\\
(1+\nu)^4, & \mbox{otherwise},
\end{array}
\right.
$$
where $d=d_1d_2d_3d_4$.
Then $r(n_1)r(n_2)r(n_3)r(n_4)\leq 2^8 r_1(n_1n_2n_3n_4),$
and it is not hard to see that $r_1$ belongs to the class of
non-negative arithmetic functions considered previously by the first
two authors \cite{bretechebrowning:sums}. An
application of \cite[corollary 1]{bretechebrowning:sums} now reveals that
\begin{align*}
\mathcal{U}(T)
\ll
(d\ell)^\varepsilon (T_1T_2 + T_1^{1+\varepsilon})
&\ll (d \ell)^\varepsilon \Big(\frac{T}{s_1s_2} + \frac{T^{1/2+\varepsilon}}{s_1}\Big),
\end{align*}
for any $\varepsilon>0$.
Combining \eqref{rog6} with \eqref{eq:251.1} we therefore conclude the
proof of the lemma.
\end{proof}
The main purpose of lemma \ref{lem:UU-upper} is to reduce the range of
summation of the various parameters appearing in proposition \ref{prop:6.1}.
Let us write $E_0(B)$ for the overall contribution to the summation from
values of $\mathfrak{b}_j, \ell$ such that
\begin{equation}
\label{eq:bnf}
\max \mathrm{N}(\mathfrak{b}_j)>\log (B)^{D} \quad \mbox{or} \quad \ell> \log (B)^L,
\end{equation}
for parameters $D,L>0$ to be selected in due course.
We will denote by $N_1(B)$ the remaining contribution, so that
\begin{equation}
\label{eq:N1E0}
N(B)=N_1(B)+E_0(B).
\end{equation}
Henceforth, the
implied constants in our
estimates will be allowed to depend on $D$ and $L$, in addition to
the coefficients of the linear forms $L_j$.
We proceed to establish the following result.
\begin{lemma}\label{lem:M(B)}
We have $E_0(B)\ll B\log (B)^{1-\min\{D/4,L/2\}+\varepsilon}$, for any $\varepsilon>0$.
\end{lemma}
\begin{proof}
We begin observing that $\mathcal{U}(B/t)=0$ in $E_0(B)$, unless
$D_j\leq \sqrt{B/t}$,
in the notation of \eqref{eq:thed}. But then it follows that we must have
$$
t \leq \frac{B}{\sqrt{D_1D_2D_3D_4}}\leq \frac{B
\sqrt{\gcd(\mathrm{N}(\mathfrak{b}_1),\ell)\gcd(\mathrm{N}(\mathfrak{b}_2),\ell)}}{\ell\sqrt{\mathrm{N}(\mathfrak{b}_1)\cdots
\mathrm{N}(\mathfrak{b}_4)}} =B_0,
$$
say, in the summation over $t$.
Here we have used the fact that $m_jN(\mathfrak{a}^+_j)=O(1)$ whenever ${\boldsymbol{m}}\in
\Sigma$ and ${\boldsymbol{\mathfrak a}}\in\Sigma'$.
We now apply lemma \ref{lem:UU-upper} to bound $\mathcal{U}(B/t)$, giving
\begin{align*}
E_0(B)\ll
\sum_{\substack{{\boldsymbol{m}}\in \Sigma\\ {\boldsymbol{\mathfrak a}}\in \Sigma'}}
&\sum_{\ell} \ell^\varepsilon
\sum_{\mathfrak{b}_1,\ldots,\mathfrak{b}_4} (\mathrm{N}(\mathfrak{b}_1)\cdots \mathrm{N}(\mathfrak{b}_4))^\varepsilon \\
&\times \sum_{\substack{t\leq B_0\\ \mathrm{N}(\bigcap
\mathfrak{b}_j)\mid t}}
r\Big(\frac{t}{\mathrm{N}(\bigcap \mathfrak{b}_j)}\Big)
\Big(
\frac{B}{t[D_1D_2,\ldots, D_3D_4]}+\frac{B^{1/2+\varepsilon}}{t^{1/2+\varepsilon}\ell}\Big),
\end{align*}
for any $\varepsilon>0$, where the summations over $\ell$ and $\mathfrak{b}_j$ are
subject to \eqref{eq:bnf}. In view of the elementary estimates
\begin{equation}
\label{eq:familiar}
\sum_{n\leq x} \frac{r(n)}{n^\theta} \ll \begin{cases}
\log (2x) & \mbox{if $\theta\geq 1$,}\\
x^{1-\theta} & \mbox{if $0\leq \theta<1$,}
\end{cases}
\end{equation}
we easily conclude that
\begin{align*}
E_0(B)\ll
\sum_{\substack{{\boldsymbol{m}}\in \Sigma\\ {\boldsymbol{\mathfrak a}}\in \Sigma'}}
&\sum_{\ell} \ell^\varepsilon
\sum_{\mathfrak{b}_1,\ldots,\mathfrak{b}_4} (\mathrm{N}(\mathfrak{b}_1)\cdots \mathrm{N}(\mathfrak{b}_4))^\varepsilon \\
&\times \frac{1}{\mathrm{N}(\bigcap \mathfrak{b}_j)}\Big(
\frac{B \log (B)}{[D_1D_2,\ldots,
D_3D_4]}+\frac{B^{1/2+\varepsilon}B_0^{1/2-\varepsilon}}{\ell}\Big).
\end{align*}
The second term in the inner bracket is
$$
\frac{B^{1/2+\varepsilon}B_0^{1/2-\varepsilon}}{\ell}\ll
B \cdot \frac{\gcd(\mathrm{N}(\mathfrak{b}_1),\ell)^{1/4}\gcd(\mathrm{N}(\mathfrak{b}_2),\ell)^{1/4}}
{\ell^{3/2-\varepsilon}\mathrm{N}(\mathfrak{b}_1)^{1/4-\varepsilon}\cdots \mathrm{N}(\mathfrak{b}_4)^{1/4-\varepsilon}}.
$$
Similarly, a rapid consultation with \eqref{eq:thed} reveals that the first term is
\begin{align*}
\frac{B \log (B)}{[D_1D_2,\ldots,
D_3D_4]}&\ll
\frac{B \log (B)}{(D_1D_2)^{3/4}(D_3D_4)^{1/4}}\\
&\ll
B\log (B)\cdot \frac{\gcd(\mathrm{N}(\mathfrak{b}_1),\ell)^{1/4}\gcd(\mathrm{N}(\mathfrak{b}_2),\ell)^{1/4}}
{\ell^{3/2}\mathrm{N}(\mathfrak{b}_1)^{1/4}\cdots
\mathrm{N}(\mathfrak{b}_4)^{1/4}}.
\end{align*}
Bringing these estimates together we may now conclude that
\begin{align*}
E_0(B)\ll B\log (B)
\sum_{\ell}
\sum_{\mathfrak{b}_1,\ldots,\mathfrak{b}_4}
\frac{1}{\mathrm{N}(\bigcap \mathfrak{b}_j)}\cdot \frac{\gcd(\mathrm{N}(\mathfrak{b}_1),\ell)^{1/4}\gcd(\mathrm{N}(\mathfrak{b}_2),\ell)^{1/4}}
{\ell^{3/2-\varepsilon}\mathrm{N}(\mathfrak{b}_1)^{1/4-\varepsilon}\cdots
\mathrm{N}(\mathfrak{b}_4)^{1/4-\varepsilon}},
\end{align*}
where the sums are over $\ell \in \mathbf{Z}_{>0}$ and $\mathfrak{b}_1,\ldots,\mathfrak{b}_4 \subseteq
\widehat{\mathfrak{D}}$ such that \eqref{eq:bnf} holds.
For fixed $\ell\in \mathbf{Z}_{>0}$
and $\varepsilon>0$ we proceed to estimate the sum
$$
S_{\ell}(T)=\sum_{\substack{\mathfrak{b}_1,\ldots,\mathfrak{b}_4\subseteq {\mathbf Z}[i]\\ \max
\mathrm{N}(\mathfrak{b}_j)\geq T}}
\frac{\gcd(\mathrm{N}(\mathfrak{b}_1),\ell)^{1/4}\gcd(\mathrm{N}(\mathfrak{b}_2),\ell)^{1/4}}
{\mathrm{N}(\bigcap \mathfrak{b}_j)\mathrm{N}(\mathfrak{b}_1)^{1/4-\varepsilon}\cdots
\mathrm{N}(\mathfrak{b}_4)^{1/4-\varepsilon}}.
$$
This is readily achieved via Rankin's trick and the observation
that $\mathrm{N}(\mathfrak{a})\mid \mathrm{N}(\mathfrak{a}\cap\mathfrak{b})$ for
any $\mathfrak{a}, \mathfrak{b}\subseteq {\mathbf Z}[i]$. Thus it follows that
$\mathrm{N}(\bigcap \mathfrak{b}_j)\geq [\mathrm{N}(\mathfrak{b}_1),\ldots,\mathrm{N}(\mathfrak{b}_4)]$, whence
\begin{align*}
S_{\ell}(T)
&\leq \frac{1}{T^\delta}\sum_{\mathfrak{b}_1,\ldots,\mathfrak{b}_4\subseteq{\mathbf Z}[i]}
\frac{\gcd(\mathrm{N}(\mathfrak{b}_1),\ell)^{1/4}\gcd(\mathrm{N}(\mathfrak{b}_2),\ell)^{1/4}}
{[\mathrm{N}(\mathfrak{b}_1),\ldots,\mathrm{N}(\mathfrak{b}_4)]^{1-\delta}\mathrm{N}(\mathfrak{b}_1)^{1/4-\varepsilon}\cdots
\mathrm{N}(\mathfrak{b}_4)^{1/4-\varepsilon}}\\
&\ll \frac{1}{T^\delta}\sum_{b_1,\ldots,b_4=1}^\infty
\frac{\gcd(b_1,\ell)^{1/4}\gcd(b_2,\ell)^{1/4}}
{[b_1,\ldots,b_4]^{1-\delta}b_1^{1/4-\varepsilon}\cdots
b_4^{1/4-\varepsilon}}\\
&\ll \frac{1}{T^\delta}\sum_{[k_1,k_2]\mid \ell} (k_1k_2)^\varepsilon
\sum_{b_1,\ldots,b_4=1}^\infty
\frac{1}{[b_1,\ldots,b_4]^{1-\delta}b_1^{1/4-\varepsilon}\cdots
b_4^{1/4-\varepsilon}}\\
&\ll_{\delta} \ell^\varepsilon T^{-\delta},
\end{align*}
provided that $\delta<1/4,$
as can be seen by considering the corresponding Euler product.
Armed with this we see that the overall contribution to the above
estimate for $E_0(B)$ arising from
$\ell, \mathfrak{b}_1,\ldots,\mathfrak{b}_4 $ for which $\ell>\log (B)^L$
is
\begin{align*}
&\ll B\log (B) \sum_{\ell> \log (B)^L} \ell^{-3/2+\varepsilon}
S_{\ell}(1) \ll B\log (B)^{1-L/2+\varepsilon},
\end{align*}
which is satisfactory. In a similar fashion we
see that the overall contribution to $E_0(B)$ arising from
$\ell, \mathfrak{b}_1,\ldots,\mathfrak{b}_4 $ for which $\max \mathrm{N}(\mathfrak{b}_j)>\log (B)^D$ is
\begin{align*}
&\ll
B\log (B)
\sum_{\ell} \ell^{-3/2+\varepsilon}
S_{\ell}(\log (B)^D)
\ll B\log (B)^{1-D/4+\varepsilon},
\end{align*}
which is also satisfactory. The statement of lemma \ref{lem:M(B)} is
now obvious.
\end{proof}
\section{Estimating \texorpdfstring{$\mathcal{U}(T)$}{U(T)}: an asymptotic formula}
\label{section.estimate}
In view of our work in the previous section it remains to estimate
$N_1(B)$, which we have defined
as the contribution to $N(B)$ from values of $\mathfrak{b}_j, \ell$ for which
\eqref{eq:bnf} fails. Thus
\begin{align*}
N_1(B)&=\frac 1{\cardT_{\NS}({\mathbf Q})_\textinmath{tors}}
\sum_{\substack{{\boldsymbol{m}}\in \Sigma\\ {\boldsymbol{\mathfrak a}}\in \Sigma'}}\mu({\boldsymbol{\mathfrak a}})
\hspace{-0.2cm}
\sum_{\substack{\ell \leq \log (B)^L \\ 2\nmid \ell}}
\hspace{-0.2cm}
\mu(\ell)
\hspace{-0.5cm}
\sum_{\substack{\mathfrak{b}_1,\ldots,\mathfrak{b}_4 \in \widehat{{\mathfrak{D}}}\\
\mathrm{N}(\mathfrak{b}_j)\leq \log (B)^{D}}}
\hspace{-0.2cm}
\prod_{j=1}^4 \mu(\mathfrak{b}_j)
\hspace{-0.3cm}
\sum_{\substack{t \in {\mathfrak{D}}\cap[1,B]\\ \gcd(t,\mathrm{N}({\boldsymbol{\mathfrak a}}))=1\\ \mathrm{N}(\bigcap
\mathfrak{b}_j)\mid t}}
\hspace{-0.3cm}
r\Big(\frac{t}{\mathrm{N}(\bigcap \mathfrak{b}_j)}\Big)
\mathcal{U}\Big(\frac{B}{t}\Big).
\end{align*}
Here we have inserted the condition $t\leq B$ in the summation over
$t$, since the innermost summand is visibly zero otherwise.
Whereas the previous section was primarily concerned with a uniform
upper bound for the sum $\mathcal{U}(T)$ defined in \eqref{eq:UUT}, our work in
the present section will revolve around a uniform asymptotic formula
for $\mathcal{U}(T)$. The error term that arises in our analysis will involve
the real number
\begin{equation}\label{defeta}
\eta=1-\frac{1+\log(\log (2))}{\log (2)},
\end{equation}
which has numerical value $0.086071\ldots$.
Before revealing our result for $\mathcal{U}(T)$, we must first
introduce some notation for certain local densities that emerge in the
asymptotic formula.
In fact estimating $\mathcal{U}(T)$ boils down to counting integer points on the affine variety
\begin{equation}
\label{eq:torsor}
L_j(U,V)=d_j(S_j^2+T_j^2), \quad (1\leq j\leq 4),
\end{equation}
in ${\mathbf A}_{{\mathbf Q}}^{10}$, with $U,V$ restricted to lie in a lattice
depending on $\mathbf{D}$.
Thus the expected leading constant admits an interpretation as a product of local
densities.
Given a prime $p>2$ and
$\mathbf{d},\mathbf{D}$ as in~\eqref{eq:thed}, let
$$
N_{\mathbf{d},\mathbf{D}}(p^n)=\sharp\Big\{(u,v,\mathbf{s},\mathbf{t})\in
({\mathbf Z}/p^n{\mathbf Z})^{10},\ \begin{array}{l}
L_j(u,v)\equiv d_j(s_j^2+t_j^2) \bmod{p^n}\\
D_j \mid L_j(u,v)
\end{array}
\Big\}.
$$
The $p$-adic density on \eqref{eq:torsor}
is defined to be
\begin{equation}
\label{eq:def-sigp}
\omega_{\mathbf{d},\mathbf{D}}(p)=\lim_{n\rightarrow
\infty}p^{-6n-\lambda_1-\cdots-\lambda_4}N_{\mathbf{d},\mathbf{D}}(p^n),
\end{equation}
when $p>2, $ where
\begin{equation}\label{eq:lm}
\boldsymbol{\lambda}=\big(v_p(d_{1}),\ldots,v_p(d_{4})\big),
\quad \boldsymbol{\mu}=\big(v_p(D_{1}),\ldots,v_p(D_{4})\big).
\end{equation}
When $\mathbf{d},\mathbf{D}$ are as in \eqref{eq:thed} and $p>2$, we will set
\begin{equation}
\label{eq:sig>2}
\sigma_p(\mathbf{d},\mathbf{D})=\omega_{\mathbf{d},\mathbf{D}}(p).
\end{equation}
Turning to the case $p=2$, we define
\begin{equation}
\label{eq:sig=2}
\sigma_2(\mathbf{d},\mathbf{D})=
\lim_{n\rightarrow \infty}2^{-6n}N_{\mathbf{d},\mathbf{D}}(2^n
\end{equation}
where
$$
N_{\mathbf{d},\mathbf{D}}(2^n) =
\sharp\Big\{(u,v,\mathbf{s},\mathbf{t})\in
({\mathbf Z}/2^n{\mathbf Z})^{10},\ \begin{array}{l}
L_j(u,v)\equiv d_j
(s_j^2+t_j^2) \bmod{2^n}\\
2\nmid \gcd(u,v)
\end{array}
\Big\}.
$$
Finally, we let $\omega_{\mathcal{R}_{\boldsymbol{m}}}(\infty)$ denote the usual
archimedean density of solutions to the system of equations
\eqref{eq:torsor}, with $(u,v,\mathbf{s},\mathbf{t})\in\mathcal{R}_{\boldsymbol{m}}\times
{\mathbf R}^{8}$ and where $\mathcal{R}_{\boldsymbol{m}}$ is defined in \eqref{eq:RR}.
We are now ready to record our main estimate for $\mathcal{U}(T)$.
\begin{lemma}\label{lem:Tfinal}
Recall the definitions of $\mathbf{d},\mathbf{D}$ from \eqref{eq:thed}.
Then for any $\varepsilon>0$ and $T>1$ we have
$$
\mathcal{U}(T)=c_{\mathbf{d},\mathbf{D},\mathcal{R}_{\boldsymbol{m}}} T
+O\Big(\frac{(d_1d_2d_3d_4 \ell)^\varepsilon T}{\log (T)^{
\eta-\varepsilon}}\Big),
$$
where
\begin{equation}
\label{eq:cfinal}
c_{\mathbf{d},\mathbf{D},\mathcal{R}_{\boldsymbol{m}}}=\omega_{\mathcal{R}_{\boldsymbol{m}}}(\infty)
\prod_{p\in{\mathcal P}}\sigma_p(\mathbf{d},\mathbf{D}).
\end{equation}
\end{lemma}
\begin{proof}
Our primary tool in estimating $\mathcal{U}(T)$ asymptotically is the subject
of allied work of the first two authors \cite{bretechebrowning:4linear}. We begin by bringing our
expression for $\mathcal{U}(T)$ into
a form that can be tackled by the main
results there. According to \eqref{eq:resultant}
we may assume that the binary linear forms $L_j$ are
pairwise non-proportional and primitive. Furthermore, it is clear that the region
$\mathcal{R}_{\boldsymbol{m}}\subset {\mathbf R}^2$
defined in \eqref{eq:RR} is open, bounded and convex, with a piecewise continuously
differentiable boundary such that $m_jL_j(u,v)>0$ for each $(u,v)\in \mathcal{R}_{\boldsymbol{m}}$.
A key step in applying the work of \cite{bretechebrowning:4linear} consists in
checking that the ``normalisation hypothesis''
\textsf{NH}$_2(\mathbf{d})$
is satisfied in the present context.
In fact it is easy to see that
$L_j,\mathcal{R}_{\boldsymbol{m}}$ will satisfy \textsf{NH}$_2(\mathbf{d})$
provided that
$$
L_1(U,V)\equiv d_1U \pmod{4}, \quad L_2(U,V)\equiv V \pmod{4}.
$$
The second congruence is automatic since $L_2(U,V)=V$.
Recalling that $L_1(U,V)=U$, we therefore conclude that
\textsf{NH}$_2(\mathbf{d})$ holds if $d_1\equiv 1
\bmod{4}$. Alternatively, if $d_1\equiv 3
\bmod{4}$, we make the unimodular change of variables
$(U,V)\mapsto (-U,V)$ to place ourselves in the setting of
\textsf{NH}$_2(\mathbf{d})$.
We leave the reader to check that this ultimately leads to an
identical estimate in the ensuing argument. Thus, for the purposes of
our exposition here, we may freely assume that
$L_j,\mathcal{R}_{\boldsymbol{m}}$ satisfy \textsf{NH}$_2(\mathbf{d})$ in $\mathcal{U}(T)$.
We proceed by writing
\begin{equation}
\label{eq:0501}
\mathcal{U}(T)=U_1(T)+U_2(T)+U_3(T),
\end{equation}
where $U_1(T)$ denotes the contribution to $\mathcal{U}(T)$ from $(u,v)$ such that
$2\nmid uv$,
$U_2(T)$ denotes the contribution from $(u,v)$ such that
$2\nmid u$ and $2\mid v$, and finally
$U_3(T)$ is the contribution from $(u,v)$ such that
$2\mid u$ and $2\nmid v$.
Beginning with an estimate for $U_1(T)$,
we observe that
$$
U_1(T)=S_1(\sqrt{T},\mathbf{d},\mathsf{\Gamma}_{\mathbf{D}}),
$$
in the notation of \cite[eq. (1.9)]{bretechebrowning:4linear}, with $\mathbf{d},\mathbf{D}$
given by \eqref{eq:thed}. An application of \cite[theorems 3 and
4]{bretechebrowning:4linear} with $(j,k)=(1,2)$ therefore reveals that
there exists a constant $c_1$ such that
$$
U_1(T)
=c_1T +
O\Big(\frac{(d\ell)^{\varepsilon} T}{\log(T)^{ \eta-\varepsilon}}\Big),
$$
where $d=d_1d_2d_3d_4$.
The value of the constant is given by
$$
c_1=\omega_{\mathcal{R}_{\boldsymbol{m}}}(\infty)\omega_{1,\mathbf{d}}(2)\prod_{p>2}
\omega_{\mathbf{d},\mathbf{D}}(p).
$$
Here $\omega_{\mathbf{d},\mathbf{D}}(p)$ is given by
\eqref{eq:def-sigp}
and $\omega_{\mathcal{R}_{\boldsymbol{m}}}(\infty)$ is defined prior to the
statement of the lemma.
Finally, if
$$
N_{i,\mathbf{d}}'(2^n)=\sharp\Big\{(u,v,\mathbf{s},\mathbf{t})\in
({\mathbf Z}/2^n{\mathbf Z})^{10},\ \begin{array}{l}
L_j(u,v)\equiv d_j(s_j^2+t_j^2) \bmod{2^n}\\
u \equiv 1 \bmod{4}, ~v\equiv i \bmod{2}
\end{array}
\Big\},
$$
for any $i \in \{0,1\}$, then
the corresponding $2$-adic density is given by
\begin{equation*}
\omega_{i,\mathbf{d}}(2)=\lim_{n\rightarrow \infty}2^{-6n}N_{i,\mathbf{d}}'(2^n).
\end{equation*}
Note that the notation introduced in \cite{bretechebrowning:4linear} involves an
additional subscript in $\omega_{i,\mathbf{d}}(2)$ whose presence indicates
which of the various normalisation hypotheses the $L_j,\mathcal{R}_{\boldsymbol{m}}$ are
assumed to satisfy. Since we have placed ourselves in the context of
\textsf{NH}$_2(\mathbf{d})$ in each case, we have found it reasonable to
suppress mentioning this here.
Let us now shift to a consideration of the sum $U_2(T)$ in
\eqref{eq:0501}, for which one finds that
$$
U_2(T)=S_0(\sqrt{T},\mathbf{d},\mathsf{\Gamma}_{\mathbf{D}}).
$$
Applying \cite[theorems 3 and
4]{bretechebrowning:4linear} with $(j,k)=(0,2)$ therefore yields
$$
U_2(T)
=c_2T +
O\Big(\frac{(d\ell)^{\varepsilon} T}{\log (T)^{ \eta-\varepsilon}}\Big),
$$
where now
$$
c_2=\omega_{\mathcal{R}_{\boldsymbol{m}}}(\infty)\omega_{0,\mathbf{d}}(2)\prod_{p>2}
\omega_{\mathbf{d},\mathbf{D}}(p),
$$
with notation as above.
Finally we turn to the sum $U_3(T)$ in \eqref{eq:0501}.
Making the unimodular change of variables
$(U,V)\mapsto (V,U)$, one now sees that
$$
U_3(T)=S_0(\sqrt{T};\mathbf{d},\mathsf{\Gamma}_{\mathbf{D}}^{\flat}),
$$
where now the underlying region is $\mathcal{R}_{\boldsymbol{m}}^\flat=
\{(u,v) \in {\mathbf R}^2,\ (v, u)\in \mathcal{R}_{\boldsymbol{m}}\}$
and $\mathsf{\Gamma}_{\mathbf{D}}^\flat$ is defined as for $\mathsf{\Gamma}_{\mathbf{D}}$, but with
the linear forms $L_j(U,V)$ replaced by $L_j(V,U)$.
Thus an application of \cite[theorems 3 and
4]{bretechebrowning:4linear} with $(j,k)=(0,2)$ produces
$$
U_3(T)
=c_3T +
O\Big(\frac{(d\ell)^{\varepsilon} T}{\log (T)^{ \eta-\varepsilon}}\Big),
$$
with
$$
c_3=\omega_{\mathcal{R}_{\boldsymbol{m}}^\flat}(\infty)\omega_{0,\mathbf{d}}^\flat(2)\prod_{p>2}
\omega_{\mathbf{d},\mathbf{D}}^\flat(p)=
\omega_{\mathcal{R}_{\boldsymbol{m}}}(\infty)\omega_{0,\mathbf{d}}^{\flat}(2)\prod_{p>2}
\omega_{\mathbf{d},\mathbf{D}}(p),
$$
where the superscripts $\flat$ indicate that the local densities
are taken with respect to the linear forms $L_j(V,U)$.
We are now ready to bring together our various estimates for
$U_1(T), U_2(T)$ and $U_3(T)$
in \eqref{eq:0501}. This leads to the asymptotic formula in the
statement of the lemma, with leading constant
$$
c_{\mathbf{d},\mathbf{D},\mathcal{R}_{\boldsymbol{m}}}=
\omega_{\mathcal{R}_{\boldsymbol{m}}}(\infty)\big(\omega_{1,\mathbf{d}}(2)+\omega_{0,\mathbf{d}}(2)+
\omega_{0,\mathbf{d}}^{\flat}(2)\big) \prod_{p>2} \omega_{\mathbf{d},\mathbf{D}}(p).
$$
The statement of the lemma easily follows with recourse to the
definitions
\eqref{eq:sig>2}, \eqref{eq:sig=2} of the local densities
$\sigma_p(\mathbf{d},\mathbf{D})$.
\end{proof}
We will need to consider the effect of the error term
in lemma \ref{lem:Tfinal} on the quantity $N_1(B)$ that was described
at the start of the section. Accordingly, let us write
\begin{equation}
\label{eq:vas}
N_1(B)=N_2(B)+E_1(B),
\end{equation}
where $N_2(B)$ denotes the overall contribution from the main term in
lemma \ref{lem:Tfinal} and $E_1(B)$ denotes the contribution from the
error term.
\begin{lemma}\label{lem:M(B)e1}
We have $E_1(B)\ll B\log (B)^{1+L-\eta+\varepsilon}$, for any $\varepsilon>0$.
\end{lemma}
\begin{proof}
Inserting the error term in lemma \ref{lem:Tfinal} into our expression
for $N_1(B)$, we obtain
\begin{align*}
E_1(B)
&\ll
B \log (B)^\varepsilon
\sum_{\substack{\ell \leq \log (B)^L}}
\sum_{\substack{\mathfrak{b}_1,\ldots,\mathfrak{b}_4 \in \widehat{{\mathfrak{D}}}\\
\mathrm{N}(\mathfrak{b}_j)\leq \log (B)^{D}}}
\sum_{\substack{t\leq B \\ \mathrm{N}(\bigcap
\mathfrak{b}_j)\mid t}}
r\Big(\frac{t}{\mathrm{N}(\bigcap \mathfrak{b}_j)}\Big)\cdot \frac{1}{t\log (2B/t)^{
\eta}}\\
&\ll
B \log (B)^{L+\varepsilon}
\sum_{\substack{\mathfrak{b}_1,\ldots,\mathfrak{b}_4 \in \widehat{{\mathfrak{D}}}\\
\mathrm{N}(\mathfrak{b}_j)\leq \log (B)^{D}}} \frac{1}{\mathrm{N}(\bigcap \mathfrak{b}_j)}
\sum_{\substack{t\leq B_1}} \frac{r(t)}{t\log (2B_1/t)^{\eta}},
\end{align*}
where we have written $B_1=B/\mathrm{N}(\bigcap \mathfrak{b}_j)$, for ease of notation.
Combining the familiar \eqref{eq:familiar} with partial summation, we
therefore conclude that
\begin{align*}
E_1(B)
&\ll
B \log (B)^{1+L-\eta+\varepsilon}
\sum_{\substack{\mathfrak{b}_1,\ldots,\mathfrak{b}_4 \in \widehat{{\mathfrak{D}}}\\
\mathrm{N}(\mathfrak{b}_j)\leq \log (B)^{D}}} \frac{1}{\mathrm{N}(\bigcap \mathfrak{b}_j)}\\
&\ll
B \log (B)^{1+L-\eta+\varepsilon}
\sum_{b_1,\ldots,b_4=1}^\infty
\frac{1}{[b_1,\ldots,b_4](b_1b_2b_3b_4)^\varepsilon}\\%E%
&\ll
B \log (B)^{1+L-\eta+\varepsilon}.
\end{align*}
This concludes the proof of the lemma.
\end{proof}
To be useful we will also need a uniform upper bound for the constant
\eqref{eq:cfinal} appearing in lemma \ref{lem:Tfinal}.
This is achieved in the following result.
\begin{lemma}\label{lem:c-upper}
Let $\varepsilon>0$. Then we have
$$
c_{\mathbf{d},\mathbf{D},\mathcal{R}_{\boldsymbol{m}}}\ll \frac{(D_1D_2D_3D_4)^\varepsilon}{[D_1D_2,\ldots, D_3D_4]},
$$
where $\mathbf{d},\mathbf{D}$ are given by \eqref{eq:thed}.
\end{lemma}
\begin{proof}
Now it follows from \cite[theorem 4]{bretechebrowning:4linear} that
$\omega_{\mathcal{R}_{\boldsymbol{m}}}(\infty)=\pi^4 \Vol(\mathcal{R}_{\boldsymbol{m}})\ll 1$. Similarly, it
is easy to see that
$\sigma_2(\mathbf{d},\mathbf{D})\leq 2^4$, since for any $A\in {\mathbf Z}$ there are
at most $2^{n+1}$ solutions of the congruence $s^2+t^2 \equiv A \bmod{2^n}$ by
\cite[eq. (2.5)]{bretechebrowning:4linear}.
Thus we have
$$
c_{\mathbf{d},\mathbf{D},\mathcal{R}_{\boldsymbol{m}}}\ll \prod_{p>2}|\sigma_p(\mathbf{d},\mathbf{D})|,
$$
where $\sigma_p(\mathbf{d},\mathbf{D})$ is given by \eqref{eq:sig>2}.
Assume that $p>2$. A further application of \cite[theorem~4]{bretechebrowning:4linear}
now yields
$$
\sigma_p(\mathbf{d},\mathbf{D})=
\Big(1-\frac{\chi(p)}{p}\Big)^4
\sum_{\nu_1,\ldots,\nu_4=0}^{\infty}
\frac{\chi(p)^{{\nu_1}+{\nu_2}+{\nu_3}+{\nu_4}}}{\rho(p^{\max\{\mu_1,\lambda_1+\nu_1\}},
\ldots, p^{\max\{\mu_4,\lambda_4+\nu_4\}})},
$$
where $\rho$ is the determinant given in \eqref{eq:def-rho} and
$\boldsymbol{\lambda}, \boldsymbol{\mu}$ are given by \eqref{eq:lm}. Using the multiplicativity of
$\rho$ we may clearly write
$$
\prod_{p>2}|\sigma_p(\mathbf{d},\mathbf{D})|=\frac{1}{\rho(\mathbf{D})}
\prod_{p>2}|\sigma_p'(\mathbf{d},\mathbf{D})|,
$$
where now
$$
\sigma_p'(\mathbf{d},\mathbf{D})=
\Big(1-\frac{\chi(p)}{p}\Big)^4
\sum_{\nu_1,\ldots,\nu_4=0}^{\infty}
\frac{\chi(p)^{{\nu_1}+{\nu_2}+{\nu_3}+{\nu_4}}\rho(p^{\mu_1},
\ldots, p^{\mu_4})}{\rho(p^{\max\{\mu_1,\lambda_1+\nu_1\}},
\ldots, p^{\max\{\mu_4,\lambda_4+\nu_4\}})}.
$$
In view of \eqref{eq:251.1}, it will suffice to
show that
\begin{equation}
\label{eq:suffice}
\prod_{p>2}|\sigma_p'(\mathbf{d},\mathbf{D})|\ll (D_1D_2D_3D_4)^\varepsilon,
\end{equation}
in order to complete the proof of the lemma.
Recall the definition \eqref{eq:resultant} of $\Delta$ and write $D=D_1D_2D_3D_4$. Then
for $p\nmid \Delta D$
it follows from \eqref{eq:tenn''} that
$$
\sigma_p'(\mathbf{d},\mathbf{D})=
\Big(1-\frac{\chi(p)}{p}\Big)^4
\sum_{\nu_1,\ldots,\nu_4=0}^{\infty}
\frac{\chi(p)^{{\nu_1}+{\nu_2}+{\nu_3}+{\nu_4}}}{p^{m(\boldsymbol{\nu})}},
$$
where $m(\boldsymbol{\nu})$ is defined in section \ref{geom-series}.
On refamiliarising oneself with the notation $S_0^\varepsilon(z)$ introduced
in \eqref{30-S0pm}, lemma \ref{31-s0-} therefore yields
$$
\sigma_p'(\mathbf{d},\mathbf{D})=
\Big(1-\frac{1}{p}\Big)^4
S_0^{+}(1/p)=
\frac{1+2/p+6/p^2+2/p^3+1/p^4}{(1+1/p)^2},
$$
if $p\equiv 1 \bmod 4$, and
$$
\sigma_p'(\mathbf{d},\mathbf{D})=
\Big(1+\frac{1}{p}\Big)^4
S_0^{-}(1/p)=\frac{(1-1/p^2)^2}{(1+1/p^2)},
$$
if $p\equiv 3 \bmod 4$. Thus
$\sigma_p'(\mathbf{d},\mathbf{D})=1+O(1/p^2)$ for $p\nmid \Delta D$.
Suppose now that $p\mid \Delta D$. Then \eqref{eq:tenn-} implies that
$$
\sigma_p'(\mathbf{d},\mathbf{D})\ll
\sum_{\nu_1,\ldots,\nu_4=0}^{\infty}
\frac{1}{p^{m(\mathbf{n})-m(\boldsymbol{\mu})}}\ll 1
$$
where $\mathbf{n}=
(\max\{\mu_1,\lambda_1+\nu_1\},\ldots, \max\{\mu_4,\lambda_4+\nu_4\})$.
Putting this together with our treatment of the factors corresponding
to $p\nmid \Delta D$, we are easily led to the desired upper bound in \eqref{eq:suffice}.
This therefore concludes the proof of the lemma.
\end{proof}
\section{The d\'enouement}
Take $D=4$ and $L=2\eta/3$ in lemmas \ref{lem:M(B)}
and lemma \ref{lem:M(B)e1}, and let $\varepsilon>0$ be given. We therefore deduce that
$$
N(B)=N_2(B)+O\big(B\log (B)^{1-\eta/3+\varepsilon} \big)
$$
via \eqref{eq:N1E0} and \eqref{eq:vas}, where $\cardT_{\NS}({\mathbf Q})_\textinmath{tors}
\times N_2(B)$ is equal to
\begin{align*}
B\sum_{\substack{{\boldsymbol{m}}\in \Sigma\\ {\boldsymbol{\mathfrak a}}\in \Sigma'}}\mu({\boldsymbol{\mathfrak a}})
\hspace{-0.1cm}
\sum_{\substack{\ell \leq \log (B)^{2\eta/3} \\ 2\nmid \ell}}
\hspace{-0.1cm}
\mu(\ell)
\hspace{-0.1cm}
\sum_{\substack{\mathfrak{b}_1,\ldots,\mathfrak{b}_4 \in \widehat{{\mathfrak{D}}}\\
\mathrm{N}(\mathfrak{b}_j)\leq \log (B)^{4}}}
\prod_{j=1}^4 \mu(\mathfrak{b}_j)
c_{\mathbf{d},\mathbf{D},\mathcal{R}_{\boldsymbol{m}}}
\hspace{-0.1cm}
\sum_{\substack{t \in {\mathfrak{D}}\cap[1,B]\\ \gcd(t,\mathrm{N}({\boldsymbol{\mathfrak a}}))=1\\ \mathrm{N}(\bigcap
\mathfrak{b}_j)\mid t}}
\hspace{-0.1cm}
\frac{r(t/\mathrm{N}(\bigcap \mathfrak{b}_j))}{t}.
\end{align*}
Here $ c_{\mathbf{d},\mathbf{D},\mathcal{R}_{\boldsymbol{m}}}$ is given by
\eqref{eq:cfinal}, with $\mathbf{d},\mathbf{D}$ being given by \eqref{eq:thed}
and $\mathcal{R}_{\boldsymbol{m}}$ given by \eqref{eq:RR}.
The following simple result allows us to carry out the inner summation over
$t$.
\begin{lemma}\label{lem:r}
Let $m\in \mathbf{Z}_{>0}$ and let $T\geq 1$. Then for any $\varepsilon>0$ we have
$$
\sum_{\substack{t \in {\mathfrak{D}}\cap[1,T]\\ \gcd(t,m)=1}}
\frac{r(t)}{t}= C_m \log (T)
+O(m^\varepsilon),
$$
where
$$
C_m=2L(1,\chi)
\prod_{p\equiv 3
\bmod{4}}\Big(1-\frac{1}{p^{2}}\Big)
\prod_{\substack{p\mid m\\ p\equiv 1\bmod{4}}} \Big(1-\frac{1}{p}\Big)^2.
$$
\end{lemma}
\begin{proof}
Recall the definition \eqref{18-D} of the set ${\mathfrak{D}}$.
We consider the Dirichlet series
$$
F_m(s)=\sum_{\substack{t\in {\mathfrak{D}}\\ \gcd(t,m)=1}}
\frac{r(t)}{t^s}
=4\prod_{\substack{p\nmid m\\ p\equiv 1 \bmod{4}}}
\sum_{k\geq 0}\frac{k+1}{p^{ks}}
=4\prod_{\substack{p\nmid m\\ p\equiv 1 \bmod{4}}}
\Big(1-\frac{1}{p^s}\Big)^{-2},
$$
for $\Re e(s)>1$.
Thus we may write $F_m(s)=F_1(s)H(s)$, with
$$
H(s)=\prod_{\substack{p\mid m\\ p\equiv 1 \bmod{4}}}
\Big(1-\frac{1}{p^s}\Big)^{2}=\sum_{d=1}^\infty \frac{h(d)}{d^s},
$$
say, for an appropriate arithmetic function $h$. One calculates
\begin{align*}
F_1(s)=4\zeta(s)L(s,\chi) \Big(1-\frac{1}{2^{s}}\Big)\prod_{p\equiv 3
\bmod{4}} \Big(1-\frac{1}{p^{2s}}\Big),
\end{align*}
whence an application of Perron's formula yields
$$
\sum_{\substack{t \in {\mathfrak{D}}\cap[1,T]}}
\frac{r(t)}{t}= C_1\log (T) +O(1),
$$
with $C_1$ defined as in the statement of the lemma.
We may complete the proof of the lemma using an argument based on
Dirichlet convolution. Thus it follows that
\begin{align*}
\sum_{\substack{t \in {\mathfrak{D}}\cap[1,T]\\ \gcd(t,m)=1}}
\frac{r(t)}{t}
&= \sum_{\substack{d \mid m^2\\ d \in {\mathfrak{D}}\cap [1,T]}}
\frac{h(d)}{d}\Big(C_1 \log\Big(\frac{T}{d}\Big) +O(1)\Big)\\
&= \prod_{\substack{p\mid m\\ p\equiv 1\bmod{4}}} \Big(1-\frac{1}{p}\Big)^2
C_1 \log (T) +O\Big(\sum_{d\mid m^2} \frac{|h(d)|\log (2d)}{d}\Big).
\end{align*}
The main term confirms the prediction in the statement of the lemma
and the error term is easily seen to be $O(m^\varepsilon)$ for any $\varepsilon>0$,
which is satisfactory.
\end{proof}
Making the obvious change of variables it now follows from lemma
\ref{lem:r} that
\begin{align*}
\sum_{\substack{t \in {\mathfrak{D}}\cap[1,B]\\ \gcd(t,\mathrm{N}({\boldsymbol{\mathfrak a}}))=1\\ \mathrm{N}(\bigcap
\mathfrak{b}_j)\mid t}}
\frac{r(t/\mathrm{N}(\bigcap \mathfrak{b}_j))}{t}
&= \frac{c_{{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}}}\log (B/\mathrm{N}(\bigcap
\mathfrak{b}_j))}{\mathrm{N}(\bigcap
\mathfrak{b}_j)}
+O(1)\\
&= \frac{c_{{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}}}\log (B)}{\mathrm{N}(\bigcap
\mathfrak{b}_j)}
+O(1),
\end{align*}
where
$$
c_{{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}}}=
\begin{cases}
C_{\mathrm{N}({\boldsymbol{\mathfrak a}})} & \mbox{if $\gcd(\mathrm{N}(\bigcap \mathfrak{b}_j),\mathrm{N}({\boldsymbol{\mathfrak a}}))=1$,}\\
0 & \mbox{otherwise}.
\end{cases}
$$
In particular it is clear that
$c_{{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}}}=O(1)$.
Applying lemma \ref{lem:c-upper} it is easy to conclude that
the overall contribution to $N_2(B)$ from the error term in this estimate is
\begin{align*}
&\ll
B\sum_{\ell \leq \log (B)^{2\eta/3}}
\ell^\varepsilon
\sum_{\mathrm{N}(\mathfrak{b}_j)\leq \log (B)^{4}}
\frac{(\mathrm{N}(\mathfrak{b}_1)\cdots \mathrm{N}(\mathfrak{b}_4))^\varepsilon}{[
\mathrm{N}(\mathfrak{b}_1)\mathrm{N}(\mathfrak{b}_2), \ldots, \mathrm{N}(\mathfrak{b}_3)\mathrm{N}(\mathfrak{b}_4)]}\\
&\ll
B\log (B)^{2\eta/3+\varepsilon} \sum_{b_1,\ldots,b_4\leq \log (B)^{4}}
\frac{1}{[b_1b_2, \ldots, b_3b_4]}\\
&
\leq B\log (B)^{2\eta/3+\varepsilon} \prod_{p\leq \log (B)^4} S_0^+(1/p),
\end{align*}
in the notation of \eqref{30-S0pm}. This is therefore seen to be
$O(B\log (B)^{2\eta/3+\varepsilon})$ via lemma \ref{31-s0-}.
In conclusion, we may write
$$
N(B)=N_3(B)+O\big(B\log (B)^{1-\eta/3+\varepsilon} \big),
$$
where now
\begin{align*}
N_3(B)=
\frac{B\log (B)}{\cardT_{\NS}({\mathbf Q})_\textinmath{tors}}
\sum_{\substack{{\boldsymbol{m}}\in \Sigma\\ {\boldsymbol{\mathfrak a}}\in \Sigma'}}\mu({\boldsymbol{\mathfrak a}})
\sum_{\substack{\ell \leq \log (B)^{2\eta/3} \\ 2\nmid \ell}}
\mu(\ell)
\sum_{\substack{\mathfrak{b}_1,\ldots,\mathfrak{b}_4 \in \widehat{{\mathfrak{D}}}\\
\mathrm{N}(\mathfrak{b}_j)\leq \log (B)^{4}}} \frac{c_{{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}}}c_{\mathbf{d},\mathbf{D},\mathcal{R}_{\boldsymbol{m}}}}{\mathrm{N}(\bigcap
\mathfrak{b}_j)} \prod_{j=1}^4 \mu(\mathfrak{b}_j) .
\end{align*}
Here we have used \eqref{defeta} to observe that $1-\eta/3>2\eta/3$.
Finally, through a further application of lemma \ref{lem:c-upper}, it
is now a trivial matter to re-apply the proof of lemma \ref{lem:M(B)}
to show that the summations over $\ell$ and
$\mathfrak{b}_j$ can be extended to infinity with error
$O(B\log (B)^{1-\eta/3+\varepsilon} )$.
This therefore leads to the final outcome that
$$
N(B)=cB\log (B)+O\big(B\log (B)^{1-\eta/3+\varepsilon} \big),
$$
for any $\varepsilon>0$, where
\begin{equation}
\label{equ.final.constant}
c=\frac 1{\cardT_{\NS}({\mathbf Q})_\textinmath{tors}}
\sum_{\substack{{\boldsymbol{m}}\in \Sigma\\ {\boldsymbol{\mathfrak a}}\in \Sigma'}}\mu({\boldsymbol{\mathfrak a}})
\sum_{\substack{\ell=1\\ 2\nmid \ell}}^\infty
\mu(\ell)
\sum_{\substack{\mathfrak{b}_1,\ldots,\mathfrak{b}_4 \in \widehat{{\mathfrak{D}}}}}
\frac{c_{{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}}}c_{\mathbf{d},\mathbf{D},\mathcal{R}_{\boldsymbol{m}}}}{\mathrm{N}(\bigcap
\mathfrak{b}_j)} \prod_{j=1}^4 \mu(\mathfrak{b}_j) .
\end{equation}
Here $c_{\mathbf{d},\mathbf{D},\mathcal{R}_{\boldsymbol{m}}}$ is given by
\eqref{eq:cfinal}, with $\mathbf{d},\mathbf{D}$ being given by \eqref{eq:thed}.
\section{Jumping down}
We shall now relate the constant $c$
defined by equation~\eqref{equ.final.constant}
with the one expected, as required to complete
the proof of theorem~\ref{theo.main}.
\subsection{Expression in terms of volumes}
Let us first recall that the adelic set ${\matheusm T}_{\boldsymbol{n}}({\boldsymbol A}_{\mathbf Q})$
comes with a canonical measure which is defined as follows.
The canonical line bundle on $\omega_{{\matheusm T}_{\boldsymbol{n}}}$ is trivial
\cite[lemme 3.1.12]{peyre:cercle}
and the invertible functions on ${\matheusm T}_{\boldsymbol{n}}$ are constant.
Therefore up to multiplication by a constant there exists
a unique section $\formon{{\matheusm T}_{\boldsymbol{n}}}$ of $\omega_{{\matheusm T}_{\boldsymbol{n}}}$
which does not vanish.
By \cite[\S2]{weil:adeles}, this form defines a measure
$\boldsymbol\omega_{{\matheusm T}_{\boldsymbol{n}},v}$ on ${\matheusm T}_{\boldsymbol{n}}({\mathbf Q}_v)$ for any
place $v$ of ${\mathbf Q}$. According to \cite[lemme 3.1.14]{peyre:cercle},
the product $\prod_v\boldsymbol\omega_{{\matheusm T}_{\boldsymbol{n}},v}$ converges
and defines a measure on ${\matheusm T}_{\boldsymbol{n}}({\boldsymbol A}_{\mathbf Q})$. By the product
formula, this measure does not depend on the choice of the section
$\formon{{\matheusm T}_{\boldsymbol{n}}}$.
Let us now describe explicitly how to construct such
a section $\formon{{\matheusm T}_{\boldsymbol{n}}}$.
\begin{nota}
Let $\mathcal X_{\boldsymbol{n}}$ be the
subscheme of ${\mathbf A}^8_{\mathbf Z}=\Spec({\mathbf Z}[X_j,Y_j,1\leq j\leq 4])$
defined by the equations~\eqref{equ.torcz}. Then ${\mathcal Y}_{\boldsymbol{n}}$ is
the product $\mathcal X_{\boldsymbol{n}}\times{\mathbf A}^2_{\mathbf Z}$.
We denote by $\mathcal X^\circ_{\boldsymbol{n}}$ the complement of the origin
in $\mathcal X_{\boldsymbol{n}}$.
For three distinct elements $j,k,l$
of ${\{1,2,3,4\}}$, let us denote by $P_{j,k,l}$ the quadratic form
\[\Delta_{j,k}n_l(X^2_l+Y^2_l)+\Delta_{k,l}n_j(X_j^2+Y_j^2)
+\Delta_{l,j}n_k(X_k^2+Y_k^2).\]
Then we have the relations
\begin{align*}
a_jP_{k,l,m}+a_kP_{l,m,j}+a_lP_{m,j,k}+a_mP_{j,k,l}&=0\\
b_jP_{k,l,m}+b_kP_{l,m,j}+b_lP_{m,j,k}+b_mP_{j,k,l}&=0
\end{align*}
whenever $\{j,k,l,m\}={\{1,2,3,4\}}$.
Since $\Delta_{1,2}=1$, the scheme $\mathcal X^\circ_{\boldsymbol{n}}$ is the complete
intersection in ${\mathbf A}^6_{\mathbf Z}\setminus\{0\}$
of the quadrics defined by $P_{1,2,3}$ and $P_{1,2,4}$.
Therefore the corresponding Leray form is a nonzero
section of the canonical line bundle $\omega_{\mathcal X^\circ_{{\boldsymbol{n}},{\mathbf Q}}}$.
On ${\mathbf A}^2_{\mathbf Z}$, we may take the natural form
$\frac\partial{\partial X_0}\wedge\frac\partial{\partial Y_0}$.
The exterior product of these forms
gives a form on an open subset of $\mathcal Y_{\boldsymbol{n}}$,
and by restriction a form $\formon{{\matheusm T}_{\boldsymbol{n}}}$
on ${\matheusm T}_{\boldsymbol{n}}$ which does not vanish.
We denote by $\boldsymbol\omega_{{\boldsymbol{n}},v}$ the corresponding measure on
$\mathcal Y_n({\mathbf Q}_v)$ for $v\in\Val({\mathbf Q})$.
\end{nota}
\begin{lemma}
\label{lemma.volumeprime}%
Let ${\boldsymbol{m}}\in\Sigma$ and ${\boldsymbol{\mathfrak a}}\in\Sigma'$.
Let ${\boldsymbol{\mathfrak b}}=(\mathfrak{b}_j)_{j\in{\{1,2,3,4\}}}$ belong to $\widehat{\mathfrak{D}}^4$.
Let~$\ell$ be an odd integer.
Let $d_j$ and $D_j$ be defined by formula~\eqref{eq:thed}.
Then for any prime number $p$ we have
\[\boldsymbol\omega_{{\boldsymbol{n}},p}(\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell,p})
=\beta_pp^{-v_p\bigl(\mathrm{N}\bigl(\bigcap_j\mathfrak{b}_j\bigr)\bigr)}
\lim_{n\to+\infty}p^{-6n}N_{\mathbf{d},\mathbf{D}}(p^n),\]
where
\[\beta_p=\begin{cases}
\frac 12&\text{if $p=2$},\\
1-\frac 1{p^2}&\text{if $p\equiv 3\mod 4$},\\
\left(1-\frac 1p\right)^2&\text{if $p\mid \prod_j\mathrm{N}(\mathfrak{a}^+_j)$
and $p\equiv 1\mod 4$},\\
0&\text{if $p\mid\prod_j\mathrm{N}(\mathfrak{a}^+_j)$ and $p\mid\prod_j\mathrm{N}(\mathfrak{b}_j)$},\\
1&\text{otherwise.}
\end{cases}\]
\end{lemma}
\begin{proof}
In the product $\mathcal X_{N({\boldsymbol{\mathfrak a}}{\boldsymbol{\mathfrak b}}){\boldsymbol{m}}}\times{\mathbf A}^2_{\mathbf Z}$,
the domain $\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell,p}$ decomposes as
a product.
The projection on the eight coordinates $X_j,Y_j$, where $j\in{\{1,2,3,4\}}$,
gives an isomorphism from
the complete intersection in ${\mathbf A}^{10}_{\mathbf Z}-\{0\}$
given by the equations
\[L_j(U,V)=n_j(X_j^2+Y_j^2)\]
for $j\in{\{1,2,3,4\}}$ to the scheme $\mathcal X^\circ_{\boldsymbol{n}}$.
Moreover this isomorphism map is compatible with the respective Leray forms.
Since the measure defined by the Leray measure coincides
with the counting measure (see, for example,
\cite[proposition 1.14]{lachaud:waring}),
the volume of the first component is equal to
$\lim_{n\to+\infty}p^{-6n}N_{\mathbf{d},\mathbf{D}}(p^n)$.
The measure on ${\mathbf A}^2_{\mathbf Z}$ is the standard Haar measure.
On the other hand, the image of the domain in ${\mathbf Z}_p^2$ may be described
as follows:
\begin{itemize}
\item It is ${\mathbf Z}[{\mathsl i}]_{1+{\mathsl i}}\setminus(1+{\mathsl i}){\mathbf Z}[{\mathsl i}]_{1+{\mathsl i}}$ if $p=2$;
\item It is ${\mathbf Z}_p^2\setminus p{\mathbf Z}_p^2$ if $p\equiv 3\mod 4$;
\item It is the set of $(x,y)\in{\mathbf Z}_p^2$
such that $p$ does not divide $\mathrm{N}(x+{\mathsl i} y)$ if $p\mid\prod_j\mathrm{N}(\mathfrak{a}^+_j)$,
the prime~$p$ does not divide $\mathrm{N}(\bigcap_j\mathfrak{b}_j)$
and $p\equiv 1\mod 4$;
\item It is empty if $p\mid\prod_j\mathrm{N}(\mathfrak{a}^+_j)$ and $p\mid\prod_j\mathrm{N}(\mathfrak{b}_j)$;
\item It is $(\bigcap_j\mathfrak{b}_j){\mathbf Z}_p[{\mathsl i}]$ otherwise.
\end{itemize}
Therefore $\beta_pp^{-v_p\bigl(\mathrm{N}\bigl(\bigcap_j\mathfrak{b}_j\bigr)\bigr)}$
is the volume of this component.
\end{proof}
\begin{lemma}
\label{lemma.volumeinfinite}%
Let ${\boldsymbol{m}}\in\Sigma$ and ${\boldsymbol{\mathfrak a}}\in\Sigma'$.
Let ${\boldsymbol{\mathfrak b}}=(\mathfrak{b}_j)_{j\in{\{1,2,3,4\}}}$ belong to $\widehat{\mathfrak{D}}^4$.
We put ${\boldsymbol{n}}=\mathrm{N}({\boldsymbol{\mathfrak a}}{\boldsymbol{\mathfrak b}}){\boldsymbol{m}}$.
Let~$\ell$ be an odd integer. For any real number $B$, we have
\[\boldsymbol\omega_{{\boldsymbol{n}},\infty}(\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell,\infty}(B))=
\frac{4L(1,\chi)
\pi^4}{\prod_{j=1}^4n_j}\Vol(\mathcal R_{\boldsymbol{m}})f(B),\]
where $f(B)=\int_0^{\log(B)}ue^u \,\textinmath{d} u=B\log(B)-B+1$.
\end{lemma}
\begin{proof}
The functions $U$ and $V$ on ${\mathcal Y}_{\boldsymbol{n}}=\mathcal X_{\boldsymbol{n}}\times{\mathbf A}^2$
are induced by functions on $\mathcal X_{\boldsymbol{n}}$ which we shall
also denote by $U$ and $V$. Let $H_{F,\infty}:\mathcal X_{\boldsymbol{n}}({\mathbf R})\to{\mathbf R}$
and $H_{E,\infty}:{\mathbf R}^2\to{\mathbf R}$ be defined by
\[H_{F,\infty}(R)=\max(|U(R)|,|V(R)|)\qquad\text{and}\qquad
H_{E,\infty}(x_0,y_0)=x_0^2+y_0^2.\]
Then the domain $\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell,\infty}(B)$ is the set
of $(R,(x_0,y_0))\in\mathcal X_{\boldsymbol{n}}({\mathbf R})\times{\mathbf R}^2$ such that
\[H_{F,\infty}(R)\geq 1,\qquad H_{E,\infty}(x_0,y_0)\geq 1,\qquad\text{and}\qquad
H_{F,\infty}(R)^2H_{E,\infty}(x_0,y_0)\leq B.\]
Let us denot by $v_{{\boldsymbol{n}},1}(t)$ (resp. $v_2(t)$) the
volume of the set of $R\in\mathcal X_{\boldsymbol{n}}({\mathbf R})$ (resp. $(x_0,y_0)\in{\mathbf R}^2$)
such that $H_{F,\infty}(R)\leq t$ (resp. $H_{E,\infty}(x_0,y_0)\leq t$).
Then the functions $v_{{\boldsymbol{n}},1}$ and $v_2$ are monomials of respective
degrees $2$ and $1$. Therefore the volume
of the domain $\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell,\infty}(B)$
is given by
\[v_{{\boldsymbol{n}},1}(1)v_2(1)\int_{\substack{t\geq 1, u\geq 1\\t^2u\leq B}} 2t\,\textinmath{d} u\,\textinmath{d} t
=v_{{\boldsymbol{n}},1}(1)v_2(1)f(B).\]
To compute the value of $v_{{\boldsymbol{n}},1}(1)$, we may use the change
of variables $x_j'=\sqrt{|n_j|}x_j$ and $y'_j=\sqrt{|n_j|}y_j$.
Since the Leray form may be locally described as
\[\left|\begin{matrix}\frac{\partial P_{1,2,3}}{\partial X_1}
&\frac{\partial P_{1,2,3}}{\partial X_2}\\
\frac{\partial P_{1,2,4}}{\partial X_1}
&\frac{\partial P_{1,2,4}}{\partial X_2}\\\end{matrix}\right|^{-1}
\,\textinmath{d} X_3 \,\textinmath{d} X_4\prod_{j=1}^4\,\textinmath{d} Y_j=(4\Delta_{3,4}X_1X_2)^{-1}
\,\textinmath{d} X_3\,\textinmath{d} X_4\prod_{j=1}^4\,\textinmath{d} Y_j\]
we get that $v_{{\boldsymbol{n}},1}(1)=v_{\boldsymbol\varepsilon,1}(1)\prod_{j=1}^4n_j^{-1}$,
where $\varepsilon_j=\sgn(n_j)=\sgn(m_j)$. It follows
that $v_{{\boldsymbol{n}},1}(1)=(\prod_{j=1}^4n_j)^{-1}\pi^4\Vol(\mathcal R_{\boldsymbol{m}})$.
We conclude the proof with the equalities $v_2(1)=\pi=4L(1,\chi)$.
\end{proof}
\begin{prop}
Let ${\boldsymbol{m}}\in\Sigma$ and ${\boldsymbol{\mathfrak a}}\in\Sigma'$.
Let ${\boldsymbol{\mathfrak b}}=(\mathfrak{b}_j)_{j\in{\{1,2,3,4\}}}$ belong to $\widehat{\mathfrak{D}}^4$.
Let~$\ell$ be an odd integer. Then
\[\frac{c_{{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}}}c_{\mathbf{d},\mathbf{D},\mathcal{R}}}{\mathrm{N}(\bigcap
\mathfrak{b}_j)}f(B)=\Vol(\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell}(B)),\]
where $f(B)=B\log(B)-B+1$.
\end{prop}
\begin{proof}
This follows from lemmata~\ref{lemma.volumeprime}
and \ref{lemma.volumeinfinite}: indeed, by
\cite[(2.8)]{bretechebrowning:4linear},
we have $\omega_{\mathcal R_{\boldsymbol{m}}}(\infty)=
\pi^4\Vol(\mathcal R_{\boldsymbol{m}})$ and
\[\prod_{p\in{\mathcal P}}\sigma_p(\mathbf{d},\mathbf D)=\frac 1{\prod_{j=1}^4n_j}
\prod_{p\in{\mathcal P}}\lim_{k\to+\infty}p^{-6k}N_{\mathbf{d},\mathbf{D}}(p^k)\]
where ${\boldsymbol{n}}=N({\boldsymbol{\mathfrak a}}{\boldsymbol{\mathfrak b}}){\boldsymbol{m}}$.
\end{proof}
\Subsection{Moebius reversion}
\begin{prop}
Let $B$ be a real number and ${\boldsymbol{m}}$ belong to $\Sigma$.
Then
\[\Vol(\mathcal D_{{\boldsymbol{m}}}(B))=\sum_{{\boldsymbol{\mathfrak a}}\in\Sigma'}\sum_{{\boldsymbol{\mathfrak b}}\in\widehat
{\mathfrak D}^4}
\sum_{\ell\operatorname{ odd}}
\mu({\boldsymbol{\mathfrak a}})\mu({\boldsymbol{\mathfrak b}})\mu(\ell)
\Vol(\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell}(B)).\]
\end{prop}
\begin{proof}
For any ${\boldsymbol{\lambda}}\in T_{\boldsymbol\Delta}({\mathbf Q})\cap{\mathbf Z}_{\boldsymbol\Delta}$,
and any ${\boldsymbol{n}}\in{\mathbf Z}^4$,
the multiplication by ${\boldsymbol{\lambda}}$ defines an isomorphism
from ${\mathcal Y}_{N({\boldsymbol{\lambda}}){\boldsymbol{n}}}$ to ${\mathcal Y}_{{\boldsymbol{n}}}$. Therefore
it sends the canonical form on the adelic set
${\mathcal Y}_{N({\boldsymbol{\lambda}}){\boldsymbol{n}}}({\boldsymbol A}_{\mathbf Q})$ onto the
canonical form on ${\mathcal Y}_{{\boldsymbol{n}}}({\boldsymbol A}_{\mathbf Q})$.
Therefore the volume of $\mathcal D^3_{{\boldsymbol{m}},{\boldsymbol{\mathfrak a}},{\boldsymbol{\mathfrak b}},\ell}(B)$
coincides with the volume of its image in ${\mathcal Y}_{{\boldsymbol{m}}}({\boldsymbol A}_{\mathbf Q})$.
The formula then follows from lemma~\ref{lemma.firstinversion}
and the proofs of propositions~\ref{prop.secondinversion}
and~\ref{prop.aftermoebius}.
\end{proof}
\Subsection{The constant}
\begin{prop}
We have
\[C_H(S)B\log(B)=\frac 1{\cardT_{\NS}({\mathbf Q})_\textinmath{tors}}
\sum_{{\boldsymbol{m}}\in\Sigma}\Vol(\mathcal D_{{\boldsymbol{m}}}(B))+O(B).\]
\end{prop}
\begin{proof}
The following proof is based upon the ideas
of Per Salberger~\cite{salberger:tamagawa}
as des\-cri\-bed in~\cite[\S5.3]{peyre:cercle}.
We may identify $\omega_S^{-1}$ with $\mathcal O_{S'}(1)$
(see lemma~\ref{lemma.linearomega}). This enables us
to define an adelic metric on $\omega_S^{-1}$ by
\[\Vert y\Vert_v=\begin{cases}
\min\left(\left|\frac{y}{X_0(x)}\right|,
\left|\frac{y}{X_1(x)}\right|,
\left|\frac{y}{X_2(x)}\right|,
C\left|\frac{y}{X_3(x)}\right|,
C\left|\frac{y}{X_4(x)}\right|\right)&\text{if $v=\infty$,}\\
\min_{0\leq i\leq 4}\left(\left|\frac{y}{X_i(x)}\right|_v
\right)&\text{otherwise.}
\end{cases}\]
for $x\in S'({\mathbf Q}_v)$ and $y$ in the corresponding fiber
$\mathcal O_{S'}(1)_x\otimes{\mathbf Q}_v$, with the constant $C$ defined
in notation~\ref{nota.height}. This adelic metric defines the height
used throughout the text.
Let $v$ be a place of ${\mathbf Q}$. We denote by ${\boldsymbol\omega}_{H,v}$
the measure on $S({\mathbf Q}_v)$ corresponding
to the adelic metric on $\omega_S^{-1}$ (see \cite[\S2]{peyre:fano}).
Let us recall that on a split torus ${\mathbf G}_m^n$,
the form $\bigwedge_{j=1}^n\xi_j^{-1}d\xi_j$, where
$(\xi_j)_{1\leq j\leq n}$ is a basis of $X^*({\mathbf G}_m^n)$,
up to sign does not depend on the choice of the basis.
Therefore there is a canonical Haar measure on $T_{\NS}({\mathbf Q}_v)$
which we shall denote by ${\boldsymbol\omega}_{T_{\NS},v}$.
Let ${\boldsymbol{m}}$ be an element of $\Sigma$.
The functions $H_w$ defined in definition~\ref{defi.localheight}
may been seen as the composite of the metrics on $\omega_S^{-1}$
with the natural morphism from the universal torsor
${\matheusm T}_{\boldsymbol{m}}$ to the line bundle $\omega_S^{-1}$.
Let $U\neq\emptyset$ be an open subset of $\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_v))$.
According to \cite[lemme 3.1.14]{peyre:cercle} and
\cite[\S4.4]{peyre:torseurs}, if $s:U\to{\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_v)$ is a continuous
section of $\pi_{\boldsymbol{m}}$, then the measure ${\boldsymbol\omega}_{{\boldsymbol{m}},v}$ is characterised by
the relation
\begin{equation}
\label{equ.meas.torsor}
\int_{\pi_{\boldsymbol{m}}^{-1}(U)}f(y){\boldsymbol\omega}_{{\boldsymbol{m}},v}(y)=\int_U\int_{T_{\NS}({\mathbf Q}_v)}f(t.s(x))
H_{v}(t.s(x)){\boldsymbol\omega}_{T_{\NS},v}(t){\boldsymbol\omega}_{H,v}(x)
\end{equation}
for any continuous function $f$ on $\pi_{\boldsymbol{m}}^{-1}(U)$ with compact support.
By lemmata~\ref{lemma.most.places} and~\ref{lemma.bad.places},
for any prime number $p$,
$\mathcal D_{{\boldsymbol{m}},p}$ is a fundamental domain in ${\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)$
under the action of $T_{\NS}({\mathbf Q}_p)$ modulo $T_{\NS}({\mathbf Z}_p)$.
Moreover, by definition, we have that $\mathcal D_{{\boldsymbol{m}},p}$ is
contained in $\widehat\pi_{\boldsymbol{m}}^{-1}({\mathcal T}_\textinmath{spl}({\mathbf Z}_p))$ and thus
$H_p$ is equal to $1$ on $\mathcal D_{{\boldsymbol{m}},p}$.
Using~\eqref{equ.meas.torsor}, we get that
\[{\boldsymbol\omega}_{{\boldsymbol{m}},p}(\pi_{\boldsymbol{m}}^{-1}(U)
\cap\mathcal D_{{\boldsymbol{m}},p})={\boldsymbol\omega}_{T_{\NS},p}(T_{\NS}({\mathbf Z}_p)){\boldsymbol\omega}_{H,v}(U)\]
for any open subset $U$ of $\pi_{\boldsymbol{m}}(\mathcal D_{{\boldsymbol{m}},p})$.
The maps $\log\circ H_F$ and $\log\circ H_E$ define
a map $\log_\infty:{\matheusm T}_{\boldsymbol{m}}({\mathbf R})\to\Pic(S)^\vee\otimes_{\mathbf Z}{\mathbf R}$
and using $\log_\infty\times \pi_{\boldsymbol{m}}$ we get a homeomorphism
\[{\matheusm T}_{\boldsymbol{m}}({\mathbf R})\to\Pic(S)^\vee\otimes_{\mathbf Z}{\mathbf R}\times\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf R})).\]
Let
\[T_{\NS}^1({\mathbf R})=\{\,t\inT_{\NS}({\mathbf R}),\ \forall\chi\in\Pic(S),
\vert\chi(t)\vert=1\,\}.\]
Then for any real number $B$ and any open
subset $U$ of $\pi_m(\mathcal D_{{\boldsymbol{m}},\infty}(B)$,
we get
\begin{align*}
&{\boldsymbol\omega}_{{\boldsymbol{m}},\infty}(\pi_{\boldsymbol{m}}^{-1}(U)\cap
\mathcal D_{{\boldsymbol{m}},\infty}(B))\\
&=\int_{\{\,y\in{C\eff}(S)^\vee,\ \langle\omega_S^{-1}
,y\rangle\leq \log(B)\,\}}e^{\langle\omega_S^{-1},y\rangle}\,\textinmath{d} y\times
\omega_{T_{\NS}}(T_{\NS}^1({\mathbf R}))\,\omega_{H,\infty}(U)\\
&=\alpha(S){\boldsymbol\omega}_{T_{\NS},\infty}(T_{\NS}^1({\mathbf R}))\,{\boldsymbol\omega}_{H,\infty}(U)
f(B),
\end{align*}
where ${C\eff}(S)^\vee$ is the dual to the closed cone in $\Pic(S)\otimes_{\mathbf Z}{\mathbf R}$
generated by the effective divisors.
Taking the product over all places of ${\mathbf Q}$, we get the formula
\begin{equation}
\label{equ.global.volume}
\begin{aligned}
{\boldsymbol\omega}_{{\boldsymbol{m}}}(\mathcal D_{{\boldsymbol{m}}}(B))&=
\alpha(S){\boldsymbol\omega}_{T_{\NS},\infty}(T_{\NS}^1({\mathbf R}))
{\boldsymbol\omega}_{H,\infty}(\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf R})))\int_0^{\log(B)}ue^u\,\textinmath{d} u\\
&\quad\times\left(\prod_{p\in{\mathcal P}}
L_p(1,\Pic(\overline S)){\boldsymbol\omega}_{T_{\NS},p}(T_{\NS}({\mathbf Z}_p))\right)\\
&\quad\times
\left(\prod_{p\in{\mathcal P}}L_p(1,\Pic(\overline S))^{-1}
{\boldsymbol\omega}_{H,p}(\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\mathbf Q}_p)))\right).
\end{aligned}
\end{equation}
By lemma~\ref{lemma.idelic}, the map from $T_{\NS}({\mathbf Q})$
to $\bigoplus_{p\in{\mathcal P}}X_*(T_{\NS})_p$ is surjective. It follows
that
\[T_{\NS}^1({\boldsymbol A}_{\mathbf Q})=(T_{\NS}^1({\mathbf R})\times\prod_{p\in{\mathcal P}}T_{\NS}({\mathbf Z}_p))
.T_{\NS}({\mathbf Q})\]
and we get an exact sequence
\[1\longrightarrowT_{\NS}({\mathbf Q})_\textinmath{tors}\longrightarrow
T_{\NS}^1({\mathbf R})\times\prod_{p\in{\mathcal P}}T_{\NS}({\mathbf Z}_p)
\longrightarrow T_{\NS}^1({\boldsymbol A}_{\mathbf Q})/T_{\NS}({\mathbf Q})\longrightarrow 1.\]
Combining this with formula~\eqref{equ.global.volume} and the definitions
of the adelic measures, we get the formula
\[{\boldsymbol\omega}_{{\boldsymbol{m}}}(\mathcal D_{{\boldsymbol{m}}}(B))=
\cardT_{\NS}({\mathbf Q})_\textinmath{tors}\alpha(S)\tau(T_{\NS})\,{\boldsymbol\omega}_H(\pi_m({\matheusm T}_{\boldsymbol{m}}({\boldsymbol A}_{\mathbf Q})))
\int_0^{\log(B)}ue^u\,\textinmath{d} u,\]
where $\tau(T_{\NS})$ denotes the Tamagawa number of $T_{\NS}$.
By Ono's main theorem~\cite[\S 5]{ono:tamagawa}, $\tau(T_{\NS})$ is equal to
$\sharp H^1({\mathbf Q},\Pic(\overline S)
/\sharp{\cyrille X}^1({\mathbf Q},T_{\NS})$ and using Salberger's
argument \cite[proof of lemma 6.17]{salberger:tamagawa}
and prop.~\ref{prop.uniquem},
any point in $S({\boldsymbol A}_{\mathbf Q})^{\Br}$ belongs to exactly $\sharp{\cyrille X}^1({\mathbf Q},T_{\NS})$
sets of the form $\pi_{\boldsymbol{m}}({\matheusm T}_{\boldsymbol{m}}({\boldsymbol A}_{\mathbf Q}))$.
This concludes the proof of the proposition.
\end{proof}
\let\bold\mathbf
\ifx\undefined\leavevmode\hbox to3em{\hrulefill}\,
\newcommand{\leavevmode\hbox to3em{\hrulefill}\,}{\leavevmode\hbox to3em{\hrulefill}\,}
\fi
\ifx\undefined$\hbox{n}^\circ$
\newcommand{$\hbox{n}^\circ$}{$\hbox{n}^\circ$}
\fi
\ifx\undefinedand
\newcommand{and }{and }
\fi
\ifx\undefined,
\newcommand{,}{,}
\fi
|
\section{Introduction}
Since the seminal paper by Pendry~\cite{pendry2000nrm}, subwavelength imaging at visible wavelengths has been demonstrated in much
thicker low-loss layered silver-dielectric periodic structures~\cite{wood2006,Scalora2007opex,Scalora08pra,Scalora:OE09,belov2005csi,belov2006prb,Li:prb07,Kotynski:jopa2009}. In fact, the effective medium theory (EMT) provides sufficient means to explain enhanced transmission through the metal-dielectric stacks. Subwavelength resolution results from the extreme anisotropy of the effective permittivity tensor~\cite{wood2006,belov2005csi,belov2006prb}.
Let us continue the introduction by linking the metal-dielectric multilayers for sub-wavelength imaging with the concepts taken from Fourier Optics.
We refer to the model of a linear shift invariant scalar system (LSI)~\cite{Saleh,GoodmanFourierOptics} for the description of optical multilayers in a situation when they act as imaging nano-elements for coherent monochromatic light. In-plane imaging through a layered structure consisting of uniform and isotropic materials represents a LSI, for either TE or TM polarisations. Linearity of the system is the consequence of linearity of materials and validity of the superposition principle for optical fields. Shift invariance results from the assumed infinite perpendicular size of the multilayer and the freedom in the choice of an optical axis.
A scalar description is valid for the TE and TM polarisations in 2D since all other field components may be derived from $E_y$ or $H_y$, respectively, where the co-ordinate system is oriented as in fig.~\ref{fig.schem}.
For the TM polarization, the magnetic field $H_y(x,z)$ may be represented with its spatial spectrum $\hat H_y(k_x,z)$
\begin{equation}
H_y(x,z)=\int_{-\infty}^{+\infty} \hat H_y(k_x,z) exp(\imath k_x x) dk_x,
\end{equation}
where, at least for lossless materials, the spatial spectrum is clearly separated into the propagating part $k_x^2<k_0^2 \epsilon$ and evanescent part $k_x^2>k_0^2 \epsilon$.
\begin{figure}
a.\includegraphics[width=5cm]{fig1a.eps}
b.\includegraphics[width=5cm]{fig1b.eps}
c.\includegraphics[width=5cm]{fig1c.eps}
\caption{Schematic of a periodic silver-dielectric multilayer with symmetric (\textbf{a}) or non-symmetric (\textbf{b}) composition, and the equivalent effective medium slab (\textbf{c}).
\label{fig.schem}}
\end{figure}
The transfer function $t(k_x)$ (TF) is the ratio of the output to incident fields spatial spectra and corresponds to the amplitude transmission coefficient of the multilayer
\begin{equation}
\hat H_y(k_x,z=L)=t(k_x) \cdot \hat H_y^{Inc}(k_x,z=0).
\label{eq.mtfdef}
\end{equation}
Due to reflections, the incident field $\hat H_y^{Inc}(k_x,z=0)$ differs from the total field $\hat H_y(k_x,z=0)$.
The point spread function (PSF) is the inverse Fourier transform of the TF and has the interpretation of the response of the system to a point signal $\delta(x)$. The response to an arbitrary input $H_y^{Inc}(x,z=0)$ can be further expressed as its convolution with the PSF
\begin{equation}
H_y(x,z=L)=H_y^{Inc}(x,z=0)* PSF(x)
\label{eq.psfconool}
\end{equation}
PSF of an imaging system usually provides clear information about the resolution, loss or enhancement of contrast, as well as the characteristics of image distortions. However, for subwavelength imaging, the PSF is not a straightforward measure of resolution and even imaging of objects smaller than the FWHM of PSF is possible~\cite{Kotynski:arXiv-09}.
In this paper we demonstrate a diffraction-free material for subwavelength sized optical beams. We combine the following properties of the multilayer: PSF with sub-wavelength size and little dependence on the thickness of the structure, high transmission, low losses, and a limited dependence of the imaging properties on the size of external layers. Together, these properties allow to use a multilayer as a flexible construction material for various optical imaging nano-devices.
\begin{figure}
a.\includegraphics[width=7.5cm]{fig2a.eps}
b.\includegraphics[width=8.5cm]{fig2b.eps}
\caption{\textbf{a} Intensity transmission and reflection coefficients ($T,R$) of the symmetric multilayer as a function of the filling factor, plotted over the transfer function (in vertical cross-sections of the color map). Phase isolines are separated by $\pi/4$. The pitch $\Lambda=d_1+d_2$ is fixed at $22nm$, and the total thickness is $L=30\Lambda$. \textbf{b.}~Characteristics of the symmetric and non-symmetric multilayers, and of the equivalent slab made of an effective medium, in the function of the number of periods. Intensity transmission and reflection coefficients calculated at normal incidence (top); FWHM of $PSF$ and $\arrowvert PSF\arrowvert^2$ (bottom).\label{fig.tf}}
\end{figure}
\psection{Imaging with sub-wavelength resolution in metal-dielectric multilayers}
The dispersion relation of a two-component infinite stack for the TM polarisation has the form~\cite{Wu:PRB67-235103}
\begin{equation}
\cos(K_z \Lambda)=\cos(k_{1z} f \Lambda) \cos(k_{2z} (1-f) \Lambda)
-{ \frac{\sin(k_{1z} f \Lambda) \sin(k_{2z} (1-f) \Lambda)}{2}}\left({\frac{k_{1z} \epsilon_2}{\epsilon_1 k_{2z}}} + {\frac{k_{2z} \epsilon_1}{\epsilon_2 k_{1z}}}\right)\label{eq.dispinf},
\end{equation}
where $K_z$ is the Bloch wavenumber, $\Lambda=d_1+d_2$ is the period of the stack, $d_i$ and $\epsilon_i$ are the layer thickness and permittivity of material $i=1,2$, the filing fraction of material $1$ is $f=d_1/\Lambda$ and the local dispersion relations are $k_{iz}^2+k_x^2=k_0^2 \epsilon_i$. Wavevector $k_x$ is conserved at the layer boundaries and depends on the incidence conditions. The Bloch wavenumber $K_z$ is real for a Bloch mode in a lossless stack, however for evanescent waves in a finite stack or for lossy materials $K_z$ may be complex.
In the first BZ, the real part of $K_z$ satisfies $\pi/\Lambda<=re(K_z)<=\pi/\Lambda$. The group velocity as well as the imaginary part responsible for absorption do not depend on the choice of BZ.
When the layers are thin $K_z \Lambda, k_{iz} \Lambda<<1$, the second order expansion of (\ref{eq.dispinf}) over the arguments of trigonometric functions leads to the dispersion relation for an uniaxially anisotropic effective medium
\begin{equation}
K_z^2/\epsilon_x+k_x^2/\epsilon_z=k_0^2,
\end{equation}
with $\epsilon_x = f \epsilon_1 + (1-f) \epsilon_2$ and $\epsilon_z =1/( f /\epsilon_1 + (1-f)/ \epsilon_2)$. This is the basis of the effective medium approximation thoroughly discussed by Wood et al.~\cite{wood2006}, also as the basis for applying the near-field approximation. The transmission coefficient of the Fabry-Perot (FP) slab consisting of the homogenized effective medium is then given as~\cite{wood2006},
\begin{equation}
t(k_x)=\left( \cos(k_z L)-0.5\imath (K_z\epsilon_x /k_z\epsilon+k_z\epsilon /K_z\epsilon_x )\sin(k_z L)) \right)^{-1},
\end{equation}
where $k_z$ and $\epsilon$ refer to the external medium, and $L$ is the total thickness of the slab. At the same time $t(k_x)$ is the already mentioned coherent amplitude transfer function of the imaging system~\cite{GoodmanFourierOptics}. When $\lvert \epsilon_x/\epsilon_z\rvert<<1$ (which is satisfied for a lossless metal for $\epsilon_1/\epsilon_2=-d_1/d_2$~\cite{belov2006prb}), for certain slab thickness the resonant FP condition becomes independent on the angle of incidence. This happens when $L=\lambda m/\sqrt{\epsilon_x}, m\in N$. Then $t(k_x)\equiv exp(- \imath K_z L)$ and the FP slab introduces the same phase shift for all harmonics of the spatial spectrum. Belov and Hao~\cite{belov2006prb} proposed to combine this condition with impedance matching between the external medium and the effective FP slab $\epsilon=\epsilon_x=f \epsilon_1 + (1-f) \epsilon_2$ and referred to that regime as canalization. However, Li et al.~\cite{Li:prb07} questioned the importance of impedance matching in favour of the FP resonance condition. In fact, for a lossless metal and dielectric, the FP resonance is sufficient to entirely eliminate reflections resulting in perfect imaging $t(k_x)\equiv 1$ without impedance matching.
In reality, losses make the condition $\epsilon=\epsilon_x$ only approximate and, at the same time, the finite value of $\Lambda$ limits the validity of homogenisation. Moreover, transmission through a finite slab strongly depends on the material and thickness of the external layers and appears to be the largest for a symmetrically designed slab~(Fig.~\ref{fig.schem}a) with half-width dielectric layers located at the boundaries~\cite{Scalora2007opex}.
\psection{Numerical demonstration of diffraction-free propagation of sub-wavelength sized optical beams}
After recalling the theory of sub-wavelength imaging in metal-dielectric multilayers, we now focus on an imaging regime which may be called diffraction-free. We note, that the FP resonances are accompanied with a field pattern inside the slab similar to a standing wave~\cite{wood2006,Kotynski:jopa2009}. Therefore look for structures for which $t(k_x)\approx const$ nonetheless FP resonances are weak.
Let us now focus on an example of a metal-dielectric periodic multilayer consisting of silver and high refractive index dielectric, which enables us to demonstrate and explain the diffraction-free propagation of sub-wavelength sized optical beams. The structure operates at the wavelength of $\lambda=422nm$, when the permittivity of silver is equal to $\epsilon_1=-5.637 + 0.214\imath$\cite{JohnsonChristy}, and the permittivity of the dielectric such as $TiO_2$ or $SrTiO_3$~\cite{Saleh,Palik} is $\epsilon_2=(2.6)^2$. Layer thicknesses are assumed to be $d_1=10nm$, $d_2=12nm$ with periodic symmetric or non-symmetric composition shown in fig.~\ref{fig.schem}a,b. The fraction $d_1/d_2$ may be justified with the use of EMT. The corresponding permittivity of the effective medium gives $\epsilon_x\approx1.02+0.1\imath$ and $\epsilon_z\approx-158+191\imath$, which assure impedance matching with air $\epsilon=1\approx\epsilon_x$ together with the condition for the extreme anisotropy $\lvert \epsilon_x/\epsilon_z\rvert<<1$. The imaginary part of the refractive index $n_x=\sqrt{\epsilon_x}$ is reduced by the factor of $50$ compared to bulk silver, resulting in low-loss transmission. In fig.\ref{fig.tf}a we show the transfer function calculated rigorously with TMM for a range of filling factors $0<d_1/\Lambda<0.6$, when the total thickness of the structure is fixed at $L=660nm$. Indeed, the impedance matching which occures when $d_1=10nm, d_2=12nm$ results in reduced reflections, high transmission of both propagating and evanescent spatial harmonics, and a flat phase of the transfer function for a broad range of $k_x/k_0$. The size of the corresponding PSF is of the order of $\lambda/10\approx 40nm$ assuring the super-resolving properties of the device. Fig.~\ref{fig.tf}b shows the FWHM of PSF
alongside with the reflection and transmission coefficient in the function of number of periods, for the symmetric, non-symmetric and effective-index composition of the structure. The major properties of the structure are the following: the size of PSF varies slowly with the size of the structure and is almost the same for the symmetric and non-symmetric composition, the FP resonances observed in transmission are weak (as opposed to those observed in reflections), the attenuation is uniform as the number of layers is increased. These properties assure an almost diffraction-free propagation through the structure, with a similar attenuation and the size of PSF for symmetric and non-symmetric composition and for a broad range of structure sizes $L$. This behaviour is illustrated in fig.~\ref{fig.fdtd1}a with an FDTD~\cite{Farjadpour2006ol} simulation showing the uniform non-diverging distribution of the Poynting vector inside the structure for a beam size of the order of $\lambda/10$. Transmission takes place along the direction normal to the layer boundaries with negligible divergence. Furthermore, our material may be used to fabricate slabs cut at arbitrary angle to layer surfaces. As is shown in fig.~\ref{fig.fdtd1}b,c we continue to observe the diffraction-free propagation for inclined layers, whether the aperture covers only a single or multiple periods of the grating. This may be heuristically explained with the similar PSF for a symmetric and non-symmetric composition of multilayers. Both of them are encountered at the cross sections drawn along the propagating beam, starting from various points inside the aperture and normal to the layer boundaries. Furthermore, it is possible to construct other simple optical nano-elements. In fig.~\ref{fig.fdtd2} we demonstrate the operation of a double multilayer, which may be part of a cloaking device or an optical interconnect, as well as prism for imaging subwavelength beams.
\begin{figure}
a.\includegraphics[height=4.5cm]{fig3a.eps}
b.\includegraphics[height=5cm]{fig3b.eps}
c.\includegraphics[height=5cm]{fig3c.eps}
\caption{Diffraction-free transmission of the subwavelength sized beam through the multilayer (FDTD simulations, with the incident CW beam limited by a subwavelength aperture in a perfect conductor): \textbf{a}.~time-averaged Poynting vector $S_z$ in a rectangular slab; \textbf{b,c}.~instantaneous magnetic field $H_y$ in a slab with layer boundaries oriented at $30^\circ$ towards the external boundaries for various aperture sizes.\label{fig.fdtd1}}
\end{figure}
\begin{figure}
a.\includegraphics[height=5cm]{fig4a.eps}
b.\includegraphics[height=5cm]{fig4b.eps}
\caption{Simple optical nano-elements internally made of the diffraction-free material \textbf{a}. double slab; \textbf{b}. prism;.
\label{fig.fdtd2}}
\end{figure}
\psection{Conclusions}
We have demonstrated a diffraction-free, low-loss material, which is impedance matched to air. The diffraction-free propagation is verified by the analysis of FWHM of PSF in the function of the number of periods. The material consists of silver and a dielectric with refractive index of $n=2.6$ and operates at $\lambda=422nm$. It has the effective imaginary part of refractive index smaller than that of silver by a factor of $50$. The reflections are in between $-10dB$ and $-30dB$, and the FWHM of PSF is at the order of $40nm$. Small reflections, small attenuation, and reduced Fabry Perot resonances make it a flexible metamaterial for arbitrarily shaped optical planar devices with sizes of the order of one wavelength, such as the elements of optical interconnects or cloaks.
\ack
We acknowledge support from the Polish MNiI research projects \textit{N N202 033237} and \textit{N R15 0018 06}
as well as the framework of COST actions \textit{MP0702} and \textit{MP0803}.
|
\section{Introduction and Statement of Results}
For any smooth area form $\mu$ on $S^2$, let $\Diff_{\mu}(S^2)$ be the
group of $C^\infty$ diffeomorphisms of $S^2$ that preserve $\mu$. In particular
they preserve the orientation of $S^2.$ Elements of $\Diff_{\mu}(S^2)$
are said to {\em preserve area.} Surface diffeomorphisms with
positive entropy have been studied from both the hyperbolic dynamical
systems point of view and the Nielsen-Thurston point of view. In this
paper we formulate and prove a structure theorem for $F\in
\Diff_{\mu}(S^2)$ with zero entropy. The assumptions that $S = S^2$
and that $F$ preserves area are made to simplify the problem.
Suppose that $F \in \Diff_{\mu}(S^2)$ has zero topological entropy and
that $\Fix(F)$ is the set of fixed points for $F$. To enhance the
topology of the ambient surface, we consider ${\cal M} = S^2 \setminus
\Fix(F)$ and $f=F|_{{\cal M}} :{\cal M} \to {\cal M}$.
Every $x \in {\cal M}$ has a neighborhood $B$ that is a {\em free disk},
meaning that $B$ is an open disk and that $f(B) \cap B =\emptyset$. A
very weak notion of recurrence for a point $x \in {\cal M}$ is to require
that there be $n \ne 0$ and a free disk $B$ that contains both $x$ and
$f^n(x).$ Note that this is equivalent to saying there is a free disk
$B$ which intersects $\orb(x),$ the orbit of $x$, in at least two
points. We will call such points {\em free disk recurrent}
and denote the set of these points by ${\cal W}_0.$ Clearly, if either
$\alpha(F,x)$ or $\omega(F,x)$ contains a point which is not
in $\Fix(F)$ then $x \in {\cal W}_0.$ In
particular the set ${\cal W}_0$ contains the full measure subset of ${\cal M}$
consisting of birecurrent
points. The set ${\cal W}_0$ is open and dense in ${\cal M}$. However,
its components may not be equal to the interior of their
closure, and this is a property which is technically useful.
Consequently we define the larger set ${\cal W}$ of {\em weakly free disk
recurrent} points as follows.
\begin{defn}\label{defn: omega-free}
A point $x \in {\cal M}$ is {\em free disk recurrent} for $f$
provided there is an $f$-free disk $B$ in ${\cal M}$, such that $\orb(x),$
the orbit of $x$ intersects $B$ in at least two points.
The set of free disk recurrent points in ${\cal M}$ is denoted
${\cal W}_0.$ If $W_0$ is a component
of ${\cal W}_0$ and $x \in {\cal M}$ is in $\Int(\cl_{{\cal M}}(W_0))$, then we say that $x$ is
{\em weakly free disk recurrent}. The set of weakly free disk
recurrent points in ${\cal M}$ is denoted ${\cal W}$.
\end{defn}
There are several immediate consequences of the definition which
we record for later use.
\begin{rem}\label{rem: free disk}
The set ${\cal W} \subset {\cal M}$ is open, dense, and contains all birecurrent
points for $f$. The complement of ${\cal W}$
is a closed
set with measure zero. If any point of $\omega(x,F)$ or $\alpha(x,F)$
is not in $\Fix(F)$ then a free disk containing that point will
contain infinitely many points of $\orb(x)$ so $x$ is free disk
recurrent.
\end{rem}
There are certain elements of $\Diff_\mu(S^2)$ which are trivial from
the point of view of their periodic points.
These include $F \in \Diff_\mu(S^2)$ of
finite order and $F$ for which $\Per(F)$ contains only two points. It
is known that an area preserving $F$ must have at least two fixed
points (see \cite{simon:index}). In
the case that $\Per(F)$ contains exactly two points those points must
be fixed since $F$ must have a fixed point
and the complement of
those points may be compactified to form a closed annulus and $f$
extends to this closed annulus with every point of that
annulus having the same
irrational rotation number. This is an interesting topic to
investigate but is not addressed in this article.
The main building block in our structure theorem is a partition of
${\cal W}$ into countably many disjoint $f$-invariant annuli.
\begin{thm} \label{thm: annuli} Suppose $F \in \Diff_\mu(S^2)$ has
entropy zero. Let $f = F|_{{\cal M}}$ where ${\cal M} = S^2
\setminus \Fix(F)$. Then there is a countable collection ${\cal A}$ of pairwise disjoint
open $f$-invariant annuli such that
\begin{enumerate}
\item The union $\bigcup_{U \in {\cal A}} U$ is the the set ${\cal W}$ of weakly free disk
recurrent points for $f$.
\item If $U \in {\cal A}$ and $z$ is in the frontier of $U$ in $S^2$, there
are components $F_+(z)$ and $F_-(z)$ of $\Fix(F)$ so that $\omega(F,z)
\subset F_+(z) $ and $\alpha(F,z) \subset F_-(z)$.
\item For each $U \in {\cal A}$ and each component $C_{{\cal M}}$ of the frontier of
$U$ in ${\cal M}$, $F_+(z)$ and $F_-(z)$ are independent of the choice of $z
\in C_{{\cal M}}$.
\end{enumerate}
\end{thm}
\begin{remark}\label{first centralizer invariance} If $h:S^2 \to S^2$ commutes with $F$ then it permutes the open annuli in the family
${\cal A}.$ This is because each $U \in {\cal A}$ is a
component of ${\cal W}$ and $h$ is a conjugacy from $F$ to itself and hence
preserves ${\cal W}$.
\end{remark}
To see how the elements of ${\cal A}$ arise, consider the special case that
$F$ is the time one map of an area preserving flow $\phi_t$. Given $x
\in {\cal M},$ choose a free disk neighborhood $B$ of $x$ which is also a
flow box for $\phi_t$. It is an easy consequence of the
Poincare-Bendixson theorem that if the flow line for $\phi_t$ that
contains $x$ returns to $B$ it closes up into a simple closed curve
$\rho_x$. In particular, in this case the subsets ${\cal W}_0$ and ${\cal W}$ are
equal and coincide with the union of the periodic orbits of the flow
which lie in ${\cal M}$. Denote {\em the isotopy class of $\rho_x$} in ${\cal M}$
by $[\rho_x]$. It is clear that $\rho_x$ depends only on the orbit of
$x$ and not $x$ itself and that if $z \in B$ is sufficiently close to
$x$ then $\rho_x$ and $\rho_z$ cobound an annulus in $M$; in
particular $[\rho_x] =[ \rho_z]$. In this case $U=\{ y \in {\cal W}:
[\rho_y] =[ \rho_x]\}$ is the element of ${\cal A}$ that contains $x$.
For a second special case suppose that $f$ is isotopic to the
identity. Given $x \in {\cal W}_0$, choose $B$ and $n$ as in the definition of
weakly free disk recurrent. If $f_t:{\cal M} \to {\cal M}$ is an isotopy between $f_0 = $
identity and $f_1 = f$ then the path $\mu_x \subset {\cal M}$ defined by
$\mu_x(t) = f_t(x)$ connects $x$ to $f(x)$. The path $\mu_x \cdot
\mu_{f(x)} \cdot \ldots \cdot \mu_{f^{n-1}(x)}$ can be closed by
adding a path in $B$ connecting $f^n(x)$ to $x$. Up to homotopy in
${\cal M}$, this closed path is a multiple of some non-repeating closed path
$\rho_x$. Using the hypothesis that $F$ has entropy zero, one can
show (see \cite{fh:periodic}) that the homotopy class of $\rho_x$ is
represented by a simple closed curve, (also written $\rho_x$) that is
independent of $B, n$ and the choice of isotopy $f_t$. It is easy to see that if
$z \in B$ is sufficiently close to $x$ then $[\rho_x] =[ \rho_z]$. As
in the previous case, $U=\{ y \in {\cal W}: [\rho_y] =[ \rho_x]\}$ is the element of ${\cal A}$ that contains $x$.
In the general case, we make use of the fact (Theorem 1.2, Lemma~6.3 and Remark~6.4, all in \cite{fh:periodic}) that $f$ is isotopic
to a composition of Dehn twists along a finite set of simple closed
curves ${\cal R}$.
Cutting along the elements of ${\cal R}$ produces a
decomposition of ${\cal M}$ into subsurfaces ${\cal M}_i$ such that $f|_{{\cal M}_i} : {\cal M}_i
\to {\cal M}$ is isotopic to the inclusion ${\cal M}_i\hookrightarrow {\cal M}$. The main
technical work in this proof is showing that each ${\cal M}_i$ is realized,
in a suitable sense, by an $f$-invariant subsurface; see
section~\ref{sec:domain covers}. One then defines ${\cal A}$
in a fashion similar to the second special case.
Theorem~\ref{thm: annuli} can be applied to $F^q$ for each $q\ge 2$.
This gives a countable collection ${\cal A}(q)$ of pairwise disjoiont open $F^q$-invariant annuli that (see Proposition~\ref{prop: V in U}) refines ${\cal A}$ in the sense that each $V_j \in {\cal A}(q)$ is contained in some $U_i \in {\cal A}$. This {\em renormalization process}
can be iterated with ${\cal A}(q)$ playing the role of ${\cal A}$ and
so on. The $V_j$'s may be essential or inessential in $U_i$. In the
limit, the former lead to twist-map-like behavior and the latter to
solenoid-like behavior when they are nested infinitely often.
It is important to note that replacing
$F$ with $F^q$ changes the set of fixed points and hence changes ${\cal M}$
and changes the free disk recurrent points of ${\cal M}$.
We are interested in partitioning $cl(U)$ into sets
analogous to the periodic orbits in the case of the time one map of a
flow. In particular we would like the rotation number to be constant
on these sets. The two components of the frontier of $U$ can be
somewhat problematic since such a component could be a single point or
could be a complicated fractal. To deal with this issue we introduce a
natural compactification $U_c$ of $U$ and a natural extension $f_c$ of
$f$ (see Definition~(\ref{defn: annular comp}) in the next
section ). The compactification of an end described there is either the
prime end compactification or the compactification obtained by blowing
up a fixed point, whichever is appropriate. The set of two rotation
numbers of $f_c$ on the two circles added in the compactification will
be denoted by $\rho(\partial U_c).$
We are now prepared to state the second of our main results It
describes the finer structure of the dynamics of $f$ on one of the
annuli in ${\cal A}$. The proof is based on renormalization and the details
are in Section~\ref{sec: renormalization}.
\begin{thm}\label{thm: C(x)}
Suppose $F \in \Diff_\mu(S^2)$ has entropy zero. Let $f = F|_{{\cal M}}$ where
${\cal M} = S^2 \setminus \Fix(F)$ and let ${\cal A}$ be as in Theorem~\ref{thm: annuli}.
Then for each $U \in {\cal A}$ there is a partition of $cl(U)\subset S^2$ into a family ${\cal C}$
of closed $F$-invariant sets
with the following properties:
\begin{itemize}
\item Each $C \in {\cal C}$ is compact and connected.
\item There are two elements of ${\cal C}$ (which may coincide), called {\em ends}
and denoted $C_0$ and $C_1,$
each of which contains a component of the frontier of $U$.
Every element of ${\cal C}$ which is not an end
is a subset of $U$ and is called {\em interior}.
Each interior $C$ is essential in $U$, i.e.
its complement (in $U$) has two components and it separates $C_0$ and $C_1$.
\item The rotation number $\rho_f(x) \in \mathbb R/\mathbb Z$ is well defined and
constant on any interior $C$. In fact, each $C$ is a connected
component of a level set of $\rho_f(x).$ Moreover, $\rho_f(x)$ is
is continuous on $U$
and has a unique continuous
extension to $cl(U)$. The value of this extension on $C_i, \ i=0,1$ is
the element of $\rho(\partial U_c)$ corresponding to the component of
the frontier of $U$ contained in $C_i.$
\end{itemize}
\end{thm}
We emphasize the fact that the $C$ containing a
point $x$ may be very different for $F^q$ than for $F$, and, in fact,
will be undefined for $F^q$ if $x$ is a periodic point whose period
divides $q.$
\begin{remark}
There are some degenerate cases where both Theorem~(\ref{thm: annuli})
and Theorem~(\ref{thm: C(x)}) remain true, but are not particularly
interesting. For example, if $F$ is the identity then ${\cal M}$ and ${\cal A}$
are empty and the results are vacuously true. Slightly more interesting
is the case when $\Fix(F)$ contains only two
points and all points of the annulus $U = {\cal M} = S^2 \setminus \Fix(F)$
have the same rotation number, (e.g. if $F$ is an isometry of some
smooth metric).
Then $U = {\cal M}$ is an
$f$-invariant annulus. In this case ${\cal A}$ contains the single annulus
$U = {\cal M}$ and (as we show in Proposition~(\ref{prop: free-disk A}))
every point of $U$ is free disk recurrent. Also in this case
Theorem~(\ref{thm: C(x)}) remains true since the two ``ends'' $C_0$
and $C_1$ coincide, there are no interior $C$'s, and $C_0 = C_1 =
S^2.$
\end{remark}
The sets in ${\cal C}$ are the generalizations of the closed orbits
foliating $U$ in the special case that $F$ is the time one map of a
flow. Of course in the general case $C \in {\cal C}$ can be considerably more
complicated. The main example constructed in \cite{han:pseudo-circle}
shows $C$ can be a pseudo-circle. It is also possible for $C$
to have interior.
A heuristic picture of one possibility
in the case that $\rho_f|_C$ is rational is an essential ``necklace''
in $U$ consisting of a periodic orbit of saddle periodic points
each joined to the next by a stable manifold (which is the unstable
manifold of the next) and by an unstable manifold (which is the stable
manifold of the next). This pair, stable and unstable, bound a ``bead'',
an open disk. The diffeomorphism $f$ permutes the beads and has
a periodic orbit with one point in each bead. The set $C$ containing any
$x$ in one of the beads will be the entire necklace. For such a $C$
there is an $n$ such that $f^n$ will fix each bead and each saddle
point joining them.
We now turn to applications.
Recall that a group $G$ is a {\em indicable} if there exists a
non-trivial homomorphism $G \to \mathbb Z$. For finitely generated groups
this is equivalent to $H^1(G, \mathbb Z) \ne 0$ and equivalent to the
abelianization of $G$ being infinite. If a finite index subgroup of
$G$ is indicable then we say that $G$ is {\em virtually indicable.}
Denote the centralizer of $f \in \Diff_{\mu}(S^2)$ by $Z(f)$. As an application of Theorem~\ref{thm: annuli} and Theorem~\ref{thm: C(x)} we prove
\begin{thm}\label{thm: ent0}
If $f \in \Diff_{\mu}(S^2)$ has infinite order then each finitely
generated infinite subgroup $H$ of $Z(f)$ is virtually indicable.
\end{thm}
One might expect that Theorem~\ref{thm: ent0} is proved by first
proving the existence of a finite index subgroup $H_0$ of $H$ with
global fixed points and then applying the Thurston stability theorem(
\cite{thurston:stability}; see also Theorem 3.4 of
\cite{franks:distortion}) to produce a non-trivial homomorphism from
$H_0$ to $\mathbb Z$. This is easy to do (see Proposition~(\ref{prop: +ent}))
in the case that $f$ has
positive entropy but fails when $f$ has zero entropy. Indeed, there
are examples (see (\ref{example})) for which no finite index subgroup of $Z(f)$ has a
global fixed point. We prove Theorem~\ref{thm: ent0} by analyzing the
possible ways in which the existence of global fixed points can fail
and by showing that each allows one to define a non-trivial
homomorphism to $\mathbb Z$.
As an application of Theorem~\ref{thm: ent0} we have the following
result about mapping class groups.
\begin{cor} \label{no actions} If $\Sigma_g$ is the closed orientable surface of genus $g \ge 2$ then at least one of the following holds.
\begin{enumerate}
\item \label{no faithful actions} No finite index subgroup of $\mcg(\Sigma_g)$ acts faithfully on $S^2$ by area preserving diffeomorphisms.
\item \label{non-trivial first cohomology}For all \ $1 \le k \le g-1$, there is an indicable finite index subgroup $\Gamma$ of the bounded mapping class group $\mcg(S_k,\partial S_k)$ where $S_k$ is the surface with genus $k$ and connected non-empty boundary.
\end{enumerate}
\end{cor}
Corollary~\ref{no actions} relates to the following well known questions about mapping class groups.
\begin{question} \label{mcg question} Does $\mcg(\Sigma_g)$, or any of its finite index subgroups, act faithfully on a closed surface $S$ by diffeomorphisms? by area preserving diffeomorphisms?
\end{question}
\begin{question}[{\bf Ivanov}] \label{cohomology} Does every finite index subgroup $\Gamma$ of $\mcg(\Sigma_g)$ satisfy $H^1(\Gamma, \mathbb R) = 0$?
\end{question}
Question~\ref{mcg question} is motivated in part by the sections problem (see Problem~6.5 and Question~6.7) of Farb's survey/problem list \cite{farb:survey} on the mapping class group: which subgroups of $\mcg(\Sigma_g)$ lift to $\Diff(\Sigma_g)$? It is also motivated by the analogy between mapping class groups and higher rank lattices and the fact (\cite{Po},\cite{fh:periodic}, \cite{fh:distortion}) that every action of a non-uniform irreducible higher rank lattice on $\Sigma_g$ by area preserving diffeomorphisms factors through a finite group; see Question~12.4 of Fisher's survey article \cite{fisher:survey} on the Zimmer program.
Question ~\ref{cohomology} is Problem 2.11 of \cite{kirby};
see also \cite{Iv2} and \cite{kork2}.
Corollary~\ref{no actions}-\pref{no faithful actions} is a negative
answer to the area preserving, $S = S^2$ case of Question~\ref{mcg
question}. The answer to Question~\ref{cohomology} is no for genus
$2$ (see \cite{mcC}) but is unknown for for genus
at least three. Presumably a positive answer to
Question~\ref{cohomology} would imply that Corollary~\ref{no
actions}-\pref{non-trivial first cohomology} does not hold and so
imply that Corollary~\ref{no actions}-\pref{no faithful actions} does
hold.
\section{Area preserving annulus maps.} \label{sec: annulus maps}
We will make use of a number of results on area preserving
homeomorphisms and diffeomorphisms of the annulus which we cite here.
If $A = S^1 \times [0,1]$ is the annulus its universal
covering space is $\tilde A = \mathbb R \times [0,1].$ We will
denote by $p_1$ \ the projection, $p_1 : \mathbb R \times [0,1] \to \mathbb R$,
of $\tilde A$ onto its first factor.
\begin{defn}
If $f: A \to A$ is an orientation preserving homeomorphism
isotopic to the identity and $\tilde f$ is a lift to $\tilde A$
then the {\em translation number} $\tau_{\tilde f}(\tilde x)$
of $\tilde x \in \tilde A$ is defined (if it exists)
to be
\[
\lim_{n \to \infty} \frac{p_1(\tilde f^n(\tilde x)) - p_1(\tilde x)}{n}.
\]
The {\em rotation number} $\rho_{f}(x)$ of
$x \in A$ is defined to be the projection of
$\tau_{\tilde f}(\tilde x)$ in $\mathbb T^1 = \mathbb R / \mathbb Z$ for
any lift $\tilde f$ of $f$ and any lift $\tilde x$
of $x.$
\end{defn}
There are several results about these numbers which we will
need.
\begin{thm}\label{thm: rot num}
Suppose $f: A \to A$ is an area preserving homeomorphism
of the closed annulus which is isotopic to the identity
and let $\tilde f: \tilde A \to \tilde A$ be a lift to its universal
covering space. Then $\tau_{\tilde f}(\tilde x)$ exists for almost
all $\tilde x \in \tilde A.$ Moreover if there exist
points $\tilde x, \tilde y \in \tilde {\cal A}$ with $\tau_{\tilde f}(\tilde x) =s$
and $\tau_{\tilde f}(\tilde y) =t$ then for any
rational in lowest terms $p/q \in [s, t]$ there is a point
$\tilde z \in \tilde A$ such
that $\tau_{\tilde f}(\tilde z) = p/q$ and $\tilde z$ is the lift of a periodic
point for $f$ with period $q.$
\end{thm}
\begin{proof} The fact that $\tau_{\tilde f}(\tilde x)$ exists for almost
all $\tilde x$ is a standard consequence of the Birkhoff ergodic
theorem applied to the function $\phi(x) = p_1(\tilde f(\tilde x)) - p_1(\tilde x)).$
The remainder of the result is Corollary~2.4
of \cite{franks:recurrence}.
\end{proof}
\begin{defn}
If $f: A \to A$ is an area preserving homeomorphism
isotopic to the identity and $\tilde f$ is a lift to $\tilde A,$
then the {\em translation interval} $\tau(\tilde f)$ of $\tilde f$
is the smallest closed interval that contains $\{\tau_{\tilde f}(\tilde x): \tilde x \in \tilde A\}$.
The {\em rotation interval} $\rho(f)$ of
$f$ is defined to be the projection of
$\tau(\tilde f)$ in $\mathbb T^1 = \mathbb R / \mathbb Z$ for
any lift $\tilde f$ of $f.$
\end{defn}
The terminology of the previous definition is appropriate
in light of the following theorem.
\begin{thm}[\cite{handel:rotation}]\label{thm: irr rot num}
Suppose $f: A \to A$ is an area preserving homeomorphism
of the closed annulus which is isotopic to the identity
and let $\tilde f: \tilde A \to \tilde A$ be a lift to its universal
covering space. Then the set $I$ of $r \in \mathbb R$ such that there is
$\tilde x \in \tilde A$ with $\tau_{\tilde f}(\tilde x) = r$ is closed
and hence $I$ is the closed interval $\tau({\tilde f}).$
\end{thm}
The following result can be found as Theorem (3.3) of
\cite{franks:poincare}.
\begin{prop}[\cite{franks:poincare}]\label{prop: limsup}
Suppose $f: A \to A$ is an area preserving homeomorphism
of the closed annulus which is isotopic to the identity
and let $\tilde f: \tilde A \to \tilde A$ be a lift to its universal
covering space and $\tau(\tilde f) = [r,s]$ its translation interval.
Then for all $\tilde x \in \tilde A$
\[
r \le \liminf_{n \to \infty} \frac{p_1(\tilde f^n(\tilde x)) - p_1(\tilde x)}{n}
\le \limsup_{n \to \infty} \frac{p_1(\tilde f^n(\tilde x)) - p_1(\tilde x)}{n}
\le s.
\]
\end{prop}
\begin{prop}\label{prop: +meas}
Suppose $f: A \to A$ is an area preserving homeomorphism
of the closed annulus which is isotopic to the identity.
If there is a subset $Y \subset A$ with Lebesgue
measure $\mu(Y) > 0$ and such that $\rho_f(x) = 0$ for almost all
$x \in Y$ then $f$ has a fixed point in the interior of $A$.
\end{prop}
\begin{proof}
Since $\mu(Y) >0$ there is a small open disk $D$
whose closure is in the interior of $A$
with $\mu( Y \cap D) >0.$ If $f$ has no fixed point in
$D$ then by making $D$ smaller we may assume it is a free disk.
We let $X = Y \cap D$.
Let $r: X \to X$ be the first return map so $r(x) = f^n(x)$ where
$n$ is the smallest positive integer such that $f^n(x) \in X$
The function $r$ is well defined for almost all $x \in X$, so
deleting a set of measure $0$ from $X$ we may assume it defined
for all $x \in X.$
Let $\tilde D$ be a lift of $D$. If $\tilde X$ is the set of lifts
to $\tilde D$ of points in $X$ then there is a positive measure
subset $X_0 \subset X$ and a lift $\tilde f$ of $f$ such that
$\tau( \tilde x, \tilde f) = 0$ for all $\tilde x \in \tilde X_0,$ where
$\tilde X_0$ is the lift of $X_0$ to $\tilde D.$
Suppose the first return time for $x$ is $n$, so $r(x) = f^n(x)$.
Then $\tilde f^n(\tilde x) \in T^k(\tilde D)$ for a unique integer $k.$
We define $h(x, \tilde f),$ the {\em homological displacement} of $x$, to
be $k$. It depends on $\tilde f$ but not on the choice
of lift $\tilde D$ of $D.$
It suffices to prove that $h(x, \tilde f) = 0$ for some $x \in X_0$
because then $\tilde x$ is contained in a periodic disk chain
(see Proposition~(1.3) of \cite{franks:poincare}) and
$\tilde f$ has a fixed point. We note that
if there are $x,y \in X_0$ such that $h(x, \tilde f) > 0$
and $h(y, \tilde f) < 0$ then $\tilde f$ has a fixed point.
This is a consequence of Theorem~(2.1) of \cite{franks:poincare}
since there are both
positive and negative recurring disk chains for $f$.
Hence we may assume $h(x, \tilde f)$ has a constant sign.
A result of \cite{franks:open_annulus} shows that if
\[
B = \bigcup_{n \in \mathbb Z} f^n( X_0)
\]
then
\[
\int_{X_0} h(x, \tilde f) \ d\mu = \int_B \tau(x, \tilde f) \ d\mu.
\]
Since $\tau(x, \tilde f) = 0$ for all $x \in X_0$ we conclude
that $\int_{X_0} h(x, \tilde f) \ d\mu = 0.$ Since $h$ has constant
sign it follows that $h(x, \tilde f) = 0$ for almost all $x \in X_0.$
\end{proof}
Suppose $U \subset S^2$ is an open $f$-invariant annulus. We would
like to compactify $U$ to a closed annulus for which $f$ has a natural
extension. The annulus $U$ has two ends which we compactify
separately in a way depending on the nature of the end. We say that an
end of $U$ is {\em singular} if the complementary component of $U$
that it determines is a single, necessarily fixed, point $x \in S^2$.
In this case we compactify that end by blowing up $x$ to obtain a
circle on which $f$ acts by the projectivization of $Df_x$. If the
end is not singular we will take the prime end compactification (see
Mather \cite{mather:prime_end} for properties). In either case we
obtain a closed annulus $U_c$ whose interior is naturally identified
with $U$ in such a way that $f|_{U}$ extends to a homeomorphism $f_c:
U_c \to U_c.$
\begin{notn}\label{defn: annular comp}
We will call $U_c$ the {\em annular compactification}
of $U$ and $f_c: U_c \to U_c$ the annular compactification
of $f|_U.$ If there is no ambiguity about the choice of
$f$ we will denote $\rho(f_c)$ by $\rho(U)$ and the two
rotation numbers of the restriction of $f_c$ to its boundary
circles by $\rho(\partial U_c).$
\end{notn}
\begin{lemma}\label{lem: frontier fp}
Let $f$ be an area and orientation preserving diffeomorphism
of either $S^2$ or the closed disk $D^2.$
Suppose $U$ is an open $f$-invariant annulus
and $f_c : U_c \to U_c$ is the extension of $f$ to its
annular compactification. If there is a point $x \in U_c$
with $\rho_{f_c}(x) = 0$ then there is a fixed point $\bar x$ of
$f$ in $cl(U).$ If $x$ lies in the component $X$ of $\partial
U_c$ corresponding to a component $\bar X$ of the frontier of $U$
then $\bar x \in \bar X.$
\end{lemma}
\begin{proof}
The function $f_c$ has a fixed point by Theorem~(\ref{thm: rot num}).
If it is in $\Int(U_c) = U$ we
are done. So we may assume it is in a boundary component $X$ of
$U_c$. If $X$ corresponds to a singular end of $U$ then the fixed
point corresponding to that end is $\bar X \subset cl(U).$ Otherwise $X$ is the
prime end compactification of an end of $U$ and each prime
end $x \in X$ is defined by a sequence of ``cross-cuts'' $\{\gamma_n\}$ where
each $\gamma_n$ is a Jordan arc whose interior is in $U$ and whose
endpoints are in the frontier of $U$. They satisfy
\begin{enumerate}
\item The limit $\displaystyle \lim_{n \to \infty} \diam(\gamma_n) = 0.$
\item Each $\gamma_n$ has two complementary components in $U$,
one of which is an annulus and the other of which is an open
disk which we will denote $D_n$.
\item The disk $D_{n+1}$ is a subset of $D_n$ and
$\displaystyle \bigcap_n D_n = \emptyset.$
\end{enumerate}
Two such sequences of cross-cuts $\{\gamma_n\}$ and
$\{\gamma_m'\}$ determine the same prime end
if for each $n$ there is an $m$ with $D_m' \subset D_n$ and
for each $m$ an $n$ with $D_n \subset D_m'$.
Let $\{\gamma_n\}$ determine a prime end in $X$ which is fixed by $f_c$.
Then from the fact that $f$ preserves area it follows that
$f(\gamma_n) \cap \gamma_n \ne \emptyset.$ For $n\ge 1$ choose $x_n
\in \Int(\gamma_n).$ From the properties above it follows that any
point in the limit set of the sequence $\{x_n\}$ is a fixed point of
$f$. It is clearly in $\bar X.$
\end{proof}
\begin{cor}\label{cor: frontier gfp}
Let ${\cal G}$ be a group of area and orientation preserving
diffeomorphisms of $S^2$ or the closed disk $D^2.$
Suppose $U$ is an open ${\cal G}$-invariant annulus
and ${\cal G}_c$ is the group of homeomorphisms of $g_c: U_c \to U_c$,
for $g \in {\cal G}.$ If there is a point $x \in \Fix({\cal G}_c)$
then $cl(U)$ contains a point $\bar x$ of $\Fix({\cal G}).$
If $x$ lies in the component $X$ of $\partial
U_c$ corresponding to a component $\bar X$ of the frontier of $U$
then $\bar x \in \bar X.$
\end{cor}
\begin{proof} If $x$ is a point of $\Fix({\cal G}_c)$ and $x \in \Int(U_c) =
U$ we are done. So we may assume it is in a boundary component $X$ of
$U_c$. If $X$ corresponds to a singular end of $U$ then the point
corresponding to that end is in $cl(U) \cap \Fix({\cal G}).$ Otherwise $X$
is the prime end compactification of an end of $U.$ Let $\{\gamma_n\}$
be a sequence of cross-cuts that determine a prime end in $X$ which is
in $\Fix({\cal G}_c).$ Then, as in the previous lemma, the fact that each $g
\in {\cal G}$ preserves area implies that $g(\gamma_n) \cap \gamma_n \ne
\emptyset.$ For $n\ge 1$ choose $x_n \in \Int(\gamma_n).$ From the
properties above it follows that any point in the limit set of the
sequence $\{x_n\}$ is in $\Fix(g)$. Since this is independent of the
choice of $g \in {\cal G}$ it follows that any point in the limit set is in
$\Fix({\cal G}).$
\end{proof}
\begin{prop}\label{prop: free-disk A}
Suppose $f: A \to A$ is an area preserving homeomorphism
of the closed annulus which is isotopic to the identity
and suppose every point of $A$ has the same rotation number. Let
$U = int(A).$ Then either $f$ has a fixed point in $U$ or
every point of $U$ is free disk recurrent for $f|_U.$
\end{prop}
\begin{proof}
If the rotation number of all points of $A$ is $0,$ then
Proposition~(\ref{prop: +meas}) implies that $f$ has a fixed point in $U$.
Hence we may assume the common rotation number of
the points of $A$ is non-zero and consequently $\Fix(f) = \emptyset.$
Suppose $x \in U$ and $z \in \omega(x) \subset A.$ If
$z \in U,$ then any free disk containing $z$ intersects $\orb(x)$
in infinitely many points. If $z \in \partial A$ let $V$
be a free half disk neighborhood of $z$ in $A$ and let $V_0 = V \cap U.$
Then $\orb(x)$ intersects $V_0$ infinitely often.
\end{proof}
\section{Planar topology} \label{sec:planar}
In this section we record and prove two useful elementary results.
Recall that by the Riemann mapping theorem, every open, unbounded,
connected, simply connected subset of $\mathbb R^2$ is homeomorphic to
$\mathbb R^2$. A closed set $X \subset \mathbb R^2$ is said to separate two
subsets $A$ and $B$ of $\mathbb R^2$ provided $A$ and $B$ are contained
in different components of $\mathbb R^2 \setminus X.$
\begin{lemma} \label{regular point} If $A$ and $B$ are disjoint
closed connected subsets of $\mathbb R^2$ then they are separated
by a simple closed curve or a properly embedded line.
\end{lemma}
\proof Choose a smooth function
$\phi:\mathbb R^2 \to [0,1]$ such that $\phi(A) = 0$ and $\phi(B) = 1$ and a
regular value $c \in (0,1)$. Then $\phi^{-1}(c)$ is a countable
union of properly embedded lines and simple closed curves.
Each component of $\phi^{-1}(c)$ has a collar neighborhood which
is disjoint from the other components.
Let $U$ denote the component of the complement of $\phi^{-1}(c)$
which contains $B$ and let $X$ denote the frontier of $U$.
Then $X$ also separates $A$ and $B$ and $X$
consists of a countable subcollection of the components
of $\phi^{-1}(c),$ each of which is also a component of $X$. The set
$U$ is the component of $\mathbb R^2 \setminus X$ which contains $B.$
Each component $L$ of $X$ separates $\mathbb R^2$ into two
open sets, one of which contains $B$ and $X \setminus L$
and the other of which is disjoint from $X$ and $B$.
Let $V$ be the component of $\mathbb R^2 \setminus X$ containing $A.$
Consider a curve $\gamma$ running from a point of $A$ to a point of
$B$ and let $L_0$ be the first component of $X$ which $\gamma$ intersects.
The component $L_0$ is independent of the choice of $\gamma,$
since $L_0$ separates $A$ from all other components of $X.$ It
follows that $A$ and $B$ are in different components of the
complement of $L_0$ since otherwise they could be joined by a
$\gamma$ which does not intersect $L_0.$
\endproof
\begin{lemma} \label{planar topology} If $U \subset \mathbb R^2$ is open and connected then each component $Z$ of the complement of $U$ has connected frontier and connected complement.
\end{lemma}
\begin{proof}
The complement of $Z$ is the union of $U$ with some of its
complementary components and is therefore connected. If the frontier
$W$ of $Z$ is not connected then by Lemma~\ref{regular point} there is
a separation of $W$ by a set $Y \subset \mathbb R^2 $ that is either a simple
closed curve or a properly embedded line. Note that both components
of $\mathbb R^2 \setminus Y$ must intersect both the interior of $Z$ and
$\mathbb R^2 \setminus Z$. Since $Y$ is disjoint from the frontier $W$ of
$Z$, it is contained in either the interior of $Z$ or in $\mathbb R^2
\setminus Z$. In the former case $Y$ separates $Z$ and
in the latter case $Y$ separates $\mathbb R^2 \setminus Z$. This contradicts the fact that
$Z$ and $\mathbb R^2 \setminus Z$ are connected and so proves that the
frontier of $Z$ is connected.
\end{proof}
\section{Hyperbolic Structures} \label{hyperbolic}
In this section we establish notation and recall
standard results about hyperbolic structures on surfaces.
Suppose that $M$ is a connected open subset
of $S^2$ that is not homeomorphic to either the open disk or open
annulus. A closed curve $\alpha$ in $M$ is {\em essential} if it is
not freely homotopic to a point and is {\em peripheral} if it is
freely homotopic into arbitrarily small neighborhoods of an end of
$M$.
If $M$ has infinitely many ends then it can be written as an increasing union of finitely
punctured compact connected subsurfaces $M_i$ whose boundary components
determine essential non-peripheral homotopy classes in $M$. We may
assume that boundary curves in $M_{i+1}$ are not parallel to boundary
curves in $M_i$. It is straightforward (see \cite{casble:nielsen}) to put compatible
hyperbolic structures on the $M_i$'s whose union defines a complete
hyperbolic structure on $M$. Of course $M$ also has a complete hyperbolic structure when it only has finitely many ends. All hyperbolic structures in this paper are assumed to be complete.
We use the Poincar\'e disk model for the hyperbolic plane $H$. In
this model, $H$ is identified with the interior of the unit disk and
geodesics are segments of Euclidean circles and straight lines that
meet the boundary in right angles. A choice of hyperbolic structure on
$M$ provides an identification of the universal cover $\tilde M$ of $M$
with $H$. Under this identification covering translations become
isometries of $H$ and geodesics in $M$ lift to geodesics in $H$. The
compactification of the interior of the unit disk by the unit circle
induces a compactification of $H$ by the \lq circle at infinity\rq\
$S_{\infty}$. Geodesics in $H$ have unique endpoints on $S_{\infty}$.
Conversely, any pair of distinct points on $S_{\infty}$ are the endpoints
of a unique geodesic.
Each covering translation $T: H \to H$ extends to a
homeomorphism (also called) $T : H \cup S_{\infty} \to H \cup
S_{\infty}$. The fixed point set of a non-trivial $T$ is either one or
two points in $S_{\infty}$. We denote these point(s) by $T^+$ and $T^-$,
allowing the possibility that $T^+ = T^-$. If $T^+ = T^-$, then $T$
is said to be {\it parabolic}; a root-free parabolic covering
translation with fixed point $P$ is sometimes written $T_P$. If $T^+$
and $T^-$ are distinct, then $T$ is said to be {\it hyperbolic} and we
may assume that $T^+$ is a sink and $T^-$ is a source; the unoriented
geodesic connecting $T^-$ and $T^+$ is called the {\em axis} of $T$.
A root-free covering translation with axis $\tilde \gamma$ is sometimes
denoted $T_{\tilde \gamma}$.
Each essential non-peripheral $\alpha \subset M$ is homotopic to a unique closed geodesic $\gamma$. For each lift $\tilde \alpha \subset H$, the homotopy between $\alpha$ and $\gamma$ lifts to a bounded homotopy between $\tilde \alpha$ and a lift $\tilde \gamma$ of $\gamma$ which is the axis of a hyperbolic covering translation $T$. The ends of both lines $\tilde \alpha$ and $\tilde \gamma$ converge to $T^-$ and $T^+$.
Similarly, each essential peripheral $\alpha$ is homotopic to a (non-unique) horocycle $\gamma$. For each lift $\tilde \alpha \subset H$, the homotopy between $\alpha$ and $\gamma$ lifts to a bounded homotopy between $\tilde \alpha$ and a lift $\tilde \gamma$ of $\gamma$. Both ends of both lines $\tilde \alpha$ and $\tilde \gamma$ converge to a single point $T^{\pm}$ that is the unique fixed point of a parabolic covering translation $T$.
Suppose now that $f : M \to M$ is a homeomorphism. Identify $H$ with $\tilde M$ and write $\tilde f : H \to
H$ for lifts of $f : M \to M$ to the universal cover. If $f:M \to M$ and $g:M \to M$ are homotopic and the homotopy between $f$ and $g$ lifts to a homotopy between $\tilde f:H \to H$ and $\tilde g : H \to H$ then we say that $\tilde f$ and $\tilde g$ are equivariantly homotopic. A proof of the following fundamental
result of Nielsen theory appears in Proposition 3.1 of
\cite{han:fpt}.
\begin{prop} \label{nielsen} Every lift $\tilde f : H \to H$ extends
uniquely to a homeomorphism (also called) $\tilde f : H \cup S_{\infty} \to
H \cup S_{\infty}$.
If $\tilde f $ and $\tilde g $ are equivariantly homotopic lifts of $f:M \to M$ and $g:M \to M$ then $\tilde f|_{S_{\infty}} = \tilde g|_{S_{\infty}}$.
\end{prop}
For any extended lift $\tilde f : H \cup S_{\infty} \to
H \cup S_{\infty}$ there is an {\it associated action $\tilde f_\#$ on
geodesics [horocycles] in $H$ }defined by sending the geodesic with endpoints $P$
and $Q$ to the geodesic [horocycle] with endpoints $\tilde f(P)$ and $\tilde f(Q)$.
The action $\tilde f_\#$ projects to an {\it action $f_\#$ on geodesics [horocycles]
in $M$}.
Proposition~\ref{nielsen} implies that $f_\#$ depends only on the isotopy class of $f$.
The following results are well known and follow easily from the definitions.
\begin{lemma} \label{basic lemma 1}Suppose for $i=1,2$, that $\alpha_i$ is an essential closed curve homotopic to the geodesic [horocycle] $\gamma_i$. Then $f(\alpha_1)$ is freely homotopic to $\alpha_2$ if and only if $f_\#(\gamma_1) = \gamma_2$.
\end{lemma}
\begin{lemma} \label{basic lemma 2} For any extended lift $\tilde f : H \cup S_{\infty} \to H \cup S_{\infty}$ and extended covering translation $T: H \cup S_{\infty} \to H \cup S_{\infty}$, the following are equivalent:
\begin{enumerate}
\item $\tilde f$ commutes with $T$.
\item $\tilde f$ fixes $T^{+}$ or $T^-.$
\item $\tilde f$ fixes $T^{+}$ and $T^-.$
\end{enumerate}
\end{lemma}
\section{An intermediate proposition} \label{sec: intermediate}
Recall that $F \in \Diff_\mu(S^2)$ has entropy zero.
By Theorem~1.2, Lemma~6.3 and Remark~6.4 of \cite{fh:periodic}, $F$ is
isotopic rel $\Fix(F)$ to a composition $F_1$ of non-trivial Dehn
twists along the elements of a finite (possibly empty) set ${\cal R}_F$
of disjoint simple closed
curves. By the main theorem of \cite{brnkist}, each component $M$ of
${\cal M} = S^2 \setminus \Fix(F)$ is $F$-invariant. Let $f =F|_M : M\to M$
and let ${\cal R}$ be the subset of ${\cal R}_F \cap M$ consisting of elements that
are non-peripheral in $M$. The restriction to $M$ of the isotopy from
$F$ to $F_1$ can be modified to an isotopy from $f$ to a a composition
of non-trivial Dehn twists along the elements of ${\cal R}$. The fact that
$F$ preserves $\mu$ implies that $M$ has $f$-recurrent points. The
Brouwer plane translation theorem and the fact that $\Fix(f) =
\emptyset$ imply that $M$ is not an open disk.
The proof of the first item of Theorem~\ref{thm: annuli} requires renormalization and so is completed after the proofs of the second and third items. To clarify the logic we introduce Proposition~\ref{intermediate}, which contains the second and third items of Theorem~\ref{thm: annuli} plus two additional properties of the annuli.
\begin{prop} \label{intermediate} Suppose that $M$ is a component of ${\cal M} = S^2 \setminus \Fix(F)$ and that $f=F|_M:M \to M$. Then there is a countable collection ${\cal A}$ of pairwise disjoint
open $f$-invariant annuli such that
\begin{enumerate}
\item For each compact set $X \subset M$ there is a constant $K_X$
such that any $f$-orbit that is not contained in some $U \in {\cal A}$
intersects $X$ in at most $K_X$ points. In particular each $x \in
{\cal B}(f)$ is contained in some $U \in {\cal A}$.
\item For each $U \in {\cal A}$ and $z$ in the frontier of $U$ in $S^2$, there
are components $F_+(z)$ and $F_-(z)$ of $\Fix(F)$ so that $\omega(F,z)
\subset F_+(z) $ and $\alpha(F,z) \subset F_-(z)$.
\item For each $U \in {\cal A}$ and each component $C_M$ of the frontier of
$U$ in $M$, $F_+(z)$ and $F_-(z)$ are independent of the choice of $z
\in C_M$.
\item If $U \in {\cal A}$, and $f_c:U_c \to U_c$ is the extension
to the annular compactification of $U$, then each component of $\partial U_c$
corresponding to a non-singular end of $U$ contains a fixed point of $f_c.$
\end{enumerate}
\end{prop}
In the special case that $M$ is an annulus, let ${\cal A}$ be the single annulus $M$. Items (1) - (3) are obvious and item (4) follows from Lemma~5.1 of \cite{fh:periodic}. The constructions and analysis needed for the case that $M$ is not an annulus are carried out in sections \ref{sec: endpoints} through \ref{sec: annuli}. The final formal proof of Proposition~\ref{intermediate} occurs at the end of section~\ref{sec: annuli}.
\section {The endpoint maps $\tilde \alpha$ and $\tilde \omega$ and annular covers} \label{endpoint} \label{sec: endpoints}
In this section we begin the proof of Proposition~\ref{intermediate} in the case that $M$ has at least three ends.
Equip $M$ with
a complete hyperbolic structure as described in
section~\ref{hyperbolic}. The full pre-image in $H$ of the
reducing set ${\cal R}$ is denoted $\tilde {\cal R}$. The closure of a component of $H \setminus \tilde {\cal R}$ is called a {\em domain}. If ${\cal R}=\emptyset$ then $H$ is the unique domain but otherwise there are infinitely many domains. The subgroup of covering translations that preserves a domain $\tilde C$ is denoted $\Stab(\tilde C)$. For each domain $\tilde C$, there is an associated lift $\tilde f_{\tilde C} : H \to H$ that commutes with each $T \in \Stab(\tilde C)$. This lift is also characterized by the property that $\Fix(f|_{S_\infty})$ is equal to the intersection of the closure of $\tilde C$ with $S_\infty$; thus $\Fix(f|_{S_\infty})$ is a Cantor set if ${\cal R} \ne \emptyset$ and all of $S_{\infty}$ if ${\cal R} = \emptyset$.
\begin{lemma} For each lift $\tilde f$ of $f$
and each $\tilde x \in H$, $ \alpha(\tilde f, \tilde x)$ and
$\omega(\tilde f, \tilde x)$ are single points in $S_{\infty}$.
\end{lemma}
\proof
The Brouwer translation theorem implies that $\omega(\tilde f, \tilde x)
\subset S_\infty$. We assume that $\omega(\tilde f_, \tilde x)$ is not a
single point and argue to a contradiction. It must be the
case that $\omega(\tilde f, \tilde x) \subset S_\infty \cap \Fix(\tilde f).$
If not, a non-fixed point $z\in \omega(\tilde f_, \tilde x)$
would have a free neighborhood whose intersection with $H$ would be a free disk visited by the
orbit of $\tilde x$ more than once (indeed infinitely often).
According to Proposition~(1.3) of \cite{franks:poincare} this
implies $\tilde f$ has a fixed point in $H$ -- a contradiction.
Since $\omega(\tilde f, \tilde x)$ consists of fixed points it is
straightforward to see that it is also connected.
If $\Fix(\tilde f)$ does not contain an interval we are done. Otherwise, every covering translation with one endpoint in this interval commutes with $\tilde f$ and so preserves $\Fix(\tilde f)$. It follows that $\Fix(\tilde f) = S_{\infty}$. A proof of the lemma in this special case is given in Proposition~9.1 of \cite{fh:periodic}.
\endproof
In addition to lifts of $f$ to the universal cover $H$ we will also use lifts of $f$ to infinite cyclic covers.
\begin{defns} \label{annulus cover} Suppose that $\sigma$ is a closed geodesic that is either
equal to an element of ${\cal R}$ or disjoint from every element of ${\cal R}$. For
each lift $\tilde \sigma$, let $T_{\tilde \sigma}$ be a root free
covering translation with axis $\tilde \sigma$. Choose a domain $\tilde C$ that contains
$\tilde \sigma$. (If $\sigma \in {\cal R}$ then there are two choices but
otherwise there is just one.) Since $\tilde f_{\tilde C}$ fixes the
ends of $\tilde \sigma,$ it commutes with $T_{\tilde \sigma}$,
by Lemma~(\ref{basic lemma 2}).
The {\em annular cover} $A_{\sigma}$ is the closed
annulus that is the quotient space of $(H \cup S_{\infty}) \setminus
T_{\tilde \sigma} ^{\pm}$ by the action of $T_{\tilde \sigma}$ and $
f_\sigma : A_{\sigma} \to A_{\sigma}$ is the homeomorphism induced
by $\tilde f_{\tilde C}$. If
$\alpha(\tilde f_{\tilde C}, \tilde x)$ is not an endpoint of $\tilde \sigma$ then
$\alpha(f_\sigma, \hat x)$ is a single point in $\partial A_{\sigma}$
and similarly for $\omega( f_\sigma, \tilde x)$.
Similarly, if $\tilde \sigma$ is a lift of a horocycle $\sigma$ then both
ends of $\tilde \sigma$ converge to a point $P \in S_{\infty}$ and there is
a root free covering translation $T_P$ that preserves $\tilde
\sigma$. Let $\tilde C$ be the unique domain that contains $\tilde
\sigma$. In this case, the {\em annular cover} $A_{\sigma}$
is the half-open annulus that is the
quotient space of $(H \cup S_{\infty}) \setminus P$ by the action of $T_P
$ and the boundary is a single circle denoted
$\partial A_\sigma.$
As in the previous case, $\tilde f_{\tilde C}$ induces a homeomorphism $ f_\sigma :
A_{\sigma} \to A_{\sigma}$. The end of $A_{\sigma}$ corresponding to $P$
projects homeomorphically to the end of $M$ circumscribed by $\sigma$.
We can compactify this end exactly as in Definition~\ref{defn: annular comp} to form a closed annulus
$A_\sigma^c$. There is an extension of $f_\sigma$ (also called
$f_\sigma$) to a homeomorphism of $A_\sigma^c$.
\end{defns}
As the notation suggests, $f_\sigma$ is
independent of the choice of $\tilde C$ and, up to conjugacy, the choice
of lift $\tilde \sigma$. The former follows from the fact that if $\tilde C_1$
and $\tilde C_2$ contain $\tilde \sigma$ then $\tilde f_{\tilde C_1}$ and $\tilde
f_{\tilde C_1}$ differ by an iterate of $T_{\tilde \sigma}$ and the latter
from the fact that if $\tilde \sigma$ is replaced with $S(\tilde \sigma)$
for some covering translation $S$ then $\tilde C$ is replaced by $S(\tilde
C)$ and $ T_{\tilde \sigma}$ is replaced by $ST_{\tilde \sigma}S^{-1}$.
\begin{lemma}\label{twist or not}
Suppose that $\sigma$ is a horocycle or a closed geodesic that is either
equal to an element of ${\cal R}$ or disjoint from every element of ${\cal R}$.
\begin{enumerate}\item For each closed geodesic $\sigma$,
$\Fix(f_\sigma|_{\partial A_{\sigma}})$ intersects both components
of $\partial A_{\sigma}$. If $\sigma \in {\cal R}$ then $f_\sigma$ is
isotopic rel $\Fix(f_\sigma|_{\partial A_{\sigma}})$ to a Dehn twist
of the same index that $f$ twists around $\sigma$. If $\sigma \not
\in {\cal R}$ then $f_\sigma$ is isotopic rel $\Fix(f_\sigma|_{\partial
A_{\sigma}})$ to the identity.
\item For each horocycle $\sigma$, $\Fix(f_\sigma|_{\partial
A_{\sigma}}) \ne \emptyset$.
\end{enumerate}
\end{lemma}
\begin{proof}
Suppose at first that $\tilde \sigma$ is a lift of the closed geodesic $\sigma$.
If $\sigma \notin {\cal R}$ then the closure of the domain $\tilde C$ that contains $\tilde \sigma$ intersects both components of $S_{\infty} \setminus \tilde \sigma^{\pm}$. The points in this intersection are fixed by $\tilde f_{\tilde C}$ and project to fixed points for $f_\sigma$ in both components of $\partial A_{\sigma}$, all of which are in the same Nielsen class for $f_{\sigma}$. It follows that $f_\sigma$ is isotopic rel $\Fix(f_\sigma|_{\partial A_{\sigma}})$ to the identity.
If $\sigma \in {\cal R}$ and $\tilde C_1$ and $\tilde C_2$ are the domains that contain $\tilde \sigma$ then points in the intersection of the closure of $\tilde C_1$ with $S_{\infty}$ are fixed by $\tilde f_{\tilde C_1}$ and project to fixed points for $f_\sigma$ in one component of $\partial A_{\sigma}$ and points in the intersection of the closure of $\tilde C_2$ with $S_{\infty}$ are fixed by $\tilde f_{\tilde C_2}$ and project to fixed points for $f_\sigma$ in the other component of $\partial A_{\sigma}$. If $f$ twists with degree $k$ around $\sigma$ then $\tilde f_{\tilde C_1}$ and $ \tilde f_{\tilde C_2}$ differ by $T_{\tilde \sigma}^{k}$ so $f_\sigma$ is
isotopic rel $\Fix(f_\sigma|_{\partial A_{\sigma}})$ to a Dehn twist of index $k$. This completes the proof of (1).
The proof for (2) is similar.
\end{proof}
\section{Reducing Arcs in Annular Covers}
In this section we recall, adapt and improve definitions and results from section 10 of \cite{fh:periodic}, where the assumption is that $F$ is periodic point free and isotopic rel $\Fix(F)$ to the identity as opposed to our current assumption that $F$ has zero entropy and is isotopic rel $\Fix(F)$ to a composition of Dehn twists on the elements of ${\cal R}$.
\begin{notn} \label{notation}
We assume throughout this section that $f_\sigma : A_\sigma \to A_\sigma$ is as in
Definition~\ref{annulus cover} and that $\hat x_1,\ldots, \hat
x_r$ are points in the interior of $A_{\sigma}$ such that the $
\alpha(f_\sigma,\hat x_i)$'s are distinct points in $\partial
A_{\sigma}$ and the $ \omega(f_\sigma,\hat x_i)$'s are distinct points
in $\partial A_{\sigma}$.
Let $N = \Int(A_{\sigma})$, let $\hat X \subset N$ be the
union of the $f_\sigma$-orbits of the $\hat x_i$'s and let $N_{\hat X} =N \setminus \hat X$ equipped with a complete hyperbolic structure.
\end{notn}
A properly embedded line $\ell \subset N_{\hat X}$ is {\em essential}
if it is not properly isotopic into arbitrarily small neighborhoods of
some end of $N_{\hat X}$. Each essential $\ell$ is properly isotopic
to a unique geodesic. The action of $f_\sigma$
on isotopy classes of properly
embedded lines in $N_{\hat X}$ is captured by the map ${f_\sigma} _\#$ on
geodesics defined in section~\ref{hyperbolic}.
If an embedded path $\beta \subset N$ has endpoints in $ \hat X$ but is otherwise disjoint from $\hat X$ then the interior of $\beta$ determines a properly embedded line $\ell \subset N_{\hat X}$. Proper isotopy of $\ell$ in $N_{\hat X}$ corresponds to isotopy rel $\hat X$ of $\beta$ in $N$. If $\ell$ is essential [resp. a geodesic in $N_{\hat X}$] then we say that {\em $\beta$ is essential [resp. a geodesic rel $\hat X$ in $N$]}. There is an induced map ${f_\sigma} _\#$ on geodesics rel $\hat X$ in $N$ such that ${f_\sigma} _\#(\beta)$ is the unique geodesic path in the isotopy class rel $\hat X$ of $f_{\sigma}(\beta)$.
\begin{defn} \label{defn: translation arc geodesic} An arc $\beta' \subset N$ connecting $\hat x \in \hat X$ to $f_{\sigma}(\hat x)$ is called a {\em translation arc} for $x$ if $f_{\sigma}(\beta') \cap \beta' = f_{\sigma}(\hat x)$. The geodesic rel $\hat X$ $\beta \subset N$ determined by $\beta'$ is called a {\em translation arc geodesic} for $x$ relative to $\hat X$.
\end{defn}
\begin{lemma} \label{proper rays} For each $1 \le i \le r$
there are translation arc geodesics
$\hat \beta_i^+$ and $\hat \beta_i^-$ for some points in the
orbit of $\hat x_i$ such that
\begin{enumerate}
\item $\hat B_i^+ = \cup_{j=0}^\infty {f^j_\sigma} _\#(\hat \beta^+_i)$ is an embedded ray that converges to
$\omega(f_{\sigma}, \hat x_i)$.
\item $\hat B_i^- = \cup_{j=0}^\infty {f^j_\sigma}_\#(\hat \beta^-_i)$ is an embedded ray that converges to
$\alpha(f_{\sigma}, \hat x_i)$.
\item the $\hat B_i^{\pm}$'s
are all disjoint.
\end{enumerate}
\end{lemma}
\proof Items (1) and (2) are proved explicitly and item (3) is unstated but proved implicitly in Lemma~10.6 of \cite{fh:periodic} under the
hypotheses that $\Per(f_\sigma) =\emptyset$ and ${\cal R} = \emptyset$.
The former is only used to conclude that $\Fix(f_\sigma) = \emptyset$ and the latter is only used to guarantee that $\tilde f_{\tilde C}$ commutes
with $T_{\tilde \sigma}$. Since these hold in our context, the proof given in \cite{fh:periodic} applies in this context as well.
\endproof
\begin{remark} \label{proper streamline}Assume the notation of Lemma~\ref{proper rays}. For each $i$ the map ${f_{\sigma}}_\#$ defined on geodesics rel $\hat X$ induces a self-map of $ B_i^+$ that is conjugate to a standard tranlslation of $[0,\infty)$ into itself. The analogous statement also holds for $B_i^-$ with respect to ${f^{-1}_{\sigma}}_\#$.
\end{remark}
In the next definition we recall the main items associated to the $\hat \beta_i^\pm$'s. Further details can be found in section 10 of \cite{fh:periodic}.
\begin{defn} Assume the notation of Lemma~\ref{proper rays}. Let $V^\pm_i $ be the neighborhood of $B^\pm_i$ that has connected geodesic boundary rel $X$, intersects $X$ exactly in $B^\pm_i \cap X$ and deformation retracts to $B^\pm_i$. The nested intersection $V^+_i \supset {f_{\sigma}}_\#(V^+_i) \supset {f^2_\sigma}_\#(V^+_i) \supset \ldots$ is empty and similarly for $V^-_i \supset {f^{-1}_\sigma}_\# (V^-_i) \supset {f^{-2}_\sigma}_\# (V^-_i) \supset \ldots$. We say that $V^\pm_i$ is the {\em translation neighborhood} determined by $\hat \beta_i^\pm$. Item (3) in Definition~\ref{proper rays} implies that the $V_i^\pm$'s are disjoint.
The subsurface $W = N \setminus (\hat X \cup (\bigcup_{i=1}^r V^\pm_i))$ is finitely punctured. We write $\partial W = \partial_+W \cup \partial_- W$ where $\partial_{\pm}W = \cup_{i=1}^r \partial V_i^\pm$. Then ${f_{\sigma}}_\#(\partial_+W) \cap W = \emptyset$ and $\partial_-W \cap {f_{\sigma}}_\#(W) = \emptyset$. We say that $W$ is the {\em Brouwer subsurface determined by the $\hat \beta_i^\pm$'s}.
Let $RH(W,\partial_+ W)$ be the set of non-trivial relative homotopy classes $[\tau]$ determined by embedded arcs $(\tau, \partial \tau) \subset (W, \partial_+W)$. Denote $\tau$ with its orientation reversed by $-\tau$ and
$[-\tau]$ by $-[\tau]$. Let ${\cal T}_i \subset RH(W,\partial_+ W)$ consist of one representative ($[\tau]$ or $-[\tau]$) of each unoriented homotopy class that is represented by a component of ${f_{\sigma}^n}_\#(\hat \beta_i^-) \cap W$ for some $n> 0$. Since the ${f_{\sigma}^n}_\#( \hat \beta^-_i )$'s are disjoint and simple and since W is finitely punctured, ${\cal T}_i$ is finite. We say that ${\cal T}_i$ is the {\em fitted family} determined by the $\hat \beta_i^{\pm}$'s.
\end{defn}
If ${f_{\sigma}^n}_\#(\hat \beta_i^-) \subset V_i^+$ for some $n$ then $\hat B_i = \cup_{j=-\infty}^\infty {f_{\sigma}}_\#^j(\hat \beta^-_i)$ is a properly embedded ${f_{\sigma}}_\#$-invariant line in $N$ whose endpoints converge to $\alpha(f_{\sigma}, \hat x_i)$ and $\omega(f_{\sigma}, \hat x_i)$ by Lemma~\ref{proper rays} and the assumption that $V_i^+$ is a translation neighborhood.
To analyze the limit set of $\hat B_i$ when no such $n$ exists, we make use of a map that assigns to each finite set ${\cal T}\subset RH(W,\partial_+ W)$ another finite set ${f_{\sigma}}_\#({\cal T} ) \cap W$. We abuse notation slightly and write ${\cal T} =\{[\tau_i]\}$ where each $[\tau_i] \in RH(W,\partial_+ W)$; we do not assume that the $[\tau_i] $'s are distinct.
Choose a homeomorphism $h :N \to N$ that is isotopic to $f_{\sigma}$ rel $\hat X$ such that $h(L) ={f_{\sigma}}_\#(L)$ for each component $L$ of $\partial W$. For any arc $\tau \subset W$ with endpoints on $\partial_+W$, $ h(\tau)$ is an arc in $ h(W) = {f_{\sigma}}_\#(W)$ with endpoints on $ {f_{\sigma}}_\#(\partial_+W)$; in particular, $ h(\tau) \cap \partial_-W = \emptyset$ and $\partial h(\tau) \cap W = \emptyset$. Let $ {f_{\sigma}}_\#(\tau) \subset {f_{\sigma}}_\#(W)$ be the geodesic arc that is isotopic rel endpoints to $ h(\tau)$. The components $\tau_1,\ldots,\tau_r$ of $ {f_{\sigma}}_\#(\tau) \cap W$ are arcs in $W$ with endpoints in $\partial_+W$. Define $ {f_{\sigma}}_\#([\tau]) \cap W = \{[\tau_1],\ldots,[\tau_r]\}$. It is shown in \cite{han:fpt} (see pages 249 - 250) that ${f_{\sigma}}_\#([\tau]) \cap W$ is well defined.
More generally if ${\cal T} $ is a finite collection of elements of
$RH(W,\partial_+ W)$ then we define $ {f_{\sigma}}_\#({\cal T}) \cap W = \cup_{[\tau]
\in {\cal T} }( {f_{\sigma}}_\#([\tau]) \cap W)$. Note that $
{f_{\sigma}}_\#(\cdot )\cap W$ can be iterated. Recursively define $
({f_{\sigma}}_\#)^n([\tau]) \cap W = ({f_{\sigma}}_\#)^{n-1}(
{f_{\sigma}}_\#([\tau]) \cap W) \cap W)$.
The next lemma states that one gets the same answer by either iterating the intersection operator or by first iterating $f_{\sigma}$ and then appying the intersection operator once.
\begin{lemma}\label{iteration is well defined} For all $\tau \in RH(W,\partial_+ W)$, $({f_{\sigma}}_\#)^n([\tau]) \cap W=(f^n_\sigma)_\#([\tau]) \cap W$.
\end{lemma}
\proof This is Lemma 5.4 of \cite{han:fpt}.
\endproof
We next record some properties of the fitted family ${\cal T}_i$.
\begin{cor} The fitted family ${\cal T}_i = \{[\tau_1],\ldots [\tau_m]\}$ satisfies
\begin{itemize}
\item [(a)] $\tau_j$ and $\tau_k$ are disjoint and $[\tau_j] \ne \pm [\tau_k]$ for $j \ne k$
\item [(b)] each element of $ {f_\sigma}_\#({\cal T}_i) \cap W$ is $\pm [\tau_j]$ for some $1 \le j \le m$.
\end{itemize}
\end{cor}
\proof The first item is by construction and the second follows from
Lemma~\ref{iteration is well defined}.
\endproof
We used the assumption that $F:S^2 \to S^2$ has zero entropy to conclude that $F$ and hence $f: M \to M$ is isotopic to a composition of Dehn twists along disjoint simple closed curves. Lemma~\ref{no horseshoes} below (c.f. Theorem 5.5(b) of \cite{han:fpt}) is the only other place in which the entropy zero hypothesis is applied.
We say that
{\em $[\tau] \in RH(W,\partial_+ W)$ disappears under iteration} if $ ({f_\sigma}_\#)^n([\tau]) \cap W = \emptyset$ for some $n > 0$ and that {\em ${\cal T}_i$ disappears under iteration} if each element in ${\cal T}_i$ does.
\begin{lemma} \label{no horseshoes} Assume that $\hat \beta^\pm_i$ are translation arc geodesics as in Lemma~\ref{proper rays}, that $W$ is the associated Brouwer subsurface and that ${\cal T}_i$ is the $i^{th}$ associated fitted family. Then for each $n \ge 1$ and $[\tau] \in RH(W,\partial_+ W)$, $ ({f^n_\sigma}_\#)([\tau]) \cap W$ contains at most one element that equals either $[\tau]$ or $-[\tau]$.
\end{lemma}
\proof We assume to the contrary that there exist $[\tau]$ and $n \ge 1$ such $ ({f_\sigma}_\#)^n([\tau]) \cap W$ contains at least two elements that equal either $[\tau]$ or
$-[\tau]$ and argue to a contradiction. The obvious induction argument on $k$ implies that $ ({f_\sigma}_\#)^{kn}([\tau]) \cap W$ contains at least $2^k$ elements that equal
$\pm[ \tau]$.
Let $L_1$ and $L_2$ be the components of $\partial_+W$ that contain the endpoints of $\tau$.
As a first case suppose that $L_1 = L_2$ and that $\tau$ and the
interval in $L_1$ connecting the endpoints of $\tau$ bound a disk $D$
in $N$. Choose an element $\hat x \in \hat X \cap D$ and a compact
essential subannulus $A \subset A_\sigma$ that separates $\hat x$ from
$L_1$. There are at least $2^k$ subpaths $\mu_j$ of $
({f^{kn}_\sigma}) (\tau)$ that cross $A$. Each $\mu_j$ projects to a
path in $M$ whose distance in the metric inherited from $S^2$ is
bounded below by a constant that does not depend on $j$ . Letting
$\tau_M \subset M$ be the projected image of $\tau \subset N$ it
follows that the length of $f^{kn}(\tau_M)$ in the metric on $M$
inherited from $S^2$ grows exponentially in $k$. This contradicts
part (2) of Theorem 2 of \cite{newhouse} and the fact that $F$ has
entropy zero.
In the case that the ends of $L_1$ and $L_2$ converge to distinct components of $\partial A_{\sigma}$, the same argument works with respect to a compact essential subannulus $A \subset A_\sigma$ that separates $L_1$ and $L_2$.
The third case is that $L_1 = L_2$ and that $\tau$ and the interval in $L_1$ connecting the endpoints of $\tau$ define a simple closed curve that is essential in $A_\sigma$. Choose a compact essential annulus $A \subset A_{\sigma}$ that is disjoint from $L_1 \cup \tau$. Choose half-disks whose frontier consist of intervals $I_1$ and $I_2$ in the component of $\partial A_\sigma$ that contains the endpoints of $L_1$ and half-circles $\tilde \rho_1$ and $\tilde \rho_2$ that project to the same simple closed curve $\rho \subset M$. We may assume that the half- disks are disjoint from each other and from $L_i$. Choose thickened arcs $J_1$ and $J_2$ connecting $\tilde \rho_1$ and $\tilde \rho_2$ to the far end of $A$. (In other words $J_1$ and $J_2$ overlap with $A$ in rectangles that cross $A$.)
There are at least $2^k$ subpaths $\mu_j$ of $ ({f^{kn}_\sigma}) (\tau)$ such that each $\mu_j$ either crosses $J_1$, crosses $J_2$, crosses $A$ or is an arc with one endpoint on $\tilde \rho_1$ and the other on $\tilde \rho_2$. We claim that $\mu_j$ projects to a path in $M$ whose distance in the metric inherited from $S^2$ is bounded below by some constant that is independent of $j$. If $\mu_j$ crosses $J_1$, $J_2$ or $A$ then this follows from the compactness of these sets. In the remaining case $\mu_j$ projects to an arc in $M$ with endpoints on $\rho$ that is not homotopic rel endpoints into $\rho$ and again the claim is clear. The proof of the lemma in this case concludes as in the previous cases.
The fourth and final case is that $L_1 \ne L_2$ have endpoints in the same component of $\partial A_\sigma$. The obvious modification of the argument from the third case applies here.
\endproof
The next lemma is based on Theorem~5.5(c) of \cite{han:fpt}. See also Lemma~10.8 of \cite{fh:periodic}.
Suppose that $W$ is a Brouwer subsurface, that $\tau \in RH(W,\partial_+ W)$ and that both endpoints of $\tau $ are contained in a single component $L$ of $\partial_+W$. If the simple closed curve that is the union of $\tau$ with an interval in $L$ does not bound a disk in $N$ then we say that $\tau$ is {\em essential}.
\begin{lemma} \label{lemma:fitted family} Assume that $\hat
\beta^\pm_i$ are translation arc geodesics as in Lemma~\ref{proper
rays}, that $W$ is the associated Brouwer subsurface and that ${\cal
T}$ is an associated fitted family that does not disappear under
iteration. Then there exists $[\tau] \in {\cal T}$ such that
${f_{\sigma}}_\#([\tau])\cap W =
\{[\tau],\pm[\sigma_1],\ldots,\pm[\sigma_{m}]\}$ where each $[\sigma_j] \in {\cal T}$
disappears under iteration. If $\tau$ has both endpoints on the same
component $L$ of $\partial_+W$ then $[\tau]$ is essential.
\end{lemma}
\proof Let $\Gamma$ be the directed graph with one vertex for each
element of ${\cal T}$ and one oriented edge from the vertex
corresponding to $\tau_i \in {\cal T}$ to the vertex corresponding to
$\tau_j \in {\cal T}$ for each occurrence of $\pm[\tau_j]$ in
${f_{\sigma}}_\#([\tau_i]) \cap W$. Lemma~\ref{no horseshoes} implies
that each vertex of $\Gamma$ is contained in at most one
non-repeating oriented closed path in $\Gamma$. There is a partial
order on the vertices of $\Gamma$ defined by $v_1 > v_2$ if there is
an oriented path in $\Gamma$ from $v_1$ to $v_2$ but no oriented path
from $v_2$ to $v_1$. Among all vertices that are contained in an
oriented closed path, choose one, $v$, that is smallest with respect
to the partial order. Let $[\tau]$ correspond to $v$ and let $n$ be
the length of the unique oriented non-repeating closed path $\rho$
through $v$. Then $(f_{\sigma}^n) _\#([\tau])\cap W =
\{\epsilon[\tau],\pm[\sigma_1],\ldots,\pm[\sigma_{m}]\}$ for some $[\sigma_j] \ne [\tau]$ in ${\cal T}$ and
$\epsilon=\pm 1$. We claim that each $[\sigma_j]$ disappears under
iteration. If not then the vertex of $\Gamma$ corresponding to $[\sigma_j]$ is on the oriented closed path containing $\tau$ so there would be distinct paths in $\Gamma$ from $\tau$ to $\sigma_j$, each of which could be completed to a closed path from $\tau$ to itself.
If $n =1$, let $\tau' = \tau$; otherwise let $v'$ be the
vertex following $v$ in $\rho$ and let $\tau'$ be the corresponding
element of ${\cal T}$. Thus $\tau'$ is the unique element of
${f_{\sigma}}_\#([\tau])\cap W$ that does not disappear under iteration.
For the remainder of the proof we make use of the end (pages 253 - 254) of the proof of Theorem~5.5 of \cite{han:fpt}. The case that $\tau$ has endpoints on distinct components of $\partial_+W$ is treated in the last two paragraphs of that proof. (The labels $\tau_i^k$ in the diagram on the bottom of page 253 are incorrect; they should be $\tau_i^\ast$.) The argument given there applies without change in our present context so we may assume that $\tau$ has both endpoints on the same component, say $L$, of $\partial_+W$.
The endpoints of ${f_{\sigma}}_\#(\tau)$ are contained in the component of the complement of $ W$ bounded by $L$. If ${f_{\sigma}}_\#(\tau)$ intersects any other component of the complement of $W$ then at least two elements of ${f_{\sigma}}_\#([\tau]) \cap W$ would be represented by paths with endpoints on distinct components of $\partial_+W$. Since no such paths disappear under iteration, this can not happen and we conclude that each element of ${f_{\sigma}}_\#([\tau]) \cap W$, and in particular $\tau'$, has both endpoints on $L$.
Both ends of $L$ converge to the same end of $N$. The argument given in the first and second paragraphs on page 253 of \cite{han:fpt} (which is a proof by contradiction) carries over without change to this context and proves that $\tau$ is essential. By symmetry, $\tau'$ is also essential.
If $\tau \ne \tau'$ then either $\tau$ is contained in the annulus bounded by $\tau'$ and an interval in $L$ or $\tau'$ is contained in the annulus bounded by $\tau$ and in interval in $L$. There is a punctured rectangle $D $ bounded by $\tau, \tau'$ and intervals in $L$. There are only finitely many punctures contained in disks bounded by an element of ${\cal T}$ and an arc in $L$. Thus, for all sufficiently large $k$, $f_{\sigma}^{kn}(D)$ does not contain any punctures in $W$. It follows that $(f_{\sigma}^{kn})_\#([\tau])\cap W = (f_{\sigma}^{kn})_\#([\tau'])\cap W $. But $\tau$ [resp. $\tau'$] is the unique element of $(f_{\sigma}^{kn})_\#([\tau])\cap W $ [resp. $(f_{\sigma}^{kn})_\#([\tau'])\cap W $] that does not disappear under iteration. This contradicts the assumption that $\tau \ne \tau'$. We conclude that $\tau = \tau'$ and hence that $n =1$. Since $f_\sigma$ is orientation preserving and ${f_{\sigma}}_\#(L)$ is parallel to $L$, it follows that $\epsilon = 1$.
\endproof
\begin{defns} \label{disappears} Suppose that $W$ is a Brouwer subsurface, that ${\cal T}$ is a fitted family and that $\tau \in {\cal T}$. If $L_1$ and $L_2$ are the components of $\partial_+W$ that contain the initial and terminal endpoints $x_1$ and $x_2$ of $\tau$ respectively, and if some component of the complement of $L_1 \cup L_2 \cup \tau$ is contractible in $N_X$ then we say that $[\tau] $ is {\em peripheral}. Equivalently, $\tau$ is peripheral if there are rays $R_1 \subset L_1$ and $ R_2 \subset L_2$ whose initial points are $x_1$ and $x_2$ such that the line $ R_1^{-1}\tau R_2$ can be properly isotoped rel $X$ into any neighborhood of some end of $N$.
\end{defns}
Our next result is based on Lemma~6.4 of \cite{han:fpt}.
\begin{lemma} \label{reducing line} Suppose that $\hat \beta^\pm_i$ are translation arc geodesics as in Lemma~\ref{proper rays}, that $W$ is the associated Brouwer subsurface, that ${\cal T}$ is an associated fitted family and that $\tau \in {\cal T}$ is non-peripheral and satisfies ${f_{\sigma}}_\#([\tau])\cap W = \{[\tau],[\sigma_1],\ldots,[\sigma_{m}]\}$ where each $[\sigma_j]$ disappears under iteration. Let $L_1$ and $L_2$ be the (possibly equal) components of $\partial_+W$ that contain the initial and terminal endpoints $v_1$ and $v_2$ of $\tau$. Then there are rays $R_1 \subset L_1$ and $R_2 \subset L_2$ such that $R_1^{-1}\tau R_2$ is isotopic to an ${f_{\sigma}}_\#$-invariant geodesic line $\mu$. \end{lemma}
\proof To simplify the notation somewhat let $h = f_{\sigma}$.
Choose a lift $ \tilde \tau$ of $\tau$ in the compactified universal cover $H \cup S_{\infty}$ of $N _X$ and for $k =1,2$, let $\tilde L_k$ be the lift of $L_k$ that contains the endpoint $\tilde v_k$ of $\tilde \tau$ that projects to $v_k$. Since $\tau$ is not peripheral, $\tilde L_1$ and $\tilde L_2$ do not have a common endpoint. The endpoints of $\tilde L_1$ and the endpoints of $\tilde L_2$ bound disjoint intervals $J_1$ and $J_2$ in $S_\infty$.
Since $[\tau] \in h_\#([\tau])\cap W$, there is a lift $\tilde h : H \cup S_{\infty} \to H \cup S_{\infty}$ such that $\tilde L_1$ separates $\tilde h_\#(\tilde L_1)$ from $\tilde L_2$ and $\tilde L_2$ separates $\tilde h_\#(\tilde L_2)$ from $\tilde L_1$. The sequences $\{h_\#^n(L_k): n=0,1,\ldots\}$ have no accumulation points in $N_X$ so the sequences $\{\tilde h_\#^n(\tilde L_k): n=0,1,\ldots\}$ converge to distinct single points $P_k \in S_\infty$. The geodesic $\tilde \rho$ connecting $P_1$ to $P_2$ in $H$ projects to an $h_\#$-invariant geodesic $\rho \subset N_X$.
For all $k > 1$,\ $ h^k_\#(\tau) \cap W$ is the union of $\tau$ with $\cup_{j=1}^m\cup_{i=0}^{k-1}h_\#^i(\sigma_j) \cap W$. Since each $\sigma_j$ disappears under iteration, there exists $k_0$ such that $ h^k_\#(\tau) \cap W = h^{k_0}_\#(\tau) \cap W$ for all $k \ge k_0$. In particular, $\rho \cap W$ is a finite union of disjoint arcs.
Given $n > 0$, we could replace the translations arc geodesics $\hat \beta^+_i$ by their images under $h^n_\#$. The resulting Brouwer subsurface $W_n$ satisfies $\partial_-W_n = \partial_-W$ and $\partial_+W_n = h^n_\#(\partial_+ W)$. The resulting fitted family ${\cal T}_n$ is the $h^n_\#$-image of ${\cal T}$. Repeating the above construction of $\tilde \rho$ using $\tilde h^n_\#(\tilde \tau)$, $\tilde h^n_\#(\tilde L_1)$ and $\tilde h^n_\#(\tilde L_2)$, we still get $\tilde \rho$. Thus $\rho \cap W_n$ is a finite union of disjoint arcs for all $n \ge 0$. This proves that $\rho$ is a properly embedded line.
It remains to show that $P_i$ is an endpoint of $J_i$. This is
equivalent to one of the endpoints of $\tilde L_i$ being a fixed point of
$\tilde h$ and so also equivalent to $\tilde h_\#(\tilde L_1)$ sharing an
endpoint with $\tilde L_1$. Let $V_1^+$ be the component of the complement of $W$ bounded by $L_1$. The closure $S$ of $V_1^+ \setminus h_\#(V_1^+)$
is a once-punctured strip. Let $\tilde \mu$ be the subpath of $\tilde h_\#(\tilde \tau)$ connecting $\tilde L_1$ to $\tilde h_\#(\tilde L_1)$. The projected
image $\mu \subset h_\#(W)$ is disjoint from $h_\#( V_1^+)$ and
connects $ L_1$ to $ h_\#( L_1)$. If $\mu$ is entirely contained
in $S$ then one of the complementary components of $\mu$ in $S$ is
unpunctured. There are rays $R' \subset L_1$ and $R'' \subset h_\#(L_1)$ so that ${R'}^{-1} \mu R''$ is peripheral. Lifting this to a line that contains $\tilde \tau$ shows that $\tilde L_1$ and $\tilde h_\#(\tilde L_1)$ have a
common endpoint and so completes the proof in this case.
To complete the proof we assume that $\mu$ is not entirely contained in $S$
argue to a contradiction. There is a component $\mu_0$ of $\mu \cap S$ that has both endpoints in $L_1$. Let $\tilde \mu_0$ be the subpath of $\tilde \mu$ that projects to $\mu_0$, let $\tilde S_0$ be the component of the full pre-image of $S$ that contains $\tilde \mu_0$ and let $\tilde L_1'$ and $\tilde L_1''$ be the components of the frontier of $\tilde S_0$ that contain the endpoints of $\tilde \mu_0$. As the notation indicates, both $\tilde L_1'$ and $\tilde L_1''$ are lifts of $L_1$. By construction, there is a subpath $\tilde \rho_0$ of $\tilde \rho$ that connects $\tilde L_1'$ and $\tilde L_1''$. The projection of $\tilde \rho_0$ is a subpath $\rho_0$ of $\rho$ that is contained in $S$ and has both endpoints on $L_1$.
By the same logic there is a subpath $\rho_n$ of $\rho$ that is
contained in the closure $S_n$ of $h^n_\#(V_1^+) \setminus
h^{n+1}_\#(V_1^+)$ with both endpoints on $h^n_\#(L_1)$. Since $\rho$
is an embedded line and each $S_n$ is a once punctured strip no
component of $\rho \cap S_n$ has both endpoints in
$h^{n+1}_\#(L_1)$.
But this means that once $\rho$ crosses from $S_{n+1}$ to $S_n$ it
cannot re-enter $S_{n+1}$ without first entering $W$. Since this
happens only finitely many times, we have a contradiction to the
existence of $\rho_n$ for all $n$. \endproof
\begin{defn} An embedded arc $\rho \subset A_\sigma \setminus \hat X$ with endpoints in $\Fix(f_\sigma) \cap \partial A_{\sigma}$ is a {\em reducing arc} for $f_\sigma$ rel $\hat X$ if it is $f_\sigma$-invariant up to isotopy rel $\hat X$ and rel its endpoints and is non-peripheral in the sense that it is not homotopic rel endpoints and rel $\hat X$ into $\partial A_{\sigma}$.
\end{defn}
The following lemma is similar to Proposition 10.10 of \cite{fh:periodic}. The conclusions of the lemma are more detailed than those of that proposition and apply to $f_\sigma$ and not just some iterate of $f_\sigma$.
\begin{lemma} \label{reducible} Let $\alpha = \cup_{i=1}^r \alpha(f_\sigma,\hat x_i) $ and $\omega = \cup_{i=1}^r \omega(f_\sigma,\hat x_i) $. Assume that for each component $\partial_lA$ of $\partial A$, $\alpha_l = \alpha \cap \partial_lA$ and $\omega_l = \alpha \cap \partial_lA$ have the same cardinality $c_l$ and that if $c_l > 1$ then the elements of $\alpha_l$ and $\omega_l$ alternate around $\partial_lA$. Then
\begin{enumerate}
\item There is a reducing arc $\rho$ for $f_\sigma$ with respect to $\hat X$.
\end{enumerate}
If $r =1$ or if $\sigma \in {\cal R}$ then
\begin{enumeratecontinue}
\item For each $1 \le i \le r$, $\alpha(f_{\sigma},\hat x_i)$ and $\omega(f_{\sigma},\hat x_i)$ belong to the same component of $\partial A_{\sigma}$.
\item For each $1 \le i \le r$, there is a reducing arc $\rho_i$ whose endpoints are $\alpha(f_{\sigma},\hat x_i)$ and $\omega(f_{\sigma},\hat x_i)$.
\item For each $1 \le i \le r$ there is a translation arc geodesic $\hat \beta_i$ for $\hat x_i$ such that $\hat B_i = \cup_{j=-\infty}^\infty h^j_\#(\hat \beta _i)$ is a properly embedded line whose ends converge to $\alpha(f_{\sigma},\hat x_i)$ and to $\omega(f_{\sigma},\hat x_i)$.
\item The $\hat B_i$'s are disjoint.
\end{enumeratecontinue}
If $r =1$
\begin{enumeratecontinue}
\item There is a unique translation arc geodesic for $\hat x_1$.
\end{enumeratecontinue}
\end{lemma}
\begin{remark} Item (6) holds when $\sigma \in {\cal R}$. We limit our proof to the $r=1$ case because this is all we need.
\end{remark}
\proof Choose translation arc geodesics $\hat \beta^\pm_i$ as in Lemma~\ref{proper rays}. Let $V_i^\pm$ , $W$ and ${\cal T}_i$ be the associated translation neighborhoods, Brouwer subsurface and fitted families. We may assume that both ends of $\partial V_i^-$ converge to $\alpha(f_{ \sigma}, \hat x_i)$ and that both ends of $\partial V_i^+$ converge to $\omega(f_{ \sigma}, \hat x_i)$. After performing an isotopy rel $\hat X \cup \partial A_{\sigma}$ we may assume that $f_{\sigma}(L_i) = {f_{\sigma}}_\#(L_i) $ for each $L_i = \partial V_i^+$. In particular, the ends of each $L_i$ are asymptotic to the ends of ${f_\sigma}(L_i)$.
If ${\cal T}_i$ disappears under iteration then $\hat B_i = \cup_{n=-\infty}^{\infty} {f_{\sigma}^n}_\#(\hat \beta^-)$ is a properly embedded ${f_{\sigma}}_\#$-invariant line whose ends converge to $\alpha(f_{\sigma}, \hat x_i)$ and $\omega(f_{ \sigma}, \hat x_i)$. A reducing arc $\rho \subset W$ with endpoints $\alpha(f_{\sigma}, \hat x_i)$ and $\omega(f_{ \sigma}, \hat x_i)$ is obtained by pushing $\hat B_i$
off of itself into a non-contractible complementary component of $N \setminus \hat X$.
Suppose next that ${\cal T}_i$ does not disappear under iteration. Let $[\tau ]\in {\cal T}_i$ satisfy the conclusions of Lemma~\ref{lemma:fitted family}. Let $L_p$ and $L_q$ be the components of $\partial_+W$ containing the initial endpoint $v_p$ and terminal endpoint $v_q $ of $\tau$ respectively. Our hypotheses on $\alpha_l$ and $\omega_l$ imply that each component of the complement of $L_p \cup L_q \cup \tau$ either intersects $\hat X$ or contains an essential subannulus of $A_{\sigma}$. This proves that that $\tau$ is non-peripheral.
Lemma~\ref{reducing line} implies that there are rays $R_p \subset L_p$ and $R_q \subset L_q$ such that $ R_p^{-1}\tau R_q$ is properly isotopic rel $\hat X$ to a geodesic line $\mu$ that is ${f_{\sigma}}_\#$-invariant. To see that $\mu$ is the interior of a reducing arc $\rho$ with endpoints $\alpha(f_{\sigma}, \hat x_p)$ and $\omega(f_{ \sigma}, \hat x_q)$ we need only check that the isotopy rel $\hat X$ between $\mu$ and $f_{\sigma}(\mu)$ can be performed relative to $\hat X \cup \partial A_{\sigma}$. This follows from the fact that the ends of $\mu $ are asymptotic to the ends of $f_{\sigma}(\mu)$. We have now completed the proof of (1).
If $r=1$ then (2) follows from our
assumption on $c_l$. If $\sigma \in {\cal R}$ then Lemma~\ref{twist or not}
implies that no reducing arc can connect the two components of
$\partial A_{\sigma}$ and hence that (2) is satisfied. This completes the proof of (2).
For the remainder of the proof we assume that either $r=1$ or $\sigma
\in {\cal R}$. We prove that each ${\cal T}_i$ disappears
under iteration by assuming that some ${\cal T}_i$ does
not disappear under iteration and arguing to a contradiction. By (2) we may assume that the
endpoints of $L_p$ and $L_q$ belong to belong to the same component
$\partial_lA_{\sigma}$ of $\partial A_{\sigma}$. If $L_p \ne L_q$
then $c_l \ge 2$ and $\alpha_l$ and $\omega_l$ alternate around
$\partial_lA_{\sigma}$. But then $\rho$ must separate
$\alpha(f_{\sigma}, \hat x_i)$ from $\omega(f_{ \sigma}, \hat x_i)$
for some $i$ which contradicts the fact that $\rho$ is a reducing arc.
In the remaining case, $p=q$. Let
$Y$ and $Z$ be the components of the complement of $\rho$ with $Y$
containing the open end of $A_{\sigma}$ if $\sigma$ is a horocycle
and containing the component of $\partial A_{\sigma}$ that does not
contain the endpoints of $L_p$ otherwise. Since $\rho$ is
non-peripheral, $Z$ must contain at least one orbit in $\hat X$. From
$\rho \cap \hat B_i^- = \emptyset$ it follows that $\rho \cap \tau =
\emptyset$. Combined with the fact that $\tau$ is essential (see
Definition~\ref{disappears}) and the fact that $\rho$ is disjoint from
$L_p$, it follows that $V_p $ is contained in $Y$ and so contains at
least one orbit in $\hat X$; in particular, $r > 1$. This completes
the proof that ${\cal T}_i$ disappears under iteration in the $r=1$
case so we may assume by induction that if one works relative to $\hat
X \cap Y$ or relative to $\hat X \cap Z$ then ${\cal T}_i$ disappears
under iteration. In other words, if $\hat x_i \in Y$ ( the argument for
$\hat x_i\in Z$ is symmetric) then for all sufficiently large $n$,
${f_{\sigma}^n}_\#(\hat \beta_i^-)$ is isotopic rel $\hat X \cap Y$ to
an arc $\gamma_{i,n} \subset V_i \subset Y$. Since $\rho$ is a
reducing arc, ${f_{\sigma}^n}_\#(\hat \beta_i^-) \subset Y$. It is a
standard fact that the isotopy rel $\hat X \cap Y$ of
${f_{\sigma}^n}_\#(\hat \beta_i^-)$ to $\gamma_{i,n}$ can be taken
with support in $Y$. It follows that this isotopy is rel $\hat X$
which implies that ${f_{\sigma}^n}_\#(\hat \beta_i^-) \subset V_i$ in
contradiction to the assumption that ${\cal T}_j$ does not disappear
under iteration.
We have now proved that each ${\cal T}_i$ disappears under iteration.
Items (3) -(5) are immediate.
To verify (6), denote the translation arc geodesics that make up $\hat
B_1$ by $\hat \beta_{1,m}$ where ${f_\sigma}_\#(\hat \beta_{1,m}) =
\hat \beta_{1,m+1}$ and where $\hat \beta_{1,0}$ is a translation arc geodesic for $\hat x_1$. We assume that there is a translation arc geodesic $\hat \tau \ne \hat \beta_{1,0}$ for $\hat x_1$ and
argue to a contradiction. Let $\hat \mu$ be the maximum initial
segment of $\hat \tau$ whose interior is disjoint from $\hat B_1$ and
let $\hat y$ be the terminal endpoint of $\hat \mu$. Let $\hat \nu$
be the maximum initial segment of $\hat {f_\sigma}_\#(\hat \tau)$
whose interior is disjoint from $\hat B_1$ and let $\hat z$ be the
terminal endpoint of $\hat \nu$. If $\hat y \in \hat X$ then $\hat z
= \hat f(\hat y)$; otherwise $\hat y$ is in the interior of some $\hat
\beta_{1,m}$ and $\hat z$ is in the interior of $\hat \beta_{1,m+1}$.
If $\hat y \not \in \hat \beta_{1,-1} \cup \hat \beta_{1,0}$ then the endpoints of $\hat \mu$ and $\hat \nu$ are linked in $\hat B_1$ in contradiction to the fact that the interiors of $\hat \mu$ and $\hat \nu$ are disjoint and lie on the same side of $\hat B_1$. We may therefore assume that $\hat y \in \hat \beta_{1,-1} \cup \hat \beta_{1,0}$. In this case, the endpoints of $\hat \mu$ and $\hat \nu$ bound intervals $I_\mu$ and $I_\nu$ in $\hat B_1$ that meet in at most one point. It follows that either the simple closed curve $\hat \mu \cup I_\mu$ or the simple closed curve $\hat \nu \cup I_\nu$ is inessential in $A$ and so bounds a disk that is disjoint from $\hat X$ in contradiction to the fact that these simple closed curves are composed of two geodesic segments. This completes the proof that $\hat \beta_1^-$ is unique and hence the proof of the lemma.
\endproof
We conclude this section with three corollaries of Lemma~\ref{reducible}.
\begin{cor} \label{consistent crossing}If $\sigma \in {\cal R}$ then there do not exist $\hat x_1, \hat x_2 \in A_{\sigma}$ such that $\alpha(f_\sigma, \hat x_1)$ and $\omega( f_\sigma, \hat x_2)$ are contained in one component of $\partial A_\sigma$ and $\alpha( f_\sigma, \hat x_2)$ and $\omega( f_\sigma, \hat x_1)$ are contained in the other component of $\partial A_\sigma$.
\end{cor}
\proof
This follows from item (2) of Lemma~\ref{reducible}
with $r =2$ and $c_0 =c_1 = 1$.
\endproof
\begin{cor} \label{simple} If $\alpha( \tilde f_{\tilde C}, \tilde x) \ne \omega( \tilde f_{\tilde C}f, \tilde x)$, then
the geodesic $\tilde \gamma$ with initial endpoint
$\alpha(\tilde f_{\tilde C}, \tilde x)$ and terminal endpoint $\omega( \tilde f_{\tilde C}, \tilde x)$ projects to a simple geodesic in $C$.
\end{cor}
\proof Let $\tilde f = \tilde f_{\tilde C}$. If the corollary fails then there exists a covering translation $T$ such that $T(\tilde \gamma) \cap \tilde \gamma \ne \emptyset$. Assuming without loss that $\tilde \gamma \subset \Int(\tilde C)$ we have $T(\Int(\tilde C)) \cap \Int(\tilde C) \ne \emptyset$ which implies that $T \in \Stab (\tilde C)$ commutes with $\tilde f$. If $T$ is hyperbolic, let $\tilde \sigma$ be its axis; otherwise let $\tilde \sigma$ be a horocycle that is preserved by $T$. Define $f_\sigma: A_\sigma \to A_\sigma$ as in Definition~\ref{annulus cover}. Since the endpoints $\{\alpha(\tilde f, \tilde x), \omega(\tilde f, \tilde x)\}$ of $\tilde \gamma$ link the endpoints of $T(\tilde \gamma)$, it follows that $\alpha(\tilde f, \tilde x)$ and $ \omega(\tilde f, \tilde x)$ are contained in the same component of $S_{\infty} \setminus \{T^{\pm}\}$. Thus $\alpha(f_\sigma, \hat x)$ and $ \omega(f_\sigma, \hat x)$ are contained in the same component, say $\partial_1A_\sigma$, of $\partial A_\sigma$. By Lemma~\ref{reducible}, applied with $r=1$ and $\hat x_1 = \hat x$, there is a line $\hat B_1\subset \Int(A_\sigma)$ that contains the $f_\sigma$-orbit $\hat X$ of $\hat x$, that is $f_\sigma$-invariant up to isotopy rel $\hat X$ and whose ends converge to $\alpha(f_\sigma, \hat x)$ and $ \omega(f_\sigma, \hat x)$. The lift $\tilde B_1 \subset H$ of $\hat B_1$ that contains $\tilde x$ has endpoint set $\{\alpha(\tilde f, \tilde x), \omega(\tilde f, \tilde x)\}$. The endpoints set $T\{\alpha(\tilde f, \tilde x), \omega(\tilde f, \tilde x)\}$ of $T(\tilde B_1)$ links $\{\alpha(\tilde f, \tilde x), \omega(\tilde f, \tilde x)\}$ since these are also the endpoint sets of $\tilde \gamma$ and $T(\tilde \gamma)$. But then $\tilde B_1$ and $T(\tilde B_1)$ intersect in contradiction to the fact that $\hat B_1$ is an embedded line. \endproof
The next corollary states that if there is twisting across an annular cover then orbits that start and end on one boundary component can not get to close to the other.
\begin{cor} \label{PB} Suppose that $h: A \to A$ is either
\begin{enumerate}
\item $f_\sigma:A_\sigma \to A_\sigma$ for some $\sigma \in {\cal R}$.
\end{enumerate}
or
\begin{enumeratecontinue}
\item $f_\sigma: A^c_\sigma \to A^c_\sigma$ for some horocycle $\sigma$ corresponding to an isolated end of $M$.
\end{enumeratecontinue}
Let $\partial_0 A$ and $\partial_1 A$ be the components of $\partial A$. In case (2) assume that $\partial _0 A$ is the unique component of $\partial A_{\sigma}$ and that if $\Fix(f_{\sigma}|_{\partial_1 A}) \ne \emptyset$ then $f_\sigma$ is not isotopic to the identity rel $\Fix f_{\sigma}|_{\partial A}$. Then there is a neighborhood of $\partial_1 A $ that is disjoint from the $ h$ orbit of any $\hat x$ for which both $\alpha( h, \hat x)$ and $\omega(h, \hat x)$ are contained in $\partial_0 A$.
\end{cor}
\proof If the corollary fails then there exist $\hat x_t \to \hat P\in \partial_1 A$ with $\alpha(h, \hat x_t), \omega(h, \hat x_t) \in \partial_0 A$. After replacing $h$ by some $h^m$ we may assume that
the rotation number of $h|_{\partial_1A}$ is less than $\frac{1}{4}$. In particular there is an interval $J_3$ in $\partial_1 A$ that contains $\hat P,h(\hat P),h^2(\hat P)$ and $h^3(\hat P)$ in that order. Additionally we may assume (Corollary~\ref{twist or not}) that if $\eta$ is a path connecting a fixed point in $\partial_0A$ to $\hat P$ then $h(\eta)$ is not homotopic rel endpoints to the path obtained by concatenating $\eta$ with the subpath $J_1 \subset J_3$ connecting $\hat P$ to $h(\hat P)$. Let $J_2 \subset J_3$ be the subpath connecting $\hat P$ to $h^2(\hat P)$.
Choose contractible neighborhoods $\hat U_i$ of $J_i$ in $A$ such that $\hat U_1 \subset \hat U_2 \subset \hat U_3$ and such that $ h(\hat U_i) \subset \hat U_{i+1}$. Choose lifts $P \in \tilde U_1 \subset \tilde U_2 \subset \tilde U_3$ in $H \cup S_{\infty}$ and let $\tilde h $ be the lift of $h$ such that $\tilde h(P) \in \tilde U_2$. After passing to a subsequence, $\hat x_t \to \hat P$ lifts to a sequence $\tilde x_t \to P$ such that $\tilde x_t, \tilde h(\tilde x_t) \in \tilde U_1$ for all $t$. Recall that a translation arc for $\tilde x$ is a path from $\tilde x_t$ to $\tilde h(\tilde x_t)$ that intersects its $\tilde h$-image only in $\tilde h(\tilde x_t)$. By Lemma 4.1 of \cite{han:fpt}, there is a translation arc $\tilde \delta_t \subset \tilde U_2$ for $\tilde x_t$. Let $\hat \delta_t \subset \hat U_2$ be the projected image of $\tilde \delta_t$. Since $\tilde h(\tilde \delta_t) \cap \tilde \delta_t \subset \tilde U_3$, $h(\hat \delta_t) \cap \hat \delta_t$ is the projected image of $\tilde h(\tilde \delta_t) \cap \tilde \delta_t$. Thus $\hat \delta_t \subset \hat U_2$ is a translation arc for $\hat x_t$. We now fix such a $\hat x_t$ and drop the $t$ subscript.
Assume the notation of Lemma~\ref{reducible} applied with $r=1$ and
$\hat x_1 = \hat x$. The homotopy streamline $\hat B_1$ can be
thought of as an arc $\hat \mu$ with initial endpoint $\alpha(\hat
h, \hat x)$ and terminal endpoint $\omega(h, \hat
x)$. Let $\hat \mu_0$ be the initial subpath of $\hat \mu$ that ends
with $\hat x_1$ and let $\hat \nu \subset \hat U_1$ be a path
connecting $\hat x_1$ to $\hat P$. The path $\hat \eta = \hat
\mu_0\hat \nu$ connects $\alpha(\hat
h, \hat x) \in \partial_0 A$ to $\hat P \in \partial_1 A$.
By the uniqueness part of
Lemma~\ref{reducible}-(6), $\hat \delta$ is isotopic rel $\hat X$ to
the subpath of $\hat \mu$ connecting $\hat x$ to $h(\hat x)$.
It follows that the path $\hat \eta^{-1}h(\hat \eta)$ connecting $P$ to $h(P)$ is
homotopic rel endpoints to $\hat \nu^{-1}\hat \delta h(\hat \nu) \subset
\hat U_2$. Hence
$h(\hat \eta) $ is homotopic rel endpoints to $\eta J_1$. This contradiction completes the proof .
\endproof
\section{ $\omega$-lifts} \label{sec: omega lifts}
We assume throughout this section that ${\cal R} \ne \emptyset$.
The direction (left or right) of twisting on $\sigma \in {\cal R}$ and the choice of a domain $\tilde C$ containing a lift $\tilde \sigma$ induce an orientation on $\tilde \sigma$. We sometimes write $\tilde \sigma_{\tilde C}$ for $\tilde \sigma$ equipped with this orientation. We say that two geodesics in $H$ are {\em anti-parallel} if their initial endpoints are separated in $S_\infty$ by their terminal endpoints.
\begin{lemma} \label{orientations on sigma} The orientations on $\tilde \sigma$ induced from the two domains that contain it are opposite. If $\tilde \sigma_1$ and $\tilde \sigma_2$ are components of $\partial \tilde C$ that project to the same element of ${\cal R}$ then $\tilde \sigma_1$ and $\tilde \sigma_2$ are anti-parallel with the orientations induced from $\tilde C$.
\end{lemma}
\proof The first statement follows from the definitions. The second also uses the assumption that $M$ has genus zero.
\endproof
\begin{lemma} \label{moving on} Suppose that $\tilde C_1$ and $\tilde C_2$ are domains with intersection $\tilde \sigma \subset \tilde {\cal R}$ and that $\tilde f_i =\tilde f_{\tilde C_i}$. If $\tilde \omega(\tilde f_1,\tilde x) \ne \tilde \sigma_{\tilde C_1}^+$ then $\tilde \omega(\tilde f_2,\tilde x) = \tilde \sigma_{\tilde C_2}^+ = \tilde \sigma_{\tilde C_1}^-$. Symmetrically, if $\tilde \alpha (\tilde f_1,\tilde x) \ne \tilde \sigma_{\tilde C_1}^-$ then $\tilde \alpha(\tilde f_2,\tilde x) = \tilde \sigma_{\tilde C_2}^- = \tilde \sigma_{\tilde C_1}^+$.
\end{lemma}
\proof Let $ T_{\tilde \sigma}$ be the covering translation with axis $\tilde \sigma$ and orientation induced by $\tilde C_2$. Then $\tilde f_2^n = T_{\tilde \sigma}^{dn} \tilde f_1^n$ where $d>0$ is the degree of Dehn twisting about ${\cal R}$. By hypothesis and by Lemma~\ref{orientations on sigma}, $\omega (\tilde f_1, \tilde x) \ne \tilde \sigma_{\tilde C_1}^+= \tilde \sigma_{\tilde C_2}^- =T_{\tilde \sigma}^{-} $. Since $\tilde f_1^n(\tilde x)$ converges to $\omega(\tilde f_1, \tilde x)$ it follows that $T_{\tilde \sigma}^{dn} \tilde f_1^n(\tilde x) \to T^+_{\tilde \sigma}$. This proves that $\omega (\tilde f_2, \tilde x) = \tilde \sigma_{\tilde C_2}^+$.
\endproof
\begin{lemma} \label{bounded distance} There is a constant $D_1>0$ so that if $\dist(\tilde x,\tilde C) > D_1$ and $\tilde \sigma$ is the component of $\partial \tilde C$ closest to $\tilde x$, then at least one of $\tilde \alpha(\tilde f_{\tilde C}, \tilde x)$ and $ \tilde \omega(\tilde f_{\tilde C}, \tilde x)$ is an endpoint of $\tilde \sigma$.
\end{lemma}
\proof For any $\tilde x$ both $\alpha( f_{\tilde C}, \tilde x)$ and $\tilde
\omega(\tilde f_{\tilde C}, \tilde x)$ lie in $\Fix(f_{\tilde C})$, i.e. in
points of $S_\infty$ which correspond to ends of $\tilde C.$ Two of the
points of $\Fix(f_{\tilde C})$ are ends of $\tilde \sigma$ and all the
others lie on one side of $\tilde \sigma$. Consider the annular cover
$A_\sigma$ and the induced map $f_\sigma : A_\sigma \to A_\sigma$.
If the lemma is false then for any $D_1 >0$ we can find $\tilde x$ such
that
\begin{itemize}
\item $\tilde \sigma$ is the component of $\partial \tilde C$ closest to $\tilde x$
\item Neither $\alpha( f_\sigma, \hat x)$ nor $\omega(
f_\sigma, \hat x)$ is an endpoint of $\tilde \sigma.$
\item $\dist(\tilde x,\tilde C) > D_1.$
\end{itemize}
From the second item we conclude that both $\alpha( f_\sigma, \hat x)$ and $\omega(
f_\sigma, \hat x)$ lie in $\partial A_\sigma,$ and, in fact, in the same
component of $\partial A_\sigma$ since their lifts lie on the same
side of $\tilde \sigma.$ From the first and third item we conclude that the $\hat x$'s lie arbitrarily close to the other component of $\partial A_\sigma$ in contradiction to Corollary~\ref{PB}
applied to $f_\sigma.$
\endproof
\begin{lemma} \label{crossing arc} Suppose that $\tilde C_1$ and $\tilde C_2$ are domains with intersection $\tilde \sigma \subset \tilde {\cal R}$ and that $\tilde f_i =\tilde f_{\tilde C_i}$. Let $\partial_i A_{ \sigma}$ be the boundary component that contains points that lift into the closure of $\tilde C_i$. If neither $\alpha(\tilde f_1, \tilde x)$ nor $\omega(\tilde f_2, \tilde x)$ is an endpoint of $\tilde \sigma$ then $\alpha(f_\sigma, \hat x) \in \partial_1 A_{\sigma}$ and $\omega( f_\sigma, \hat x) \in \partial_2 A_{\sigma}$.
\end{lemma}
\proof This is an immediate consequence of the definitions.
\endproof
\begin{cor} \label{canonical lift} For all $\tilde x \in H$ exactly one of the following hold.
\begin{enumerate}
\item There is a unique domain $\tilde C$ such that $ \omega(f_{ \tilde C}, \tilde x)$ is not an endpoint of a component of $\partial \tilde C$.
\item There is a unique component $\tilde \sigma$ of $\tilde {\cal R}$ such that both $ \omega(f_{\tilde C_1}, \tilde x)$ and $\omega(f_{\tilde C_2}, \tilde x)$ are the
two endpoints of $\tilde \sigma$ where $\tilde C_1$ and $\tilde C_2$ are the two domains that contain $\tilde \sigma$ in their boundaries.
\end{enumerate}
\end{cor}
\proof Lemma~\ref{moving on} and the obvious induction argument imply that the two items are mutually exclusive and also imply the uniqueness parts of the items. It therefore suffices to find $ \tilde C$ satisfying (1) or $\tilde \sigma$ satisfying (2).
Choose a domain $\tilde C_1'$. If $ \omega(\tilde f _{\tilde C_1'}, \tilde x)$ is not an endpoint of a component of $\partial \tilde C_1'$ we are done. Otherwise, let $\tilde C_2'$ be the domain whose intersection with $\tilde C_1'$ is the component $\tilde \sigma_1$ of $\tilde {\cal R}$ that contains $ \omega(\tilde f _{\tilde C_1'}, \tilde x)$. If $\tilde \omega(\tilde C_2', \tilde x)$ is either not the endpoint of a component of $\partial \tilde C_2'$ or is an endpoint of $\tilde \sigma_1$ we are done. Otherwise, let $\tilde C_3'$ be the domain whose intersection with $\tilde C_2'$ is the component $\tilde \sigma_2$ of $\tilde {\cal R}$ that contains $ \omega(\tilde f _{\tilde C_2'}, \tilde x)$. Iterating this procedure we either reach the desired conclusion or produce domains $\tilde C_k'$ such that $\omega(\tilde f _{\tilde C_k'}, \tilde x)$ is an endpoint of $\tilde \sigma_{k} = \tilde C_k' \cap \tilde C'_{k+1}$. By Lemma~\ref{bounded distance}, $\alpha(\tilde f _{\tilde C_k'}, \tilde x)$ is an endpoint of $\tilde \sigma_{k-1}$ for all sufficiently large $k$. Let $ f_k : A_{\sigma_k} \to A_{\sigma_k}$ be the annulus map induced from $\tilde \sigma_k$ and let $\partial_- A_{\sigma_k}$ and $\partial_+ A_{\sigma_k}$ be the components of $\partial A_{\sigma_k}$ that contain points that lift into the closure of $\tilde C'_k$ and $\tilde C'_{k+1}$ respectively. Lemma~\ref{crossing arc} implies that $\alpha(f_k, \hat x) \in \partial_- A_{\sigma_k}$ and $\omega(f_k, \hat x) \in \partial_+ A_{\sigma_k}$.
Since ${\cal R}$ consists of a finite set of simple closed geodesics there
is an $l$ which is the first integer greater than $k$ such that $\sigma_l = \sigma_k$. Since $M$ has genus zero, the arc in $A_{\sigma_k}$ connecting $\alpha( f_k, \hat x)$ to $\omega( f_k, \hat x) $ and the arc connecting $\alpha(f_l, \hat x)$ to $\omega( f_l, \hat x) $ cross $\sigma_k$ in opposite directions in contradiction to Lemma~\ref{consistent crossing}. The process therefore terminates after finitely many steps.
\endproof
\begin{defn} If Corollary~\ref{canonical lift}(1) is satisfied then we say that $\tilde C$ is the {\em $\omega$-domain for $\tilde x$} and $\tilde f_{\tilde C}$ is the {\em $\omega$-lift for $\tilde x$}. Otherwise, Corollary~\ref{canonical lift}(2) is satisfied and we say that $\tilde C_1$ and $\tilde C_2$ are the {\em $\omega$-domains for $\tilde x$} and
$\tilde f_{\tilde C_1}$ and $\tilde f_{\tilde C_2}$ are the {\em $\omega$-lifts for $\tilde x$}.
\end{defn}
\begin{cor} \label{stays close} Let $D_1$ be the constant of Lemma~\ref{bounded distance}.
\begin{enumerate}
\item If $ \tilde C$ is the unique $\omega$-domain for $\tilde x$ then $\tilde f_{\tilde C}^n(\tilde x) \in N_{D_1}(\tilde C)$ for all sufficiently large $n$.
\item If $\tilde C_1$ and $\tilde C_2$ are $\omega$-domains for $\tilde x$ with intersection $\tilde \sigma \in \tilde {\cal R}$ then $\tilde f_{\tilde C_i}^n(\tilde x) \in N_{D_1}(\tilde C_1 \cup \tilde C_2)$ for $i=1,2$ and all sufficiently large $n$.
\end{enumerate}
\end{cor}
\proof If $ \tilde C$ is the unique $\omega$-domain for $\tilde x$ and (1) fails then there exist arbitrarily large $n$ such that $\tilde f_{\tilde C}^n(\tilde x) \not \in N_{D_1}(\tilde C)$. Since $\tilde f_{\tilde C}^n(\tilde x) \to \omega(\tilde f_{\tilde C}, \tilde x)$ we may assume that $\alpha(\tilde f_{\tilde C}, \tilde x)$ is not an endpoint of the component of $\partial \tilde C$ that is closest to $\tilde f_{\tilde C}^n(\tilde x)$. This contradicts Lemma~\ref{bounded distance} and so completes the proof of (1).
Assume the notation of (2) and that (2) fails for $i =1$; the $i=2$ case is symmetric. There exist arbitrarily large $n$ such that $\tilde f_{\tilde C_1}^n(\tilde x) \not \in N_{D_1}(\tilde C_1 \cup \tilde C_2)$. Lemma~\ref{bounded distance} implies that $\tilde \sigma$ is the component of $\partial \tilde C_1$ that is closest to $\tilde f_{\tilde C_1}^n(\tilde x)$ for all sufficiently large $n$. Since $\tilde f_{\tilde C_1}$ and $\tilde f_{\tilde C_2}$ differ by an iterate of $T_{\tilde \sigma}$ and since $T_{\tilde \sigma}$ preserves both $\tilde C_1$ and $\tilde C_2$, $\tilde f_{\tilde C_2}^n(\tilde x)\not \in N_{D_1}(\tilde C_1 \cup \tilde C_2)$ and $\tilde \sigma$ is not the the component of $\partial \tilde C_2$ that is closest to $\tilde f_{\tilde C_2}^n(\tilde x)$. This contradicts Lemma~\ref{bounded distance} and so completes the proof of (2).
\endproof
We record the following observation for easy reference.
\begin{lemma} \label{omega condition} If $\tilde f_{\tilde C}^k(\tilde x) \in N_D(\tilde C)$ for some $D > 0$ and all sufficiently large $k$ then $\tilde C$ is an $\omega$-domain for $\tilde x$.
\end{lemma}
\proof It suffices to show that if $\omega(\tilde f_{\tilde C},\tilde x)$ is an endpoint of $\tilde \sigma \in \tilde {\cal R}$ and $C'$ is the other domain whose frontier contains $\tilde \sigma$ then $\omega(\tilde f_{\tilde C'},\tilde x)$ is an endpoint of $\tilde \sigma$. The covering translation $T_{\tilde \sigma}$ preserves $N_D(\tilde C)$. Since $\tilde f_{\tilde C'}^k$ and $\tilde f_{\tilde C}^k$ differ by an iterate of $T_{\tilde \sigma}$, it follows that $\tilde f_{\tilde C'}^k(\tilde x) \in N_D(\tilde C)$ for all sufficiently large $k$ and hence that $\omega(\tilde f_{\tilde C'},\tilde x)$ is an endpoint of $\tilde \sigma$.
\endproof
\section{Domain Covers} \label{sec:domain covers}
We assume throughout this section that ${\cal R} \ne \emptyset$.
Let $\tilde C$ be a domain and let $C$ be its image in $S$. Recall that $\Stab(\tilde C)$ is the subgroup of covering
translations that preserve $\tilde C$ and that elements of
$\Stab(\tilde C)$ commute with $\tilde f_{\tilde C}$. We can not
restrict $f$ to $C$ because $C$ is not $f$-invariant and we can not
replace $f$ by an isotopic map that preserves $C$ because we might
lose the entropy zero property. Instead we lift to the $\pi_1(C)$
cover $\bar C$ of $S$. More precisely we make the following definitions.
\begin{defns}
Define $\bar C_{\core} \subset \bar C$ to be the quotient spaces
of $\tilde C \subset H$ by the action of $\Stab(\tilde C)$ and $ \bar f_C : \bar C
\to \bar C$ to be the homeomorphism induced by $\tilde f_{\tilde C}$.
Up to conjugacy, $ \bar f_C : \bar C \to \bar C$ is independent of
the choice of lift $\tilde C$ of $C$.
\end{defns}
\begin{snotn} Our convention will be that if
$\tilde x \in \tilde C$ then its image in $M$ is $x$ and its image in $\bar
C$ is $\bar x$.
\end{snotn}
Note that $\bar C_{\core}$ is homeomorphic to $C$ and (topologically) $\bar C$ is obtained from $ \bar C_{\core}$ by adding collar neighborhoods to each component of $\partial \bar C_{\core}$.
Note also that $\bar f_C$ is isotopic to the identity.
\vspace{.1in}
If $\tilde C$ is both an $\alpha$-domain and an $\omega$-domain for $\tilde x$ then we say that $\tilde C$ is a {\em home domain} for $\tilde x$. Denote the set of birecurrent points for $f$ by ${\cal B} (f)$ and the full pre-image of ${\cal B}(f)$ by $\tilde {\cal B}(f)$.
The following proposition, whose proof is delayed until the end of the section, is the main result of this section.
\begin{prop} \label{home lift} If $\tilde C$ is an $\omega$-domain for $\tilde x \in \tilde {\cal B}(f)$ then $\bar x \in {\cal B}(\bar f_C)$ and $\tilde C$ is a home domain for $\tilde x$. Moreover if $\tilde \omega(\tilde f_{\tilde C}, \tilde x)$ is an endpoint of $\tilde \sigma \in \tilde {\cal R}$ then $\tilde \alpha(\tilde f_{\tilde C}, \tilde x)$ is also an endpoint of $\tilde \sigma$.
\end{prop}
The following definition is key to the proof of Proposition~\ref{home lift}.
\begin{defn} \label{defn:near cycle} A covering translation $T : H \to H$ is a {\em near-cycle of period $m$ for $\tilde x \in H$ with respect to $\tilde f_{\tilde C}$} if there is a free disk $U$ for $f$ and a lift $\tilde U$ that contains $\tilde x$ such that $\tilde f_{\tilde C}^m(\tilde x) \in T(\tilde U)$. If $m$ is irrelevant then we simply say that {\em $T$ is a near-cycle for $\tilde x \in H$ with respect to $\tilde f_{\tilde C}$}.
\end{defn}
\begin{remark} \label{near cycle is open} It is an immediate
consequence of the definitions that if $T : H \to H$ is a
near-cycle of period $m$ with respect to $\tilde f_{\tilde C}$ for $\tilde x$
then it is also a near-cycle of period $m$ with respect to $\tilde f_{\tilde C}$
for all points in a neighborhood of $\tilde x$. Moreover, it is clear
that by shrinking the free disk $U$ slightly to $U_0,$
we may assume that $cl(U_0)$ is contained in a free disk and we still
have $\tilde f_{\tilde C}^m(\tilde x) \in T(\tilde U_0).$
\end{remark}
\begin{remark} A point $\tilde x \in H$ has at least one near cycle if and only if its image $x\in M$ is free disk recurrent.
\end{remark}
The following lemma is essentially the same as Lemma~10.5 of \cite{fh:periodic}. We reprove it here because our assumptions have changed.
\begin{lemma} \label{near cycle} If $T \in \Stab(\tilde C)$ is a near-cycle for $\tilde x \in H$ with respect to $\tilde f_{\tilde C}$ then $\alpha(\tilde f_{\tilde C}, \tilde x)$ and $ \omega(\tilde f_{\tilde C}, \tilde x)$ can not both lie in the same component of $S_{\infty} \setminus \{T^+,T^-\}$.
\end{lemma}
\proof If $T$ is parabolic let $\tilde \sigma$ be a horocycle preserved by $T$; otherwise let $\tilde \sigma$ be the axis of $T$. From $T \in \Stab(\tilde C)$ it follows that $\tilde \sigma$ is either an element of $\tilde {\cal R}$ or disjoint from $\tilde {\cal R}$. Let $f_\sigma :A_\sigma \to A_\sigma$ be as in
Definition~\ref{annulus cover}. Assume the notation of
Lemma~\ref{reducible} applied with $r=1$ and $\hat x_1$ the image of
$\tilde x$ in $A_\sigma$. The lifts $\tilde B_1$ and $\tilde B_1'$ of $\hat
B_1$ that contain $\tilde x$ and $T(\tilde x)$ respectively are disjoint and
$\tilde f$-invariant up to isotopy rel the orbits of $\tilde x$ and $T(\tilde
x)$. Lemma~8.7-(2) of \cite{fh:periodic}
implies that $\tilde B_1$ and $\tilde B_1'$ have
parallel orientations. The lemma now follows from the fact that the
endpoints of $\tilde B_1$ are $\alpha(\tilde f_{\tilde C}, \tilde
x)$ and $ \omega(\tilde f_{\tilde C}, \tilde x)$ and the
endpoints of $\tilde B_1' $ are $T\alpha(\tilde f_{\tilde C}, \tilde x)$ and $T
\omega(\tilde f_{\tilde C}, \tilde x)$. \endproof
\begin{lemma} \label{peripheral curve}Suppose $\tilde C$ is an $\omega$-domain for $\tilde x $, that $\omega(\tilde f_{\tilde C},\tilde x)$ is an endpoint of $\tilde \sigma \in \tilde {\cal R}$ and that $\bar x$ is $\bar f_C$-recurrent. Then every near cycle $T \in \Stab(\tilde C)$ for a point in the $\tilde f_{\tilde C}$-orbit of $\tilde x$ is hyperbolic with axis $\tilde \sigma$. \end{lemma}
\proof To simplify notation we write $\tilde f = \tilde f_{\tilde C}$. There is no loss in assuming that $T$ is a near cycle for $\tilde x$. Let $U$ be the free disk with respect to which $T$ is defined, let $\tilde U$ be the lift of $U$ containing $\tilde x$ and let $n$ satisfy $\tilde f^n(\tilde x) \in T(\tilde U)$. There is a neighborhood $x \in V \subset U$ such that $f^n(V) \subset U$. Let $\tilde V$ be the lift of $V$ contained in $\tilde U$.
If $\alpha(\tilde f, \tilde x)$ is an endpoint of $\tilde \sigma$ then Lemma~\ref{near cycle} and the fact that $\tilde \sigma \subset \partial \tilde C$ complete the proof. Suppose then that $\alpha(\tilde f, \tilde x)$ is not an endpoint of $\tilde \sigma$ and in particular, $\alpha(\tilde f, \tilde x) \ne \omega(\tilde f, \tilde x)$.
Since $\bar x$ is $\bar f$-recurrent, there exist $n_i \to \infty$ and
$S_i \in \Stab(\tilde C)$ such that $\tilde f^{n_i}(\tilde x) \in S_i(\tilde V)
\subset S_i(\tilde U)$. From $\tilde f^n(S_i \tilde V) = S_i \tilde f^n(\tilde V)
\subset S_iT(\tilde U)$ we see that $\tilde f^n(\tilde f^{n_i}(\tilde x)) \in
S_iTS_i^{-1}(S_i(\tilde U))$ and hence that $T_i = S_iTS_i^{-1}$ is a
near cycle for $\tilde f^{n_i}(\tilde x)$.
If $T$, and hence each $T_i$, is parabolic then each $T_i^{\pm} = \alpha( \tilde f, \tilde x)$ by Lemma~\ref{near cycle}. In this case the $T_i$'s are iterates of a single parabolic covering translation and there is a neighborhood of $\omega(\tilde f, \tilde x)$ that is moved off of itself by each $T_i$. This contradicts $\lim \tilde f^{n_i}(\tilde x) = \omega(\tilde f, \tilde x)$ and $\lim T_i(\tilde f^{n_i}(\tilde x)) = \lim (\tilde f^{n+n_i}(\tilde x)) = \omega(\tilde f, \tilde x)$.
We conclude that $T$ and $T_i$ are hyperbolic. Let $A_T$ be the axis of $T$ and $A_i = S_i(A_T)$ the axis of $T_i$.
We claim that $A_i \ne \tilde \sigma$. This is obvious if $A_T$ is not an element of $\tilde {\cal R}$ so we assume that $A_T$ is an element of $\tilde {\cal R}$ and that $\tilde \sigma = A_i = S_i(A_T)$ for some $S_i \in \Stab(\tilde C)$
and argue to a contradiction. Keeping in mind that $\tilde \sigma$ and $A_T$ are distinct components of the frontier of $\tilde C$, Lemma~\ref{near cycle} implies that $\alpha(\tilde f, \tilde x)$ is an endpoint of $A_T$. The axis of $S_i$ is contained in $\tilde C$ and is not $\tilde \sigma$ or $A_T$. It follows that the axis of $S_i$ is disjoint from $A_T$ and $\tilde \sigma$ and has no endpoints in common with either. Since $\tilde \sigma = S_i(A_T)$, the axis of $S_i$ does not separate $A_T$ from $\tilde \sigma$. This contradicts Lemma~\ref{near cycle} applied to the near cycle $S_i$ and so completes the proof that $A_i \ne \tilde \sigma$.
After passing to a subsequence we may assume that either the $A_i$'s are all the same or all different. In the former case, $T_i$ is independent of $i$ and there is a neighborhood of $\omega(\tilde f, \tilde x)$ that is moved off of itself by each $T_i$. As above this contradicts the fact that $T_i(\tilde f^{n_i}(\tilde x)) \to \omega(\tilde f, \tilde x)$. We may therefore assume that the $A_i$'s are distinct lifts of a closed curve in $M$ and hence, after passing to a subsequence, converge to some point $Q \in S_{\infty}$. If $Q \ne \omega(\tilde f, \tilde x)$ then there is a neighborhood of $\omega(\tilde f, \tilde x)$ that is moved off of itself by each $T_i$ and we have a contradiction. Thus $Q = \omega(\tilde f, \tilde x)$.
For sufficiently large $i$ the endpoints of $A_i$ are contained in a neighborhood of $\omega(\tilde f, \tilde x)$ that does not contain $\alpha(\tilde f, \tilde x)$ and does not contain the other endpoint of $\tilde \sigma$. Since $A_i $ is disjoint from $\tilde \sigma $, it does not separate $\alpha(\tilde f, \tilde x)$ from $\omega(\tilde f, \tilde x)$. This contradicts Lemma~\ref{near cycle} since neither $\alpha(\tilde f, \tilde x)$ nor $\omega(\tilde f, \tilde x)$ is an endpoint of $A_i$.
\endproof
\begin{lemma} \label{recurrence condition} Suppose that $U$ is a free disk, that $x \in U$ is recurrent [birecurrent] with respect to $f$ and that the set of lifts of $ U$ to $H$ that intersect $\{\tilde f_{\tilde C}^k(\tilde x): k \ge 0\}$ is finite up to the action of $\Stab(\tilde C)$. Then $\bar x \in \bar C$ is recurrent [birecurrent] with respect to $\bar f : \bar C \to \bar C$.
\end{lemma}
\proof The set of lifts of $ U$ to $H$ that intersect $\{\tilde f_{\tilde C}^k(\tilde x): k \ge 0\}$ is finite up to the action of $\Stab(\tilde C)$ if and only if the set of lifts of $U$ to $\bar C$ that intersect $\{\bar f_C^k(\bar x): k \ge 0\}$ is finite. We may therefore replace the former with the latter in the hypotheses of this lemma.
Suppose that $x$ is recurrent. We must prove that $\bar x$ is recurrent and that if $x$ is recurrent with respect to $f^{-1}$ then $\bar x$ is recurrent with respect to $\bar f^{-1}$.
Let $\bar U_1, \ldots, \bar U_m$ be the lifts of $U$ to $\bar C$ that intersect $\{\bar f_{C}^k(\bar x): k \ge 0\}$ and let $\bar x_j \in \bar U_j$ be the corresponding lifts of $x$. We may assume that $\bar x_1 = \bar x$. Choose a sequence $n_i \to \infty$ such that $f^{n_i}( x) \to x $ and such that each $f^{n_i}( x) \in U$. After passing to a subsequence we may assume that $\bar f^{n_i}_C(\bar x_1) \in \bar U_s$ where $s$ is independent of $i$. Then $\bar f^{n_i}_C(\bar x_1) \to \bar x_s$ and we are done if $s=1$. Otherwise we may assume that $s=2$. Since $\bar x_2$ is in the $\omega$-limit set of $\bar x_1$, each point in $\{\bar f_{C}^k(\bar x_2): k \ge 0\}$ that projects to $U$ is contained in some $\bar U_j$. We may therefore apply the previous argument with $\bar x_2$ in place of $\bar x_1$. After passing to a further subsequence we may assume that $\bar f^{n_i}_C(\bar x_2) \to x_t$ where $t \ne 2$ because $\bar f_C^{n_i}(\bar x_1)$ is the unique point in $\bar U_2$ that projects to $f^{n_i}(x)$. If $t=1$ then $\bar x_1$ is in the $\omega$-limit set of $\bar x_1$ and we are done. Otherwise we may assume $t=3$. After iterating this argument at most $m$ times, we have shown that $\bar x$ is recurrent.
From the recurrence of $\bar x$, it follows that a lift of $U$ to $\bar C$ intersects $\{\bar f_C^k(\bar x): k \ge 0\}$ if and only if it intersects $\{\bar f_C^k(\bar x): k \in \mathbb Z\}$. In particular, the set of lifts of $U$ to $\bar C$ that intersect $\{\bar f_C^{-k}(\bar x): k \ge 0\}$ is finite. If $x$ is recurrent with respect to $f^{-1}$ then by the above argument $\bar x \in \bar C$ is recurrent with respect to $\bar f^{-1} : \bar C \to \bar C$ as desired.
\endproof
\begin{remark}\label{cosets} If $\tilde U$ is a lift of a disk $U$ and $T_1,T_2$ are covering translations then $T_1(\tilde U)$ and $\tilde T_2(\tilde U)$ are in the same $\Stab(\tilde C)$-orbit if and only if $T_2 T_1^{-1} \in \Stab(\tilde C)$. Thus a collection of lifts $\{T_m(\tilde U)\}$ of $U$ is finite up to the action of $\Stab(\tilde C)$ is finite if and only if the $T_m$'s determine only finitely many right cosets of $\Stab(\tilde C)$.
\end{remark}
\noindent{\bf Proof of Proposition~\ref{home lift}} Let $U$ be a free disk that contains $x$ and let $\tilde U$ be the lift that contains $\tilde x$.
As a first case suppose that $\tilde \omega(\tilde f_{\tilde C}, \tilde x)$ is not
an endpoint of an element of $\tilde {\cal R}$. Lemma~\ref{stays close}-(1)
implies that for some $D$ and all $k \ge 0$, $\tilde f_{\tilde C}^k(\tilde x)
\in N_D(\tilde C)$ or equivalently, $\bar f_{C}^k(\bar x) \in N_D(\bar
C_{\core})$. Since $N_D(\bar C_{\core})$ is compact and by
Remark~(\ref{near cycle is open}) we may assume the closure of
$U$ is contained in a disk in $M$, it follows that $\{\bar
f_{C}^k(\bar x) \ |\ k\ge 0\}$ intersects only finitely many lifts of $U$.
Equivalently, $\{\tilde f_{\tilde C}^k(\tilde x): k \ge 0\}$
intersects only finitely many lifts of $ U$ to $H$ up to the action
of $\Stab(\tilde C)$. Lemma~\ref{recurrence condition} implies that $\bar
x $ is birecurrent under $\bar f_C$. It follows that $\bar
f_{C}^k(\bar x) \in N_D(\bar C_{\core})$ for all $k$ and hence that
$\tilde f_{\tilde C}^k(\tilde x) \in N_D(\tilde C)$ for all $k$. Lemma~\ref{omega
condition} applied to $\tilde f^{-1}$ implies that $\tilde C$ is a home
domain.
We assume now that $\omega(\tilde f, \tilde x)$ is an endpoint of $\tilde
\sigma \in \tilde {\cal R}$ and that $\tilde C_1$ and $\tilde C_2$ are the two
domains that contain $\tilde \sigma$ in their frontier. We will treat
$\tilde C_1$ and $\tilde C_2$ symmetrically and prove that the proposition
holds for $\tilde C = \tilde C_1$ and $\tilde C = \tilde C_2$. Denote $\tilde f_{\tilde
C_1}$ by $\tilde f_1$ and $\tilde f_{\tilde C_2}$ by $\tilde f_2$. When near
cycles are defined with respect to $\tilde f_i$ we refer to them as $\tilde
f_i$-near cycles. Let $S$ be a root-free covering translation with
axis $\tilde \sigma$. Lemma~\ref{stays close}-(2) implies that $\tilde
f_1^k(\tilde x), \tilde f_2^k(\tilde x) \in N_{D}(\tilde C_1 \cup \tilde C_2)$ for
some $D$ and all $k \ge 0$. We may assume without loss that $\tilde U
\subset N_{D}(\tilde C_1) \cap N_{D}(\tilde C_2)$.
After interchanging $\tilde C_1$ with $\tilde C_2$ if necessary, we may assume by Lemma~\ref{moving on} that $\alpha(\tilde f_1, \tilde x)$ is an endpoint of $\tilde \sigma$. Lemma~\ref{near cycle} implies that every $\tilde f_1$-near cycle $T \in \Stab(\tilde C_1)$ for a point in the $\tilde f_1$-orbit of $\tilde x$ is an iterate of $S$. We will apply this as follows. If $T_1$ and $T_2$ are $\tilde f_1$-near cycles for $\tilde x$ and if $T_1 T_2^{-1}$ (which is a near cycle for a point in the $\tilde f_1$-orbit of $\tilde x$) is an element of $\Stab(\tilde C_1)$ then $T_1 T_2^{-1}$ is an iterate of $S$. In particular, if $T_1$ and $T_2$ determine the same right coset of $\Stab(\tilde C_1)$ then they also determine the same right coset of $\Stab(\tilde C_2)$,
Let ${\cal U}_i$ be the set of lifts of $U$ that intersect $N_{D}( \tilde C_i)$ and contain $\tilde f_2^k(\tilde x) $ for some $k \ge 0$. To prove that $\bar x$ is $\bar f_2$-birecurrent it suffices by Lemma~\ref{recurrence condition} to prove that ${\cal U}_1 \cup {\cal U}_2$ is finite up to the action of $\Stab(\tilde C_2)$. As above, the compactness of $N_D(\bar C_{i_{\core}})$ implies that ${\cal U}_i$ is finite up to the action of $\Stab(\tilde C_i)$.
Each element of ${\cal U}_i$ has the form $T(\tilde U)$ for some covering translation $T$; let ${\cal T}_i$ be the set of all such $T$. Each $T_m \in {\cal T}_1$ is an $\tilde f_2$-near cycle for $\tilde x$. Since $\tilde f_2$ and $\tilde f_1$ differ by an iterate of $S$, there exists $j_m$ such that $S^{j_m}T_m$ is an $\tilde f_1$-near cycle for $\tilde x$. Remark~\ref{cosets} implies that $\{S^{j_m}T_m\}$ determines only finitely many right cosets of $\Stab(\tilde C_1)$ which, as observed above, implies that $\{S^{j_m}T_m\}$ determines only finitely many right cosets of $\Stab(\tilde C_2)$. Since $\{S^{j_m}T_m\}$ and $\{T_m\}$ determine the same right cosets of $\Stab(\tilde C_2)$, we have shown that $\{T_m\}$ determines only finitely many right cosets of $\Stab(\tilde C_2)$. A second application of Remark~\ref{cosets} completes the proof that $\bar x$ is $\bar f_2$-birecurrent.
Since $x$ is recurrent with respect to $f^{-1}$ there are near cycles $T'_j$ with respect to $\tilde f_2^{-1}$ for $\tilde x$ such that $T'_j(\tilde x) \to \alpha(\tilde f_2, \tilde x)$. From the fact that ${T'_j}^{-1}$ is a near cycle with respect to $\tilde f_2$ for a point in the orbit of $\tilde x$ and hence is an iterate of $S$ we conclude that $\alpha(\tilde f_2, \tilde x)$ is also an endpoint of $\tilde \sigma$. This completes the proof for $\tilde C_2$.
Now that we have established that $\alpha(\tilde f_2, \tilde x)$ is an endpoint of $\tilde \sigma$, this same argument can be applied to $\tilde C_1$.
\qed
We conclude this section by strengthening Corollary~\ref{stays close}.
\begin{cor} \label{symmetric close} Suppose that $x \in {\cal B}(f)$ and that $D_1$ is the constant of Lemma~\ref{bounded distance}.
\begin{enumerate}
\item If $ \tilde C$ is the unique home domain for $\tilde x$ then $\tilde f_{\tilde C}^n(\tilde x) \in N_{D_1}(\tilde C)$ for all $n$.
\item If $\tilde C_1$ and $\tilde C_2$ are home domains for $\tilde x$ with intersection $\tilde \sigma \in \tilde {\cal R}$
then $\tilde f_{\tilde C}^n(\tilde x) \in N_{D_1}(\tilde C_1 \cup \tilde C_2)$ for all $n$.
\end{enumerate}
\end{cor}
\proof Suppose that $\tilde C$ is the unique home domain for $\tilde x$ and that $\tilde f_{\tilde C}^n(\tilde x) \not \in N_{D_1}(\tilde C)$. Choose $\epsilon$ greater than the distance from $\tilde f_{\tilde C}^n(\tilde x) $ to $N_{D_1}(\tilde C)$. Proposition~\ref{home lift} implies that $\bar x \in {\cal B}(\bar f_{C})$ and hence that there exist arbitrarily large $k$ and $S_k \in \Stab(\tilde C)$ such that the distance from $\tilde f_{\tilde C}^k(\tilde x)$ to $S_k\tilde f_{\tilde C}^n(\tilde x)$ is less than $\epsilon$. Since $S_k$ preserves distance to $\tilde C$, $\tilde f_{\tilde C}^k(\tilde x) \not \in N_{D_1}(\tilde C)$. This contradicts Corollary~\ref{stays close} and so completes the proof of (1).
Assuming the notation of (2), suppose that $\tilde f_{\tilde C}^n(\tilde x) \not \in N_{D_1}(\tilde C_1 \cup \tilde C_2)$. There is no loss in assuming that $\tilde f_{\tilde C}^n(\tilde x)$ is closer to $\tilde C_1$ than $\tilde C_2$. If $S_k \in \Stab(\tilde C_1)$ then $S_k\tilde f_{\tilde C}^n(\tilde x)$ is closer to $\tilde C_1$ than $\tilde C_2$ and has distance greater than $D_1$ from $\tilde C_1$. The argument given for (1) therefore applies in this context as well.
\endproof
\section{Two compactifications}\label{sec: fh-periodic}
We now return to the general case, allowing the possibility that ${\cal R} = \emptyset$.
If ${\cal R} = \emptyset$ then $H$ is the only domain and $\tilde f_{H}$ commutes with all covering translations and fixes every point in $S_{\infty}$. For consistency of notation we define $H$ to be the home domain for each $\tilde x$, define $\bar M$ to be $M$ and define $\bar f: \bar C \to \bar C$ to be $f : M \to M$.
Suppose that $\tilde C$ is the home domain for $\tilde x$. The diffeomorphism $\bar f: \bar C \to \bar C$ is isotopic to the identity and so satisfies one of the two hypotheses needed to apply results from sections 10 and 11 of \cite{fh:periodic}. The other hypothesis, that $f$ have no periodic points, is used only to prove Lemma~10.8 of \cite{fh:periodic} (which is then applied in other proofs). We proved in Lemma~\ref{lemma:fitted family} that $\bar f: \bar C \to \bar C$ satisfies the conclusions of Lemma~10.8 of \cite{fh:periodic} and hence that the results of \cite{fh:periodic} apply to $\bar f: \bar C \to \bar C$.
The one subtlety in applying these results is that when ${\cal R} \ne \emptyset$, two different compactifications of the universal cover of $\bar C$ are being used.
In the {\em extrinsic compactification}, the universal cover of $\bar C$ is metrically identified with the universal cover $\tilde M$ of $M$, which is metrically identified with $H$ and is compactified by $S_{\infty}$. The covering translations of the universal cover of $\bar C$ are identified with the subgroup $\Stab(\tilde C)$ of covering translations of the universal cover of $M$; the closure in $S_{\infty}$ of the fixed points of the elements of $\Stab(\tilde C)$ is a Cantor set $K$ whose
convex hull projects to $C_{core} \subset \bar C$.
In the {\em intrinsic compactification}, $\bar C$ is viewed without regard to $M$ and is equipped with a hyperbolic structure in which the ends corresponding to the components of $\partial C$ are cusps. The universal cover of $\bar C$ is then metrically identified with $H$ and compactified with $S_{\infty}$. In this case, the set of fixed points of covering translations is dense in $S_{\infty}$. Topologically the intrinsic compactification of the universal cover is obtained from the extrinsic compactification by collapsing the closure of each component of $S_{\infty} \setminus K$ to a point.
We have defined $\bar C$ using the extrinsic metric so that geodesics in $\bar C_{\core}$ correspond exactly to geodesics in $C \subset M$. If one considers $\bar f: \bar C \to \bar C$ as a homeomorphism of a punctured surface without reference to $M$ and applies results from \cite{fh:periodic} then the intrinsic metric is used. To help separate the two, write $g: N \to N$ for $\bar f: \bar C \to \bar C$ when $\bar C$ has the intrinsic metric. Since $g$ is isotopic to the identity there is a preferred lift
$\tilde g : \tilde N \to \tilde N$ to the universal cover that commutes with all covering translations. The \lq identity map\rq\ $p: \tilde M \to \tilde N$ conjugates $\tilde f_{\tilde C} : \tilde M \to \tilde M$ to $\tilde g : \tilde N \to \tilde N$. The homeomorphism $p$, which is not an isometry, extends over the compactifying circles but not by a homeomorphism; it collapses the closure of each component of $S_{\infty} \setminus K$ to a point. In particular, $p|_K$ identifies a pair of points if and only if they bound a component of $\partial \tilde C$. The relevance to us is that if $\omega(\tilde f_{\tilde C}, \tilde x)$ and $\alpha(\tilde f_{\tilde C}, \tilde x)$ bound a component of $\partial \tilde C$ then $\omega(\tilde g, p(\tilde x)) =\alpha(\tilde g, p(\tilde x))$.
We now turn to the applications.
\begin{defn} \label{defn:iterates about 2} Assume that $\tilde C$ is a home domain for a lift $\tilde x$ of $x \in M$. A point $P \in S_{\infty}$ {\em projects to an isolated puncture $c$} in $M$ if some (and hence every) ray in $\tilde M$ that converges to $P$ projects to a ray in $M$ that converges to $c$. We say that {\em $x$ rotates about an isolated puncture $c$} if for some (and hence all) lifts $\tilde x \in H$ there exists $P \in S_{\infty}$ that projects to $c$ and a parabolic covering translation $T$ that fixes $P$ such that every near cycle $S \in \Stab(\tilde C)$ for every $\tilde f_{\tilde C}^k(\tilde x)$ is a positive iterate of $T$.
\end{defn}
\begin{lemma} \label{isolated puncture}Suppose that $x \in {\cal B}(f)$, that $\tilde C$ is a home domain for a lift $\tilde x$ and that $\alpha(\tilde f_{\tilde C}, \tilde x)\ = \omega(\tilde f_{\tilde C}, \tilde x) =P$. Then $P$ projects to an isolated puncture $c$ and $x$ rotates about $c$.
\end{lemma}
\proof If ${\cal R} =\emptyset$ then this is Lemma 11.2 of \cite{fh:periodic}.
Suppose then that ${\cal R} \ne \emptyset$ and assume notation as above. Since $\alpha(\tilde f_{\tilde C}, \tilde x)\ = \omega(\tilde f_{\tilde C}, \tilde x) =P$, it follows that $\alpha(\tilde g, p(\tilde x))= \omega(\tilde g, p(\tilde x)) =p(P)$. By Lemma~11.2 of \cite{fh:periodic}, $p(P)$ projects to an isolated puncture $c'$ in $\bar C$ and there is a parabolic covering translation $T'$ that fixes $p(P)$ such every near cycle for every point in the orbit of $p(\tilde x)$ is a positive iterate of $T'$. If $c'$ does not correspond to an end of $\bar C$ determined by a component of $\partial C$ then we are done.
To complete the proof we assume that $c'$ corresponds to an end of $\bar C$ determined by a component of $\partial C$ and argue to a contradiction. In this case the parabolic $T': \tilde N \to \tilde N$ corresponds to a hyperbolic $T \in \Stab(\tilde C)$ and every near cycle $S \in \Stab(\tilde C)$ for every point in the $\tilde f_{\tilde C}$-orbit of $\tilde x$ is a positive iterate of $T$. Since (Proposition~\ref{home lift}) $\bar x \in {\cal B}(\bar f)$, there are near cycles $S^+_i\in \Stab(\tilde C)$ such that $S^+_i(\tilde x) \to \omega(\tilde f_{\tilde C}, \tilde x) = P$ and near cycles $S^-_i$ such that $S^-_i(\tilde x) \to \alpha(\tilde f_{\tilde C}, \tilde x)=P$. Since $S^+_i$ is a positive iterate of $T$ and $S^-_i$ is a negative iterate of $T$, this is impossible.
\endproof
\begin{defn} If $\tilde C$ is a home domain for $\tilde x \in \tilde M$ and $\alpha(\tilde f_{\tilde C}, \tilde x)\ \ne \omega(\tilde f_{\tilde C}, \tilde x)$ let $\tilde \gamma(\tilde x, \tilde C)$ be the oriented geodesic with endpoints $\alpha(\tilde f_{\tilde C}, \tilde x)$ and $\omega(\tilde f_{\tilde C}, \tilde x)$; we say that {\em $\tilde x$ tracks $\tilde \gamma(\tilde x, \tilde C)$ under iteration by $\tilde f_{\tilde C}$}. Let $\gamma(x) \subset M$ be the unoriented geodesic that is the projected image of $\tilde \gamma(\tilde x)$; we say that {\em $x$ tracks $\gamma(x)$}. Note that $\gamma(x)$ is independent of the choice of lift $\tilde x$ and the choice of home lift for $\tilde x$; the latter would not be true if we imposed an orientation on $\gamma(x)$.
\end{defn}
\begin{lemma} \label{disjoint scc} If $x \in {\cal B}(f)$ tracks $\gamma(x)$ then $\gamma(x)$ is a simple closed curve. If in addition $y \in {\cal B}(f)$ tracks $\gamma(y)$ then $\gamma(x)$ and $\gamma(y)$ are either disjoint or equal.
\end{lemma}
\proof If ${\cal R} =\emptyset$ then this follows from Lemmas ~10.2 and 11.6 of \cite{fh:periodic} and the fact that $M$ has genus zero.
Suppose then that ${\cal R} \ne \emptyset$. Choose a lift $\tilde x$ and home domain $\tilde C$ for $\tilde x$. We may assume without loss that $\tilde C$ is also a home domain for a lift $\tilde y$ of $y$. By Lemma~\ref{home lift}, $\bar x, \bar y \in {\cal B}(\bar f_C)$.
Assume notation as in the beginning of this section. If $\alpha(\tilde g, p(\tilde x))= \omega(\tilde g, p(\tilde x))$ then
$\alpha(\tilde f_{\tilde C}, \tilde x)$ and $\omega(\tilde f_{\tilde C}, \tilde x)$ bound a component of $\partial \tilde C$. In this case $ \gamma(x) \in {\cal R}$ and the lemma is obvious. We may therefore assume that $\alpha(\tilde g, p(\tilde x)) \ne \omega(\tilde g, p(\tilde x))$ and similarly for $\tilde y$. As in the ${\cal R} =\emptyset$ case, the lemma follows from Lemmas ~10.2 and 11.6 of \cite{fh:periodic} and the fact that $M$ has genus zero.
\endproof
\begin{cor} \label{limited near cycles} Suppose that $x\in {\cal B}(f)$, that $\tilde C$ is a home domain for a lift $\tilde x$ and that $\tilde x$ tracks $\tilde \gamma$. Then every $\tilde f_{\tilde C}$-near cycle $S \in \Stab(\tilde C)$ for a point in the orbit of $\tilde x$ is an iterate of $T_{\tilde \gamma}$.
\end{cor}
\proof We may assume without loss that $S$ is a near cycle for $\tilde x$. There exist $m \ne 0$ and a lift $\tilde U$ of a free disk $U \subset M$
such that $\tilde x \in \tilde U$ and $\tilde f_{\tilde C}^m(\tilde x) \in S(\tilde U)$. Let $\tilde y = S(\tilde x)$. Since $S \in \Stab(\tilde C)$, $S$ commutes with $\tilde f_{\tilde C}$. Thus $S (\tilde \gamma) \subset \tilde C$ is the oriented geodesic connecting $\alpha(\tilde f_{\tilde C}, \tilde y)$ to $ \omega(\tilde f_{\tilde C}, \tilde y)$. Lemma~\ref{disjoint scc} implies that $\tilde \gamma$ and $S(\tilde \gamma)$ are disjoint or equal (up to perhaps a change of orientation). In the latter case we are done so we assume the former and argue to a contradiction. Since $\tilde y$ and $ \tilde f_{\tilde C}^m(\tilde x)$ are contained in the free disk $\tilde U$, Lemma 8.7(2) of \cite{fh:periodic} implies that the $\tilde \gamma$ and $S(\tilde \gamma)$ are parallel. Since $M$ has genus zero there is an anti-parallel translate $S'(\tilde \gamma)$ that separates $\tilde \gamma$ and $S(\tilde \gamma)$. We have $S' \in \Stab(\tilde C)$ because $S'(\tilde \gamma) \subset \tilde C$. Thus $S' (\tilde \gamma)$ is the oriented geodesic connecting $\alpha(\tilde f_{\tilde C}, \tilde z)$ to $ \omega(\tilde f_{\tilde C}, \tilde z)$ for $\tilde z = S'(\tilde x)$. This contradicts Lemmas 8.9 and 8.10 of \cite{fh:periodic} and so completes the proof.
\endproof
\section{The Set of Annuli ${\cal A}$} \label{sec: annuli}
\begin{defns} Let $\Gamma$ be the set of simple closed curves that are tracked by at least one element of ${\cal B}(f)$.
For each lift $\tilde \gamma$ of $\gamma \in \Gamma$, choose a domain
$\tilde C$ that contains $\tilde \gamma$ and let $\tilde U(\tilde \gamma)$ be the
set of points in $H$ which have a neighborhood $\tilde V$
such that every point in $\tilde V \cap \tilde {\cal B}(f)$ tracks $\tilde \gamma$. We say that $\tilde C$ is a {\em home domain} for $\tilde U(\tilde \gamma)$, that $\tilde \gamma$ is the {\em defining parameter} of $ \tilde U(\tilde \gamma)$ and that $T_{\tilde \gamma}$ is {\em the covering translation associated to $\tilde U(\tilde \gamma)$}.
For each $\gamma \in \Gamma$ define $U(\gamma)$ to be the projected image of $\tilde U(\tilde \gamma)$ for any lift $\tilde \gamma$. We say that $C$ is a {\em home domain} for $U(\gamma)$ and that $\gamma$ is the {\em defining parameter} of $U(\gamma)$.
\end{defns}
\begin{remark} As the notation suggests, $\tilde U(\tilde \gamma)$ depends only on $\tilde \gamma$ and not on the choice of $\tilde C$. Indeed, if $\tilde C$ is not unique then $\tilde \gamma \in \tilde {\cal R}$ and every element of $\tilde V \cap \tilde {\cal B}(f)$ has exactly two home domains $C$ and $C'$ (where $C'$ is the other domain that contains $\tilde \gamma$) and both $\{\alpha(\tilde f_{\tilde C}, \tilde z), \omega(\tilde f_{\tilde C}, \tilde z)\}$ and $ \{\alpha(\tilde f_{\tilde C'}, \tilde z), \omega(\tilde f_{\tilde C'}, \tilde z)\}$ are contained in $\{\tilde \gamma^\pm\}$. $U(\gamma)$ is well defined because $\tilde U(S(\tilde \gamma)) = S\tilde U(\tilde \gamma)$ for any covering translation $S$.
\end{remark}
\begin{defns}\label{defn: A}
Let ${\cal C}$ be the set of punctures in $M$ for which there is at least
one element of ${\cal B}(f)$ that rotates about $c$. For each $P \in
S_{\infty}$ that projects to $c \in {\cal C}$, let $\tilde C$ be the unique domain whose closure contains
$P$ and let $\tilde U(P)$ be the set of points in $H$ for which
there is a neighborhood $\tilde V$ such that every point in $\tilde V \cap
\tilde {\cal B}(f)$ rotates about $P$. We say that $\tilde C$ is the {\em home
domain} for $\tilde U(P)$, that $P$ is the {\em defining parameter} of
$\tilde U = \tilde U(P)$ and that $T_{P}$ is {\em the covering translation
associated to $\tilde U(P)$}.
For each $c \in {\cal C}$ define $U(c)$ to be the projected image of $\tilde U(P)$ for any puncture $P$ that projects to $c$. We say that $C$ is the {\em home domain} for $U(c)$ and that $c$ is the {\em defining parameter} of $U(\gamma)$. As in the previous remark, $U(c)$ is well defined.
Let $\tilde {\cal A}$ be the set of all $\tilde U(\tilde \gamma)$'s and $\tilde U(P)$'s
and let
\[
\tilde {\cal U} = \bigcup_{\tilde \gamma} \tilde U(\tilde \gamma) \cup \bigcup_P \tilde U(P).
\]
Let ${\cal A}$ be the set of all $U(\gamma)$'s and $U(c)$'s and let
${\cal U}$ be the projection of $\tilde {\cal U}$ into $M$.
\end{defns}
\bigskip
\begin{lemma} \label{U tilde properties}
\begin{enumerate}
\item Each $\tilde U \in \tilde {\cal A}$ is open and invariant by both $T$ and $\tilde f_{\tilde C}$ where $\tilde C$ is a home domain for $\tilde U$ and $T$ is the covering translation associated to $\tilde U$.
\item If $\tilde U, \tilde U' \in \tilde {\cal A}$ have different defining parameters then $\tilde U \cap \tilde U' = \emptyset$.
\item If $\tilde U \in\tilde {\cal A}$ and $S$ is a covering translation then $S(\tilde U) \cap \tilde U \ne \emptyset$ if and only if $S$ is an iterate of the covering translation associated to $\tilde U$.
\item Each $U \in {\cal A}$ is open and $f$-invariant; if $U_1$ and $U_2$ have different defining parameters then $U_1 \cap U_2 = \emptyset$.
\end{enumerate}
\end{lemma}
\proof (1) and (2) are immediate from the definitions. (3) follows from (2) and the fact that $S$ maps the defining parameter for $\tilde U$ to the defining parameter for $S(\tilde U)$. (4) follows from (1) - (3).
\endproof
\begin{remark} \label{second centralizer invariance} If $h:M \to M$
commutes with $f$ then $h$ permutes the elements of ${\cal A}$.
We observed in Remark~(\ref{first centralizer invariance}) that
this is an easy consequence of part (1) of Theorem~(\ref{thm: annuli}).
However, we wish to use this fact before proving part (1) of that
theorem. Therefore we observe that it also follows from what we
have shown about the elements of ${\cal A}$ and their defining parameters.
To see this, suppose that $\tilde C$ is a home domain for $\tilde x \in H$ and
that $\tilde x$ tracks $\tilde \gamma$ under iteration by $\tilde f_{\tilde C}$.
Choose a lift $\tilde h :H \to H$ of $h$. Then $\tilde h \tilde f_{\tilde
C} \tilde h^{-1} = \tilde f_{\tilde C'}$ for some domain $\tilde C'$ that is a
home domain for $\tilde h(\tilde x)$ and $\tilde h(\tilde x)$ tracks $\tilde
h_{\#}(\tilde \gamma)$ under iteration by $\tilde f_{\tilde C'}$. This proves
that $h(U(\gamma)) = U(h_\#(\gamma))$. A similar argument applies to
the $U(P)$'s.
\end{remark}
To further analyze the elements of ${\cal A}$, we need the following observations.
\begin{lemma}\label{not home} \begin{enumerate} \item If $\tilde C$ is not a home domain for $\tilde y \in \tilde B(f)$ then $\alpha(\tilde f_{\tilde C}, \tilde y)$ and $\omega(\tilde f_{\tilde C}, \tilde y)$ are both endpoints of the component of $\partial \tilde C$ that is closest to the home domain for $\tilde y$.
\item If $\tilde y \in \tilde B(f)$, $\tilde C$ is any domain and either $\alpha(\tilde f_{\tilde C} \tilde y)$ or $\omega(\tilde f_{\tilde C}, \tilde y)$ is an endpoint of a frontier component $\tilde \sigma$ of $\tilde C$ then both $\alpha(\tilde f_{\tilde C} \tilde y)$ and $\omega(\tilde f_{\tilde C}, \tilde y)$are endpoints of $\tilde \sigma$. \end{enumerate}
\end{lemma}
\proof Item (1) follows from the existence of a home domain for $\tilde y$, Lemma~\ref{moving on} and the obvious induction argument on the number of domains that separate $\tilde C$ from a home domain for $\tilde y$. Item (2) follows from (1) if $\tilde C$ is not a home domain and from Lemma~\ref{isolated puncture} and Lemma~\ref{home lift} otherwise.
\endproof
We next show that ${\cal B}(f) \subset {\cal U}$.
\begin{lemma} \label{U covers} If either $\alpha(f,y) \ne \emptyset$ or $\omega(f,y) \ne \emptyset$ then $y$ is contained in an element $U$ of ${\cal A}$. In particular, each $y \in {\cal B}(f)$ is contained in some $U \in{\cal A}$.
\end{lemma}
\proof The two cases are symmetric so we may assume that $\omega(f,y) \ne \emptyset$. Choose $z \in \omega(f,y)$ and a free disk neighborhood $V$ of $z$. After replacing $y$ by some $f^k(y)$, we may assume that $y \in V$. Since $z \in \omega(f,y)$ there exist $m_i \to \infty$ such that $f^{m_i}(y) \to z$ and such that each $f^{m_i}(y) \in V$. Choose lifts $\tilde y, \tilde z \in \tilde V$.
By Corollary~\ref{symmetric close}, the distance between a point in $ \tilde B(\tilde f)$ and a home domain for that point is uniformly bounded. It follows that there are only finitely many home domains for elements $\tilde x_l \in \tilde B(f) \cap \tilde V$ and so we may choose a sequence $\tilde x_l \to \tilde y$ all of which have the same home domain(s) $\tilde C$ and $\tilde C'$, where we allow the possibility that $\tilde C = \tilde C'$. By Corollary~\ref{symmetric close} the distance between $\tilde f_{\tilde C}^{m_i}(\tilde x_l)$ and $\tilde C \cup \tilde C'$ is uniformly bounded. It follows that the distance between $\tilde f_{\tilde C}^{m_i}(\tilde y)$ and $\tilde C \cup \tilde C'$ is uniformly bounded. After passing to a subsequence of the $m_i$'s and interchanging $\tilde C$ and $\tilde C'$ if necessary, we may assume that the distance between $\tilde f_{\tilde C}^{m_i}(\tilde y)$ and $\tilde C$ is uniformly bounded.
Let $S_i$ be the covering translation such that $\tilde f_{\tilde C}^{m_i}(\tilde y) \in S_i(\tilde V)$ and note that the distance between $S_i(\tilde z)$ and $\tilde C$ is uniformly bounded. Up to the action of $\Stab(\tilde C)$, the number of translates of $\tilde z$ that have uniformly bounded distance from $\tilde C$ is finite. We may therefore choose $k>j$ such that $S = S_kS_j ^{-1}\in \Stab(\tilde C)$. Let $\tilde W = S_j(\tilde V)$ and let $\tilde W '\subset \tilde W$ be a neighborhood of $\tilde f_{\tilde C}^{m_j}(\tilde y)$ such that $\tilde f^{m_k-m_j}(\tilde W') \subset S(\tilde W)$. Then $S$ is a $\tilde f_{\tilde C}$-near cycle for every point in $\tilde W'$ and in particular for $\tilde f_{\tilde C}^{m_j}(\tilde x_l) $ for all sufficiently large $l$. Choose such an $\tilde f_{\tilde C}^{m_j}(\tilde x_l) $ and denote it simply by $\tilde x$.
To prove that $\tilde f_{\tilde C}^{m_j}(\tilde y)$, and hence $\tilde y$, is contained in an element of $\tilde U$ with home domain $\tilde C$ it suffices to show that if $\tilde w \in \tilde B(f) \cap \tilde W'$ then $\tilde C$ is a home domain for $\tilde w$ and $\{\alpha(\tilde f_{\tilde C}, \tilde x),\omega(\tilde f_{\tilde C}, \tilde x) \} = \{\alpha(\tilde f_{\tilde C}, \tilde w),\omega(\tilde f_{\tilde C}, \tilde w)\} $.
We proceed with a case analysis. As a first case suppose that $\tilde x$ tracks a geodesic $\tilde \gamma(\tilde x)$. Corollary~\ref{limited near cycles} implies that $S$ is an iterate of $T_{\tilde \gamma(\tilde x)}$. As a first subcase suppose that $\tilde C$ is a home domain for $\tilde w$. Since $T_{\tilde \gamma(\tilde x)}$ is not parabolic, Lemma~\ref{isolated puncture} implies that $\tilde w$ tracks some geodesic $\tilde \gamma(\tilde w)$. Corollary~\ref{limited near cycles} implies that $\tilde \gamma(\tilde w) = \tilde \gamma(\tilde x)$ as desired.
The remaining subcase is that $\tilde C$ is not a home domain for $\tilde w$. Lemma~\ref{not home} implies that $\{\alpha(\tilde f_{\tilde C}, \tilde w),\omega(\tilde f_{\tilde C}, \tilde w)\}$ is contained in the set of endpoints for some $\tilde \sigma$ in the frontier of $\tilde C$. Lemma~\ref{near cycle} then implies that $\tilde \sigma = \tilde \gamma(\tilde x)$. Let $\tilde C'$ be the other domain that contains $\tilde \gamma(\tilde x)$. Since some iterate of $T_{\tilde \gamma(\tilde x)}$ is a near cycle for $\tilde w$ with respect to $\tilde f_{\tilde C}$, the same is true with respect to $\tilde f_{\tilde C'}$. Lemma~\ref{near cycle} implies that $\{\alpha(\tilde f_{\tilde C'}, \tilde w),\omega(\tilde f_{\tilde C'}, \tilde w) \} \cap \{\tilde \gamma^\pm(\tilde x)\} \ne \emptyset$ and Lemma~\ref{not home} implies that both $\alpha(\tilde f_{\tilde C'}, \tilde w)$ and $\omega(\tilde f_{\tilde C'}, \tilde w) $ are endpoints of $\tilde \gamma(\tilde x)$. This contradicts the assumption that $\tilde C$ is not home domain for $\tilde w$ and so proves that the second subcase never occurs.
By Lemma~\ref{isolated puncture}, the only remaining case is that $\alpha(\tilde f_{\tilde C}, \tilde x) = \omega(\tilde f_{\tilde C}, \tilde x) =P$ and that $S$ is an iterate of $T_P$. Lemma~\ref{near cycle} implies that $P \in \{\alpha(\tilde f_{\tilde C'}, \tilde w),\omega(\tilde f_{\tilde C'}, \tilde w) \}$, Lemma~\ref{not home} implies that $\tilde C$ is a home domain for $\tilde w$ and Lemma~\ref{disjoint scc} implies that $\alpha(\tilde f_{\tilde C'}, \tilde w) = \omega(\tilde f_{\tilde C'}, \tilde w) = P$. \endproof
\begin{lemma} \label{more home domains} Let $Y = M \setminus {\cal U}$ and let $\tilde Y \subset H$ be the full pre-image of $Y$.\begin{enumerate}
\item For each $\tilde y \in \tilde Y $ there is a domain $\tilde C$ that is the unique $\alpha$-domain, unique $\omega$-domain and unique home domain for $\tilde y$.
Moreover, $\tilde C$ is a home domain for all points in a neighborhood of $\tilde y$.
\item If $\tilde C$ is the home domain for $\tilde y \in \tilde Y$ then $\tilde y$ has no $\tilde f_{\tilde C}$-near cycles in $\Stab(\tilde C)$.
\item For any compact subset $X \subset M$ there is a constant $K_X$ such that for each $y \in Y$, $ f^i(y) \in X$ for at most $K_X$ values of $i$.
\item There exists $\epsilon > 0$ so that if $\tilde y_1, \tilde y_2 \in \tilde Y $ and $\dist(\tilde y_1, \tilde y_2) < \epsilon$ then $\tilde y_1$ and $\tilde y_2$ have the same home domain.
\end{enumerate}
\end{lemma}
\proof Suppose at first that ${\cal R} = \emptyset$. Items (1) and (4) are obvious. Every neighborhood of $\tilde y$ contains points in $\tilde B(f)$ that are contained in distinct elements of ${\cal U}$. Lemma ~\ref{isolated puncture} and Lemma~\ref{limited near cycles} imply that such points have no common near cycles. Item (2) therefore follows from Remark~\ref{near cycle is open}. Item (3) follows from item (2) and the fact that every compact set has a finite cover by free disks.
We now assume that ${\cal R} \ne \emptyset$. Write $M$ as an increasing sequence of compact connected subsurfaces $M_1 \subset M_2 \subset \ldots$ such that each complementary component of each $M_i$ is unbounded and such that
$$N_{D_1}{\cal R} \subset M_i \subset M_i \cup f(M_i) \subset \Int M_{i+1}$$
for each $i$.
Lemma~\ref{U covers} implies that $f^n(y)$ converges to some puncture $c$ of $M$. After replacing $y$ by some point in its forward orbit, we may assume that $f^i(y) \in H \setminus M_2$ for all $i \ge 0$.
Given a lift $\tilde y$, let $\tilde C $ be the domain that contains $\tilde y$ and let $\tilde W_1$ be the component of $H \setminus \tilde M_1$ that contains $\tilde y$. We claim that $\tilde f_{\tilde C}(\tilde y) \in \tilde W_1$. Choose a ray $\tilde \rho \subset H \setminus \tilde M_2$ that connects $\tilde y$ to some $Q \in S_{\infty}$ and note that $Q$ is in the closure of $\tilde C$ and hence fixed by $\tilde f_{\tilde C}$ because $\tilde \rho$ does not cross any element of $\tilde {\cal R}$. Then $\tilde f_{\tilde C}(\tilde \rho) \subset H \setminus \tilde M_1$ connects $\tilde f_{\tilde C}(\tilde y)$ to $Q$. The claim now follows from the fact that $\tilde W_1$ is the unique component of $ H \setminus \tilde M_1$ whose closure contains $Q$.
Since $f^i(y) \in H \setminus M_2$ for all $i \ge 0$ the previous
argument can be iterated to show that $\tilde f^i(\tilde y) \in \tilde W_1$
for all $i \ge 0$. There exists $n_2$ so that $f^n(y) \in M\setminus
M_3$ for all $n \ge n_2$. Let $\tilde W_2$ be the component of $H
\setminus \tilde M_2$ that contains $\tilde f_{\tilde C}^{n_2}(\tilde y)$. By
the same argument, $\tilde f_{\tilde C}^n(\tilde y) \in \tilde W_2$ for all $n
\ge n_2$. Continuing in this manner, we can choose a decreasing
sequence of components $\tilde W_i$ of $H \setminus \tilde M_i$ such
that for all $k$, $\tilde f_{\tilde C}^n(\tilde y) \in \tilde W_k$ for all
sufficiently large $n$. It follows that $\tilde f_{\tilde C}^n(\tilde y)\to
P$ where $P$ necessarily projects to some end $c$. This proves that
$\tilde f_{\tilde C}$ is the unique $\omega$-lift for $\tilde y$.
Corollary~\ref{symmetric close} implies that
$\tilde C$ is a home domain for every point in $\tilde B(f) \cap \tilde W_1$.
By the symmetric argument applied to $ f^{-1}$, there is a unique domain $\tilde C^\ast$ that is an $\alpha$-domain for $\tilde y$; moreover
there is a neighborhood of $\tilde y$ such that $\tilde C^\ast$ is a home domain for every
birecurrent point in this neighborhood. To complete the proof of (1) it suffices to prove that
$\tilde C = \tilde C^\ast$.
If $\tilde C \ne \tilde C^\ast$, then both $\tilde C$ and $\tilde C^*$ are home domains for every birecurrent point in in a neighborhood of $\tilde y$. But then $y \in U(\sigma)$ where $\sigma = \tilde C \cap \tilde C^\ast$ contradicting the assumption that $y$ is not contained in any $U \in {\cal U}$.
Every neighborhood of $\tilde y$ contains points in $\tilde B(f)$ that are contained in distinct elements of ${\cal U}$. Lemma ~\ref{isolated puncture} and Lemma~\ref{limited near cycles} imply that such points have no common $\tilde f_{\tilde C}$-near cycle in $\Stab(\tilde C)$. Item (2) now follows from Remark~\ref{near cycle is open}.
Any compact $X \subset M$ has a cover by finitely many, say $K$, free
disks. Since $N_{D_1}(\bar C_{\core})$ is a
compact subset of $\bar C$, there is a constant $L$ so that for each of these $K$ free disks $B$, there are at most $L$
$\Stab(\tilde C)$-orbits of lifts of $B$ to $H$ that intersect $N_{D_1}(\tilde C)$. From (1) and Lemma~\ref{bounded distance} it follows that $\tilde f_{\tilde C}^i(\tilde y) \in N_{D_1}(\tilde C)$ for all $i$.
Item (2) therefore implies that there are at most $K_X= KL$ values of $i$
such that $f^i(y) \in X$. This proves (3).
Applying (3) to $X= M_2$ we have that for each $y \in Y$ there exists $0 \le i \le K_{M_2}$ such $f^i(y_1) \in M \setminus M_2 $. Choose $\epsilon > 0$ so that
\begin{itemize}
\item $z \notin M_2 \implies N_{\epsilon}(z) \cap M_1 = \emptyset$.
\item $z \in M_2$, $1 \le i \le K_{M_2}$ and $f^i(z) \notin M_2 \implies f^i(N_{\epsilon}(z)) \cap M_1 = \emptyset$.
\end{itemize}
Suppose now that $ \tilde C'$ is the home domain for $\tilde y_1$ and that $\dist(\tilde y_1, \tilde y_2) < \epsilon$. There exists $0 \le i \le K_{M_2}$ such that $\tilde f_{\tilde C'} ^i(\tilde y_1)$ and $\tilde f_{\tilde C'}^i(\tilde y_2)$ belong to the same component of $H \setminus \tilde M_1$. The domain that contains $\tilde f_{\tilde C'}^i(\tilde y_1)$ and $\tilde f_{\tilde C'}^i(\tilde y_2)$ is the home domain for these points so it is $ \tilde C'$. This implies that $\tilde C'$ is also the home domain for $\tilde y_2$.
\endproof
\begin{cor} \label{frontier home domains} Suppose that $\tilde V$ is a component of $\tilde U \in \tilde {\cal A}$ and that the union $\tilde V'$ of $\tilde V$ with all of its bounded complementary components has finite area. Then each point in the frontier $\fr(\tilde V)$ of $\tilde V$ has the same home domain.
\end{cor}
\proof Choose $\epsilon > 0$ as in Corollary~\ref{more home
domains}-(4). Since $\tilde V'$ is simply connected it is the union of
an increasing sequence of compact disks $\{B_i, i=1\ldots \infty\}$.
Since $\tilde V'$ has finite area we may assume that each $\partial B_i
\subset N_{\epsilon}(\fr (\tilde V'))$. For any pair of points $\tilde y_1,
\tilde y_2 \in \fr(\tilde V')$ there exists $k$ such that $\partial B_k$
intersects both $N_{\epsilon}(\tilde y_1) $ and $N_{\epsilon}(\tilde y_2) $.
It follows that $N_{\epsilon}(\fr (\tilde V'))$ is connected.
Lemma~\ref{more home domains}-(4) implies that each point in $\fr(\tilde
V')$ has the same unique home domain $\tilde C$ and hence, by
Lemma~\ref{bounded distance}, that $\fr(\tilde f_{\tilde C}^i(\tilde V') )
= \tilde f_{\tilde C}^i(\fr(\tilde V')) \subset N_{D_1}(\tilde C)$ for all
$i$. It follows that $ \tilde f_{\tilde C}^i(\tilde V') \subset N_{D_1}(\tilde
C)$ for all $i$ which implies by Lemma~\ref{omega condition} that
$\tilde C$ is a home domain for every point in $\tilde V'$.
\endproof
Our next arguments make use of a projection to $\mathbb R$ defined with respect to the defining parameter of $\tilde U \in \tilde {\cal U}$. If $\tilde U = \tilde U(\tilde \gamma)$ choose a parameterization of the annulus $A_\gamma$ (see Definition~\ref{annulus cover}) as $S^1 \times [0,1]$. Lift this to a parameterization of $H \setminus \tilde \gamma^{\pm}$ as $\mathbb R \times [0,1]$ and let $\pi: H \setminus \tilde \gamma^{\pm} \to \mathbb R $ be projection onto the $\mathbb R$ coordinate. (Alternately, one can define this directly as orthogonal projection onto $\tilde \gamma$ parameterized as $\mathbb R$.) If $\tilde U = \tilde U(P)$ where $P$ projects to an isolated end $M$ with horocycle $\tau$ define $\pi : H \setminus P \to \mathbb R $ as above using the compactified annular cover $A_P^c = A_\tau^c$. In both case we say that the $\pi$ is the {\em projection associated to $\tilde U$.}
\begin{remark} If $\tilde \sigma \in \tilde {\cal R}$ then by Lemma~\ref{more home domains}-(1), the home domain for a point in the frontier of $\tilde U(\tilde\sigma)$ is one of the two home domains of $\tilde U(\tilde\sigma)$.
\end{remark}
\begin{cor} \label{cor: frontier-bound} Suppose that $T$ is the covering translation associated to $\tilde U \in \tilde {\cal A}$, that $\pi$ is projection associated to $\tilde U$ and that $\tilde C$ is a home domain for a component $\tilde V$ of $\tilde U$. Given $p,q>0$ define $\tilde g = T^{-p} \tilde f_{\tilde C}^q$. Then there exists $r > 0$ so that
$\pi(\tilde g^r(\tilde y)) < \pi (\tilde y) -1$
for all $\tilde y$ in the frontier of $\tilde V$ whose home domain is $\tilde C$.
\end{cor}
\proof After replacing $f$ with $f^q$ we may assume that $q=1$. Increasing $p$ makes the desired inequality easier to satisfy so we may assume that $p=1$ and $\tilde g = T^{-1}\tilde f_{\tilde C}$. The goal is to prove that
\begin{equation}\label{eqn: frontier-bound}
\pi(\tilde f^r(\tilde y)) < \pi (\tilde y) +r-1
\end{equation}
for all sufficiently large $r$.
There is a constant $B$ so that $\pi(\tilde f(\tilde y)) < \pi(\tilde y) + B$ for all $\tilde y$.
Choose compact subsurfaces $M_1 \subset M_2 \subset M$ such that
\begin{enumerate}
\item Each component $X$ of $M\setminus M_2$ and each component $Z$ of $M\setminus M_1$ is unbounded.
\item For each component $X$ of $M \setminus M_2$ there is a component $Z$ of $M \setminus M_1$ such that $X \cup f(X) \subset Z$.
\end{enumerate}
We first consider the case that $\tilde U = \tilde U(\tilde \gamma)$. We may assume that
\begin{enumeratecontinue}
\item $M \setminus M_1$ does not contain a simple closed curve that is isotopic to $\gamma$
\end{enumeratecontinue}
and hence that for any component $\tilde Z$ of $H \setminus \tilde M_1$, the diameter of $\pi(\tilde Z)$ is at most the length $L$ of a fundamental domain of $\tilde \gamma$.
Suppose that $ f^j(y) \not \in M_2$ for $0 \le j \le J$. Let $X_0$ be the component of $M \setminus M_2$ that contains $y$ and let $Z_0$ be the component of $M\setminus M_1$ satisfying $X_0 \cup f(X_0) \subset Z_0$. Lift these to $\tilde y \in \tilde X_0 \subset \tilde Z_0$. Then $\tilde f_{\tilde C} ( \tilde y) \in \tilde Z_0$. There is a component $X_ 1 \subset Z_0 $ of $M\setminus M_2$ that contains $f(y)$. Lift this to $ \tilde f_{\tilde C} ( \tilde y) \in \tilde X_1 \subset \tilde Z_0$. Repeating this argument shows that $\tilde f_{\tilde C}^j(y) \in \tilde Z_0$ for $0 \le j \le J$ and hence that $\pi(\tilde f^J(\tilde y)) < \pi (\tilde y) +L$.
Since $y$ is in the frontier of $U$ it is not in any element
of ${\cal A}.$ Hence by Lemma~\ref{more home domains}-(3) there is a constant $K$ such that there are at most $K$ values of $j$ with $ f^j(y) \in M_2$.
The inequality~(\ref{eqn: frontier-bound}) therefore holds for $r > KA + (K+1)L+1$.
The second and last case is that $\tilde U =\tilde U(P)$. One component $
Z_1$ of $M\setminus M_1$ is a neighborhood of the isolated end of $M$
that lifts to $P$. Let $A$ [resp. $A^c$] be the [resp. compactified]
annular cover corresponding to $P$ and let $\partial_1 A^c$ be the
component of $\partial A^c$ that is not part of $A$. As a first
subcase suppose that either $\Fix( \hat f |_{\partial_1A^c}) =
\emptyset$ or
$\hat f$ is isotopic to a non-trivial Dehn twist
relative to fixed points in both components of $\partial A^c$. By
Corollary~\ref{PB} we may replace (3) by
\begin{itemize}
\item [($3'$)] $\tilde Z_1 \cap \fr(\tilde V) = \emptyset.$
\end{itemize}
Since the other components of $M\setminus M_1$ do not contain simple closed curves that are isotopic to horocycles corresponding to $P$, the proof given in the first case carries over to this subcase.
It remains to consider the subcase that $\hat f$ is isotopic to the identity relative to fixed points in both components of $\partial A^c$. In this subcase we replace (3) by
\begin{itemize}
\item [($3'$)] if $J \le 10$ and $\tilde f_{\tilde C}^j(y) \in \tilde Z_1$ for all $0 \le j \le J$ then $\pi(\tilde f^{J}(\tilde y)) < \pi (\tilde y) +1$
\end{itemize} and choose $r > KA+(K+1)L + 1+r/10$.
\endproof
\begin{lemma} \label{U is an annulus}Suppose that $U \in {\cal A}$.
\begin{enumerate}
\item $U $ is an open annulus.
\item If $U =U(\gamma)$ then each essential simple closed curve in $U $ is isotopic to $\gamma$. If $U =U(P)$ then each essential simple closed curve in $U $ is isotopic to a horocycle surrounding the isolated end of $M$ corrseponding to $P$.
\item Assume that the ends of $U$ are non-singular. If $U =U(\gamma)$ then both components of $\partial U_c$ of $U$ have fixed points. If $U =U(P)$ then the frontier of $U$ in $M$ is connected and there is a fixed point in the corresponding component of $\partial U_c$.
\end{enumerate}
\end{lemma}
\proof Choose $\tilde U \in \tilde {\cal A}$ projecting to $U$ and let $T $ be the covering translation associated to $\tilde U$. By Lemma~\ref{U tilde properties}-(3), (1) and (2) follow from the following three properties that we prove below.
\begin{enumerate}
\item[(a)] each component $\tilde V$ of $\tilde U$ is simply connected.
\item [(b)] each component $\tilde V$ of $\tilde U$ is $T$-invariant.
\item [(c)]$\tilde U$ is connected.
\end{enumerate}
We first prove that each component $\tilde V$ of $\tilde U$ is unbounded by
assuming that $\tilde V$ is bounded and arguing to a contradiction. Let
$\tilde f = \tilde f_{\tilde C}$ where (Corollary~\ref{frontier home domains})
$\tilde C$ is a home domain for each point in the frontier of $\tilde V$.
Since $\tilde f$ preserves area there exists $q \ge 0$ and a covering
translation $S$ so that $\tilde f^q(\tilde V) \cap S(\tilde V) \ne \emptyset$.
Lemma~\ref{U tilde properties}-(3) implies that $S = T^p$ for some $p
\in \mathbb Z$. After replacing $T$ with $T^{-1}$ if necessary we may assume
that $p > 0$. From the fact that $\tilde f^q(\tilde V )$ and $S(\tilde V)$ are
both components of $\tilde U$, it follows that $\tilde f^q(\tilde V ) = S(\tilde
V) = T^p(\tilde V)$. Thus $\tilde V$ is $\tilde g$-invariant where $\tilde g =
T^{-p}\tilde f^q$. If $p=0$ then $\tilde f$ has bounded orbits (since
we are assuming $\tilde V$ is bounded) and hence
fixed points by the Brouwer plane translation theorem. Since $\tilde f$
is fixed point free, $p \ne 0$. This contradicts
Corollary~\ref{cor: frontier-bound} and so
completes the proof that each component $\tilde V$ of $\tilde U$ is unbounded.
If (a) fails then some component of the complement of $\tilde V$ is
bounded and intersects some $\tilde U' \ne \tilde U$. This contradicts the
above fact that no component of $\tilde U'$ is bounded, and hence proves
(a).
We next assume that (b) fails and argue to a contradiction. A subsurface that contains a curve homotopic to an iterate of $\gamma$ contains a curve isotopic to $\gamma$. Thus $T^p(\tilde V) \ne \tilde V$ for all $p \ne 0$. Lemma~\ref{U tilde properties}-(3) implies that $\tilde V$ is moved off itself by every covering translation. In particular, $\tilde V$ has finite area because the covering projection into $M$ is injective on $\tilde V$. Define $\tilde f = \tilde f_{\tilde C}$ where $\tilde C$ is a home domain for each point in the frontier of $\tilde V$. As in the previous argument, there exists $p$ and $q\ge 0$ so that $\tilde f^q(\tilde V) = T^p(\tilde V)$. If $p = 0$, then $\tilde V $ has recurrent points, and hence fixed points for $\tilde f$, which is impossible. Thus $p \ne 0$ and we assume without loss that $p > 0$. Also by hypothesis, $q \ne 0$.
Let $\pi $ be the projection associated to $\tilde U$ and let
$\tilde g = T^{-p}\tilde f^q $. Then
$\tilde g( \tilde V) = \tilde V$ and by Corollary~\ref{cor: frontier-bound}, there is an $r>0$ such that $\pi(\tilde g^r(\tilde y)) < \pi(\tilde y) -1$ for every $\tilde y$ in $\partial \tilde V.$
Consequently, by continuity, there is $\delta>0$ such that every
$\tilde x\in \tilde V$ which is within $\delta$ of $\partial \tilde V$ satisfies
$\pi(\tilde g^r(\tilde x)) < \pi(\tilde x) -1.$
Since $\tilde V$ has finite area there exists $N>0$ such that
$\tilde V_{-N} = \{ \tilde x \in \tilde V\ |\ \pi(\tilde x) < -N\}$ contains
no ball of diameter $\delta$, and hence every point of
$\tilde V_{-N}$ must be within $\delta$ of $\partial \tilde V.$
We conclude the $\tilde g^r(\tilde V_{-N})$ is a subset
of $\tilde V_{-N-1}$. Note that $\tilde V_{-N}$ is non-empty since $\tilde V$ is $\tilde g$-invariant and $\lim_{n\to \infty} \pi \tilde g^{nr}(\tilde y) = -\infty$ for all $\tilde y \in \partial \tilde V$.
Since $\tilde V_{-N}$ is open and $\tilde V_{-N-1}$ is a proper
subset the area of $\tilde V_{-N-1}$
is strictly smaller than that of $\tilde V_{-N}$ contradicting
the fact that $\tilde g$ is area preserving. This completes the proof of (b).
We have now proved that each component of $U$ contains a simple closed curve that is essential in $M$ and that all such simple closed curves in $U$ are in the same isotopy class. Moreover if $U \ne U'$ then $U$ and $U'$ do not contain isotopic simple closed curves. If (c) fails then there is an unpunctured annulus $A$ whose boundary curves are in $U$ and whose interior contains a component $V'$ of some $U' \ne U \in {\cal U}$. But every simple closed curve in $V'$ that is essential in $M$ is parallel to the boundary components of $A$. This contradiction implies that $U$ and hence $\tilde U$ is connected. This completes the proof of (1) and (2). This same argument proves that if $U$ corresponds to an isolated end of $M$ then $U$ contains a neighborhood of that isolated end and so has connected complement and connected frontier.
We now consider (3). Choose a component $Y$ of the frontier of $U$ and let $\partial_0 U_c$ be the corresponding component of $\partial U_c$. We will assume that $\rho(\partial_0 U_c) >0$ and argue to a contradiction. A symmetric argument implies that $\rho(\partial_0 U_c)$ is not negative and so is $0$, which is equivalent to $\partial_0 U_c$ having a fixed point.
Choose an accessible point $z \in Y $ and a degree one closed path $\mu$ with embedded interior in $U$ and with both endpoints at $z$. Let $\tilde \mu_0$ be a lift of the interior of $\mu$ to $\tilde U$. Since $\mu$ has degree one, the ends of $\tilde \mu_0$ converge to lifts $\tilde z$ and $T (\tilde z)$ of $z$ in $\tilde Y$. Denote the bounded area component of $\tilde U \setminus \tilde \mu_0$ by $D_0$. For each $k$, let $\tilde \mu_k = T^k (\tilde \mu_0)$ and $D_k = T^k(D_0)$.
Choose $0 < p/q < \rho(\partial_0 U)$ and let $\tilde g = T^{-p}\tilde f^q $. If $\nu$ is a cross cut in $D_k$ then the $\pi$-image of the endpoints of $\tilde g^j(\nu)$ decrease linearly in $j$ by Corollary~\ref{cor: frontier-bound}. On the other hand, for arbitrarlily large $j$ one can choose $\nu$ so that $\tilde g^j(\nu)$ is a cross cut in $D_l$ where $l$ increases linearly in $j$. It follows that $\pi(D_k) $ is not bounded below.
Choose $N$ so that $\pi(\tilde \mu_0) > -N$. For all $n > N$ and $k > 0$ consider all cross cuts $\tau_{k,n}\subset D_k$ such that $\pi(\tau_{k,n}) =-n$. (In other words, $\tau_{k,n}$ is a component of $\pi^{-1}(n) \cap D_k$.) Let $E(\tau_{k,n})$ be the complementary component of $\tau_{k,n}$ that is contained in $D_k$ and let $ d_{k,n}$ be the maximum area of all such $E(\tau_{k,n})$. Then
$$
d_{k,n} = d_{k+1,n-1} < d_{k+1,n}
$$
The equality follows from the fact that $\tau_{k+1,n-1} = T(\tau_{k,n}) \subset D_{k+1}$ is a cross cut with $\pi(\tau_{k+1,n-1}) = -n+1$. The inequality follows from the fact that each $E(\tau_{k,n})$ is contained in some $E(\tau_{k,n-1})$.
Fix $k$ and choose $\tau_{k,n}$ so that $d_{k,n} = E(\tau_{k,n})$. Since $D_k$ has finite area, we have $\lim _{n \to \infty} d_{k,n} = 0$. Arguing as in the proof of (2), there exits $r > 0$ and $N' > N$ such that $\pi(\tilde g^r(\tau_{k,n})) < -n-1$ for all $n > N'$. Our choice of $N$ guarantees that $\tilde g^r(\tau_{k,n}) \cap \tilde \mu_l = \emptyset$ for $l \ge k$. Since $p/q < \rho(\partial_0 U)$ it follows that $\tilde g^r(E(\tau_{k,n}))$ is contained in $D_l$ for some $l \ge k$ and hence that $\tilde g^r(E(\tau_{k,n}))$ is contained in some $E(\tau_{l,n+1})$. This contradicts the fact that $d_{l,n+1} < d_{k,n}$ for all $l \ge k$.
\endproof
We are now able to prove Proposition~\ref{intermediate}.
\vspace{.1in}
\noindent {\bf Proposition~\ref{intermediate}} {\em Suppose that $M$ is a component of ${\cal M} = S^2 \setminus \Fix(F)$ and that $f=F|_M:M \to M$. Then there is a countable collection ${\cal A}$ of pairwise disjoint
open $f$-invariant annuli such that }
\begin{enumerate}
\item {\em For each compact set $X \subset M$ there is a constant $K_X$
such that any $f$-orbit that is not contained in some $U \in {\cal A}$
intersects $X$ in at most $K_X$ points. In particular each $x \in
{\cal B}(f)$ is contained in some $U \in {\cal A}$.}
\item {\em For each $U \in {\cal A}$ and $z$ in the frontier of $U$ in $S^2$, there
are components $F_+(z)$ and $F_-(z)$ of $\Fix(F)$ so that $\omega(F,z)
\subset F_+(z) $ and $\alpha(F,z) \subset F_-(z)$.}
\item {\em For each $U \in {\cal A}$ and each component $C_M$ of the frontier of
$U$ in $M$, $F_+(z)$ and $F_-(z)$ are independent of the choice of $z
\in C_M$. }
\item{\em If $U \in {\cal A}$, and $f_c:U_c \to U_c$ is the extension
to the annular compactification of $U$, then each component of $\partial U_c$
corresponding to a non-singular end of $U$ contains a fixed point of $f_c.$}
\end{enumerate}
\noindent{\bf Proof of Proposition~\ref{intermediate}}
The case in which $M$ has less than three ends is proved in section~\ref{sec: intermediate} so we may assume that $M$ has at least three ends.
The index set ${\cal A}$ is defined in Definition~\ref{defn: A}. Lemma~(\ref{U is an annulus})-(1) states that each $U \in {\cal A}$ is an open annulus. Lemma~\ref{more home domains}-(3) implies (1) which implies (2). Item (4) follows from Lemma~\ref{U is an annulus}-(3).
To prove (3), suppose that $\tilde Z$ is a component of the frontier
of $\tilde U$ in $H.$ If three oriented simple closed curves in $S^2$ bound
disjoint disks then at least one pair of them have anti-parallel
orientations. Part (2) of Lemma~(8.7) of \cite{fh:periodic}
therefore implies that if a free disk in $H$ intersects
distinct $\tilde U_1, \tilde U_2$ and $\tilde U_3$ then one of these annuli
separates the other two. It follows that each $\tilde z \in \tilde Z$ has a
neighborhood that intersects exactly one component $V_{\tilde z}$ of the
complement of the closure of $\tilde U$. The open set $V_{\tilde z} =
V_{\tilde Z}$ depends only on $\tilde Z$ and not on $\tilde z$. In particular,
$\tilde Z$ is contained in the frontier of $V_{\tilde Z}$.
Let $W$ be the component of the complement of $\tilde U$ that contains
$V_{\tilde Z}$ and so contains $\tilde Z$. Lemma~(\ref{planar topology})
implies that the frontier of $W$ is connected and hence is contained
in a component of the frontier of $\tilde U$. Thus $\tilde Z$ is the
frontier of $W$. Lemma~(\ref{planar topology}) implies that the
complement of $\tilde Z$ in $H$ has exactly two components. It follows that the intersection $B$ of $S_{\infty}$ with the closure of $\tilde Z$ cannot have more than two components: if it did there would be two components of $S_{\infty} \setminus B$ with neighborhoods contained in the same component of $H \setminus \tilde Z$ and so there would be a line in $H \setminus \tilde Z$ that separates $\tilde Z$ which is impossible.
The components of $B$ must be points because translates of
$\tilde U = \tilde U(\tilde \gamma)$ by covering translations will have ends
of their parametrizing geodesics which are dense in $S_\infty$.
By Lemma~\ref{frontier home domains} and Lemma~\ref{more home domains}(1) there exists $\tilde C$ that is a home domain for each point in a neighborhood of $\tilde Z$. Since $W$ contains elements of ${\cal A}$ that intersect this neighborhood, $W$ is $\tilde f_{\tilde C}$-invariant. It follows that $\tilde Z$ is $\tilde f_{\tilde C}$-invariant.
If $B$ is a single point $P$ then $P$ is also the intersection of $S_{\infty}$ with the closure of one of the complementary components of $\tilde Z$. The only possibility is that this complementary component is $\tilde U(P)$. But this contradicts (2) applied to the projection $Z$ in $M$ of $\tilde Z$ since $Z$ is disjoint from $U(c)$ which is a neighborhood of the end corresponding to $P$. We conclude that the limit set of $\tilde Z$ in $S_{\infty}$ is a pair of points, say $a$ and $b$.
Choose a small disk $\tilde D$ that projects to a free disk $D \subset M$
and contains an arc $\beta_0 $ that has endpoints in distinct
components of $H \setminus \tilde Z$. Extend $ \beta_0$ to a
properly embedded line $ \beta \subset H$ that intersects $\tilde Z$
only in $ \beta_0 \cap \tilde Z$ and separates $a$ from $b$. Let $\tilde
Z_a$ be the set of points in $\tilde Z $ that are in the same component
of the complement of $\beta$ as $a$; define $\tilde Z_b$ similarly.
Perform an isotopy rel $\tilde Z \cup S_{\infty}$ from $\tilde f_{\tilde C}$ to a
homeomorphism $g$ such that $g(\beta) \cap \beta = \emptyset$. After
interchanging $a$ and $b$ we may assume that $g(\beta)$ separates $b$
from $\beta$. Since $g|_{\tilde Z} = \tilde f_{\tilde C}|_{\tilde Z}$ it follows
that $\tilde f_{\tilde C}(\tilde Z_b)\subset Z_b$ which implies that $b=
\omega(\tilde z, \tilde f_{\tilde C})$ for each $\tilde z \in Z_b$. Varying the
location of $\tilde D$ we conclude that $b= \omega(\tilde z, \tilde f_{\tilde C})$
for all $\tilde z \in \tilde Z$. Applying the same argument to $\tilde
f^{-1}_{\tilde C}$ we see that $a= \alpha(\tilde z, \tilde f_{\tilde C})$ for all
$\tilde z \in \tilde Z$ proving (3).
\section{Renormalization.}\label{sec: renormalization}
In this section we study the finer structure of $f|_U$, the restriction
of $f$ to one of the annuli $U \in {\cal A}$.
For each $q \ge 1$ let ${\cal M}_q = S^2 \setminus \Fix(F^q) \subset S^2
\setminus \Fix(F) = {\cal M}$. Recall that by the main theorem of \cite{brnkist}, each
component $M$ of ${\cal M}$ is $F$-invariant and
similarly each component $M_q$ of ${\cal M}_q$ is
$F^q$-invariant. Let ${\cal A}(q)$ be the family of open
$F^q$-invariant annuli obtained by applying
Proposition~\ref{intermediate} to the restriction of $F^q$ to a
component $M_q$ of ${\cal M}_q$ that is contained in the component $M$ of
${\cal M}$.
\begin{lemma}\label{lem: ess => inv}
\begin{enumerate}
\item If $V \in {\cal A}(q)$ is essential in $M$ then $V$ is $f$-invariant.
\item If $V$ is any $f$-invariant open annulus in $M$ then
$V$ is essential in $M.$
\end{enumerate}
\end{lemma}
\begin{proof}
Suppose $V \in {\cal A}(q)$ is essential in $M$. Since $f$ commutes with
$f^q$, $f(V)$ is an element of ${\cal A}(q)$ (see
Remark~(\ref{second centralizer invariance})) and hence
$f(V)$ is either equal to
or disjoint from $V$. Since $V$ is essential in $M$, $f(V)$ is
essential in $M$. If $f(V)$ is disjoint from $V$ then since $f(V)$ is
essential in $M$, $f$ maps one component of the complement of $V$ to a
proper subset of itself contradicting the fact that $f$ preserves
area. This proves (1).
If an
$f$-invariant open annulus $V \subset M$ is inessential then one of
its complementary components is contained in $M$. The union of $V$
and this component is an $f$-invariant open disk which, by the Brouwer
plane translation theorem, must contain a point of $\Fix(f) \cap M$.
This contradiction implies that $V$ is essential in $M$ and so verifies (2).
\end{proof}
The following proposition shows that elements
of the family ${\cal A}(q)$ refine the elements of ${\cal A}$.
\begin{prop}\label{prop: V in U}
Each $V \in {\cal A}(q)$ is a subset of some $U \in A.$
Moreover, each $f$-invariant open annulus
in $M$ is contained in some $U \in {\cal A}$.
\end{prop}
\begin{proof}
Assume at first that $V$ is an essential annulus in $M$. If $V \in {\cal A}(q),$
then $V$ is $f$-invariant by Lemma~(\ref{lem: ess => inv}.)
Let $\alpha$ be a simple
closed loop in $V$ representing a generator of its fundamental
group and let $\gamma$ be the simple closed geodesic or a horocycle in
$M$ that is freely homotopic to $\alpha$. Since $V$ is $f$-invariant, $\gamma$ is isotopic to $f(\gamma)$ and in particular does not
cross any reducing curves.
Choose a lift $\tilde \gamma \subset H$ of $\gamma$ and let $T= T_{\tilde \gamma}$ be a primitive covering translation that preserves ${\tilde \gamma}$.
If $\gamma$ is not itself a reducing curve then $\tilde \gamma$ lies in
a unique domain $\tilde C$.
The lift $\tilde f_1 = \tilde f_{\tilde C}$ of $f$ commutes with $T$ and fixes the
endpoints $ \tilde \gamma^{\pm}$.
If $\gamma$ is in $R$ then $\tilde \gamma$ is the common frontier of
two domains $\tilde C_1$ and $\tilde C_2$. Let $\tilde f_j,\ j = 1,2$
be the lift which fixes the ends of $\tilde C_j$. In this case too $\tilde f_j$ commutes with $T$ and fixes $ \tilde \gamma^{\pm}$.
The homotopy from $\gamma$ to $\alpha$ lifts to a homotopy between $\tilde \gamma$ and a lift $\tilde \alpha$ of $\alpha$. Since the lifted homotopy moves points a uniformly bounded distance and commutes with $T$, the ends of $\tilde \alpha$ converge to $\tilde \gamma^{\pm} $ and $\tilde \alpha$ is $T$-invariant. The components of the full pre-image of $V$ are copies of the universal cover of $V$; we refer to each component as a lift of $V$. The lift $\tilde V$ of $V$ that contains $\tilde \alpha$ is $T$-invariant and is the only lift of $V$ whose closure contains $\tilde \gamma^{\pm}$. Since $\tilde f_j$ fixes $\tilde \gamma^{\pm}$ it follows that $\tilde V$ is $\tilde f_j$-invariant.
Suppose $x \in V$ is birecurrent. Let $\tilde x$ be a lift of $x$
which lies in $\tilde V.$ Choose a small free disk $D \subset V$ with
diameter $\epsilon$ containing $x.$
Suppose that $f^m(x) \in D$. Then if $\tilde D$ is the lift of $D$
containing $\tilde x,$ we have $\tilde f_j^m(\tilde x) \in T^k(\tilde D)$ for some
$k \in \mathbb Z.$
Since $x$ is recurrent there is a sequence $\{n_i\}$ tending to
infinity such that $ f_j^{n_i}(x) \in D $ and hence a sequence $\{k_i\}$
such that $\tilde f_j^{n_i}(\tilde x) \in T^{k_i}(\tilde D).$ It follows that
$\tilde f_j^{n_i}(\tilde x)$ has distance at most $\epsilon$ from $T^{k_i}(\tilde x)$.
We conclude that
\[
\lim_{i \to \infty} \tilde f_j^{n_i}(\tilde x) = \lim_{i \to \infty} T^{k_i}(\tilde x)
= \tilde \gamma^+ \text{\ or\ }\tilde \gamma^-.
\]
This shows that $\omega(\tilde f_j, \tilde x) = \tilde \gamma^+$ or $\tilde \gamma^-.$
A similar argument shows $\alpha(\tilde f_j, \tilde x) = \tilde \gamma^+$
or $\tilde \gamma^-.$
Hence $\tilde x \in U(\tilde \gamma).$
Since $x \in V$ was an arbitrary recurrent point we conclude that $V
\subset U$ where $U \in {\cal A}$ is the projection of $\tilde U(\tilde \gamma)$
in $M$. This completes the proof when $V$ is essential. Note
that by Lemma~(\ref{lem: ess => inv}) any $f$-invariant $V$ in $M$
must be essential so we have proved the ``moreover'' part of
the result.
We are left with the case that $V \in {\cal A}(q)$ is inessential in $M$.
An essential closed curve in $V$ bounds a closed disk in $M$ and we
let $W$ be the union of $V$ and this disk. Since $W$ is open and
invariant under $f^q$ there is a periodic point $p \in W \cap
\Fix(f^q)$ by the Brouwer plane translation theorem. Choose a lift
$\tilde p$ of $p$ and let $\tilde C$ be a home domain for $\tilde p$, let $\tilde
W$ be the lift of $W$ that contains $\tilde p$ and let $\tilde U$ be the
element of $\tilde {\cal A}$ that contains $\tilde p$. It suffices to prove that
$\tilde W \subset \tilde U$ and for this it suffices to show that if $\tilde w
\in \tilde W$ is a lift of $w \in W\cap {\cal B}(f)$ then $\omega(\tilde f_{\tilde
C}, \tilde w) = \omega(\tilde f_{\tilde C}, \tilde p)$.
Choose a path $\rho \subset W$ connecting $w$ to $p$. If $K$ is
a number greater than the length of $\rho$ then
there is a sequence $n_i \to \infty$
such that $f^{n_i}(w)$ can be joined to
$p$ by a path in $W$ of length at most $K$. Thus
\[
\omega(\tilde f_{\tilde C}, \tilde w) = \lim \tilde f^{n_i}_{\tilde C}( \tilde w ) = \lim \tilde f^{n_i}_{\tilde C}( \tilde p ) = \omega(\tilde f_{\tilde C}, \tilde p)
\]
as desired.
\end{proof}
\begin{remark} $V \in {\cal A}(q)$ is essential in ${\cal M}$ if and only if it is
essential in the unique $U \in {\cal A}$ containing it. In this case we will simply say that $V$ is
essential.
\end{remark}
Recall (see Definition~(\ref{defn: annular comp})) that for any open $f$-invariant annulus $V \subset M$
there is a natural annular compactification of $V$
denoted $V_c$ and an extension of $f$ to the closed annulus
$f_c : V_c \to V_c.$ To simplify notation we will denote
the rotation interval $\rho(f_c)$ by $\rho(V)$ when there
is no ambiguity about the choice of diffeomorphism $f$
but various annuli $V$ are under consideration.
The next lemmas provide information about
the translation and rotation intervals of the extension $f_c: U_c \to U_c$ of $f$ to the
annular compactification of $U$.
\begin{lemma} \label{lem: not in V} Suppose that $q$ is prime and that $x \in U$ is not contained in any $V \in {\cal A}(q)$.
\begin{enumerate}
\item The rotation number, $\rho_f(x),$
with respect to $f_c$ is well defined.
\item If $\omega(f_c, x) $ contains a point of
$U$ then $\rho_f(x) = p/q$ for some $0 < p < q$.
\item If $\omega(f_c, x)$ is contained in a component of $\partial U_c$ corresponding to a non-singular end of $U$ then $\rho_f(x) = 0$.
\item $\omega(f_c, x)$ is contained in a component $B$ of $\partial U_c$ corresponding to a singular end of $U$ then $\rho_f(x) = \rho(B)$.
\end{enumerate}
\end{lemma}
\begin{proof} Each point in $U \setminus \Fix(f^q)$ has a compact neighborhood in $U\setminus \Fix(f^q)$. Proposition~\ref{intermediate}-(1) therefore implies that $\omega(f^q,x) \subset \Fix(f^q)$ and hence that $\omega(f_c^q,x) \subset \Fix(f_c^q) \cup \partial U_c$. Since each component of $\Fix(f_c^q) \cup \partial U_c$ is $f_c^q$-invariant, $\omega(f_c^q,x)$ is contained in a component $K$ of $\Fix(f_c^q) \cup \partial U_c$.
The rotation number $\rho_{f_c}$ is well defined and constant on each
component of $\partial U_c$. It is also well defined and locally
constant on $\Fix(f_c^q)$. Since both sets are closed, $\rho_{f_c}$ is locally constant on their union and hence constant on $K$, say $\rho(f_c, K) =\rho_K$. It follows that $\rho(f_c^q, K) = q \rho_K.$
In fact, more is true. There is a lift $\tilde f_c : \tilde U_c \to \tilde U_c$ of $f$ such that $\tau_{\tilde f_c}(\tilde y) = \rho_K$ for each $\tilde y$ that projects into $K$. Let $p_1 : \tilde U_c \to \mathbb R$ be the projection used to define $\tau_{\tilde f_c}$. Then for any $k \in \mathbb Z$
$$
|p_1\tilde f^{kq}(\tilde y) - p_1\tilde y - kq\rho_K| < 1
$$
for any $\tilde y$ that projects into $K$. For any fixed $k$, this inequality holds for any point $\tilde z$ that projects into a neighborhood, say $W_k$, of $K$. Suppose that $z$ and the forward $f^q$-orbit of $z$ is contained in $W_k$. Then by applying the above inequality with $\tilde y$ equal, in order, to $\tilde z,\tilde f^q_c( \tilde z), \tilde f^{2q}_c( \tilde z), \ldots, \tilde f^{(j-1)q}_c( \tilde z)$
we see that
$$
|p_1\tilde f^{qjk}(\tilde z) - p_1\tilde z - qjk\rho_K| < j
$$
for all $j$.
This proves that
$$
q\rho_K -1/k \le \liminf_{n \to \infty} \frac{p_1(\tilde f^{nq}(\tilde z)) - p_1(\tilde z)}{n} \le
\limsup_{n \to \infty} \frac{p_1(\tilde f^{nq}(\tilde z)) - p_1(\tilde z)}{n}
\le q\rho_K + 1/k
$$
for all $z$ with $\omega(f_c^q,z) \subset W_k.$
Since $\omega(f_c^q,x) \subset W_k$ for all $k$ it follows that
$\rho_{f_c^q}(x) = q\rho_K$ and hence that
$\rho_{f_c}(x) = \rho_K.$
If $K$ contains a point $y$ in the interior of $U_c$ then $y \in \Fix(f_c^q)$ and $\rho(f_c, K) =p/q$ for some $0 \le p < q$. If $p = 0$ then there is a lift $\tilde f : \tilde U \to \tilde U$ and a lift $\tilde y \in \Fix(\tilde f^q)$ in contradiction to the Brouwer translation theorem applied to $\tilde f$ and the fact that $\Fix(\tilde f) = \emptyset$. Thus $0 < p < q$.
If $K$ is a component of $\partial U_c$ corresponding to a non-singular end then $\rho_K = 0$ by Proposition~\ref{intermediate}-(4).
\end{proof}
\begin{lemma}\label{lem: U rot int}
Suppose that $U \in {\cal A}$ and that $X$ is a component
of $\partial U_c$ corresponding to a
non-singular end. Then the
translation number $\tau(\tilde f_c|_{\tilde X})$ of any lift of $f_c$
restricted to the universal covering space $\tilde X$ is an integer
$p.$ Moreover $\tau(\tilde {f_c})$ is a non-trivial
interval containing $p$ as an endpoint and having length
at most $1$.
\end{lemma}
\begin{proof}
No integer can be in the interior of $\tau(\tilde f_c).$ To see this
we suppose to the contrary that an integer (which without loss we
assume is $0$) is in the interior of $\tau(\tilde f_c)$ and show
this leads to a contradiction. In this case by Theorem~(\ref{thm: rot num})
there would be periodic points in $U$ with both positive and negative
rotation numbers. Theorem~(2.1) of \cite{franks:poincare} then implies
that $f$ has a fixed point in the open annulus $U$, which is a contradiction.
By part (4) of Proposition~\ref{intermediate}, $f_c$ has a fixed point
in $X$. It follows that the translation number of the lift
$\tilde f_c|_{\tilde X} : \tilde X \to \tilde X$ is an integer,
say $p.$ Hence $p \in \tau(\tilde
{f_c}).$ There is a point in the interior of $U$ with a well defined
non-integer translation number. This is because almost all points of
$U$ have a well defined translation number by Theorem~(\ref{thm: rot num})
and if these were all equal to $p$ then Proposition~(\ref{prop: +meas})
would imply $U$ contains a fixed point -- a contradiction.
Since $p \in \tau (\tilde f_c)$ and no integer can be in its interior, it
follows that $\tau(\tilde {f_c})$ is non-trivial, $p$ is one endpoint and
it must be contained in either $[p, p+1]$ or $[p-1, p].$
\end{proof}
\begin{cor} \label{cor: V(q)}
Suppose $q>1$ and $V \in {\cal A}(q)$ is a subset of $U \in {\cal A}$ and
$U$ has a non-singular end. If $V$ is $f$-invariant and $f_c : V_c \to V_c$
is the extension to its annular compactification
then for any lift $\tilde f_c$ to the universal covering
there is $p \in \mathbb Z$ such that $\tau(\tilde f_c)$ is a non-trivial
subinterval of $[p/q, (p+1)/q]$ and contains at least one of its endpoints.
\end{cor}
\begin{proof}
The fact that $U$ has a non-singular end implies that $V$
does also. The result now
follows from Lemma~(\ref{lem: U rot int}) since
\[
\tau(\tilde f_c) = \frac{\tau(\tilde f_c^q)}{q}
\]
and
\end{proof}
We are now prepared to prove that $\rho_f(x)$ is well defined
for all $x \in U.$ In fact we will show in part (d) of
Proposition~(\ref{prop: C(x)-pre}) that $\rho_f$ is a continuous
function on $U$.
\begin{lemma}\label{lem: welldefined}
If $x \in U \in {\cal A}$ then the rotation number for $x$ with respect to
$f_c: U_c \to U_c$ is well defined.
\end{lemma}
\begin{proof} We first consider forward rotation numbers. By Lemma~\ref{lem: not in V} we may assume that for all prime $q$ there exists
$V \in {\cal A}(q)$ containing
$x$. Choose lifts $\tilde f_c : \tilde U_c \to \tilde U_c$ and $\tilde x \in \tilde U_c$ to the universal cover $\tilde U_c$ and let
\[
r = \liminf_{n \to \infty} \frac{p_1(\tilde f^n(\tilde x)) - p_1(\tilde x)}{n}
\text{ \quad and \quad }
s = \limsup_{n \to \infty} \frac{p_1(\tilde f^n(\tilde x)) - p_1(\tilde x)}{n}
\]
Lemma~\ref {lem: U rot int} and Proposition~\ref{prop: limsup} imply that $s-r \le 1/q$ for all $q$ and hence that $r =s$.
This proves that $x$ has a well defined forward rotation number. Similar arguments
shows that $x$ has a well defined backward rotation number and a well defined rotation number.
\end{proof}
\begin{lemma}\label{prop: cA_q}
If $V \in {\cal A}(q)$ is a subset of $U \in {\cal A},$ then
\begin{enumerate}
\item If $V$ is essential in $U$, then every
$x \in V$ has the same rotation number in $V_c$
as in $U_c$.
\item The annulus $V$ is inessential in $U$ if and only if all
recurrent points in $V$ have the same rotation number for $f_c : U_c
\to U_c.$ In this case there is $p \in \mathbb Z$ with $0< p < q,$ such that
the common rotation number is $p/q$.
\end{enumerate}
\end{lemma}
\begin{remark}
In fact if $V \in {\cal A}(q)$ is inessential in $U$ then
{\em all} points of $V$ have the same rotation number,
not just the recurrent points. This will follow
when we show as part of Proposition~(\ref{prop: C(x)-pre}) that
$\rho_f$ is continuous on $U,$ since the recurrent
points are dense in $V$.
\end{remark}
\begin{proof}
If $V$ is essential in $U$ then it is $f$-invariant by
Lemma~(\ref{lem: ess => inv}). The inclusion of $V$ in $U$
induces an isomorphism on the fundamental group and any
point $x \in V$ has the same rotation number in $V$ and $U$.
Since $\rho(V_c)$ is non-trivial by Lemma~(\ref{lem: U rot int}),
points of $V$ have a non-trivial
interval of rotation numbers in $U$.
If $V$ is inessential in $U$ there is a component $X$ of its
complement in $S^2$ contained in $U$. The set $W = V \cup X$ is an open disk
invariant under $f^q.$ Since $f$ is area preserving and $f^q(W) =
W,$ by the Brouwer plane translation theorem there is a point $x_0
\in W \cap \Fix(f^q).$ The rotation number of $x_0$ is $p/q$ for some
$0 < p < q$ by Lemma~\ref{lem: not in V}-(1). If $x \in V$ is recurrent there is a constant $K_x$ and a
sequence $\{n_i = k_iq\}$ such that the distance in $U_c$ from
$f^{n_i}(x)$ to $f^{n_i}(x_0)$ is less than $K_x$. Hence the distance
in the covering space $\tilde U_c$ from $\tilde f^{n_i}(\tilde x)$ to
$f^{n_i}(\tilde x_0)$ is less than $K_x$, where $\tilde f, \tilde x,$ and $\tilde
x_0$ are lifts of $f, x,$ and $x_0$ respectively. It follows that
$x$ and $x_0$ have the same rotation number in $U$, namely $p/q.$
\end{proof}
\begin{cor} \label{cor: ess V} If $x \in U$ is not asymptotic to
an end of $U$ then, with at most two exceptions, for each prime $q$
there is $V \in {\cal A}(q)$ which is essential in $U$ and contains $x.$
\end{cor}
\begin{proof} If $x \in V_1 \in {\cal A}(q_1)$ and $x \in V_2 \in {\cal A}(q_2)$ for distinct primes $q_1$ and $q_2$ then $V_1 \cap V_2$ is an open set and so contains a recurrent point $x'$. Since the rotation number of $x'$ with respect to $f_c :U_c \to U_c$ can not have both the form $p/q$ and $p'/q'$, Lemma~(\ref{prop: cA_q})-(2) implies that either $V$ or $V'$ is essential in $U$. We conclude that there is at most one $q$ for which $x$ is contained in an inessential element of ${\cal A}(q)$. The corollary then follows from Lemma~(\ref{lem: not in V}) which implies that there is at most one $q$ for which $x$ is not contained in any $V \in {\cal A}(q)$.
\end{proof}
\begin{lemma}\label{lem: cl(V) in U}
Suppose
$U \in {\cal A}.$
If $V \in {\cal A}(q)$ is essential in\ $U$
and $\rho(V)$ is disjoint from $\rho(\partial U_c)$, then
$\cl(V) \subset U.$ Moreover, if $x \in U$ and $\rho_f(x)
\notin \rho(\partial U_c)$, then for every sufficiently large prime $q$
there exists an essential $V \in {\cal A}(q)$ such that
$x \in V,\ cl(V) \subset U.$
\end{lemma}
\begin{proof}
Let $X$ be one component of the complement of $V$ in $U$ and define
$W = U \setminus X$. Then $W$ is an open subannulus of $U$.
One of the components of its frontier, which we denote $Y$,
coincides with a component
of the frontier of $U$ and the other is the common frontier
of $V$ and $X$. The rotation interval $\rho(W)$ contains $\rho(V)$ and
the number $r_0 \in \rho(\partial U_c)$ corresponding to the end of $U$
determined by $Y$. The number $r_0 \notin \rho(V)$ and
we assume without loss that it is less than $v = \min \rho(V)$.
Choose a rational number (in lowest terms) $r_1 = p_1/q_1$ in $(r_0, v)$. By
Theorem~(\ref{thm: rot num}) there is a periodic point $x \in W$
with rotation number $r_1$ and period $q_1.$
If $q_2$ is a sufficiently large prime, then $1/q_2 < v - r_1$ and,
by Corollary~\ref{cor: ess V}, the point $x$ lies in
some $V_0 \in {\cal A}(q_2)$ which is essential in $U$.
Corollary~\ref{cor: V(q)} implies that $\rho(V_0)$ and $\rho(V)$ are disjoint and hence that $V_0 \cap V = \emptyset.$
Every point of $V$ is separated by $V_0$ from the component $Y$ of the
frontier of $U$. Hence $cl(V)$ does not intersect $Y$. The
same argument shows $cl(V)$ does not intersect the other component of
the frontier of $U.$ It follows that $cl(V) \subset U.$
To prove the moreover part of this lemma suppose
$x \in U$ and $\rho_f(x) \notin \rho(\partial U_c).$ Then
by Corollary~\ref{cor: ess V},
for every sufficiently large prime $q,$\ there is an essential $V \in {\cal A}(q)$ that contains $x.$ Also we may assume $1/q$ is less than the distance
from $\rho_f(x)$ to $\rho(\partial U_c)$.
Since $\rho(V)$ has length at most $1/q$ and contains $\rho_f(x)$, it is disjoint
from $\rho(\partial U_c)$ and so $\cl(V) \subset U$ by the first part of this lemma.
\end{proof}
We will write $|\rho(V)|$ for the
length of the interval $\rho(V)$.
\begin{lemma}\label{lem: monotonic} Suppose that $Y$ is a component of the frontier of $U$ in $S^2$ and that
$\{V_i\} $ is an infinite sequence of distinct essential elements
of ${\cal A}(q)$ such that $V_{i+1}$ separates $V_i$ from $Y$. Then
\[
\lim_{i \to \infty} |\rho(V_i)| = 0.
\]
\end{lemma}
\begin{proof}
Let $W_i$ be the open annulus that is the union of $V_1, V_i$ and a closed annulus bounded by an essential curve in $V_1$ and and essential curve in $V_i$. The complementary components of $W_i$ in $U$ [resp. $S^2$] are the component of $U \setminus V_i$ [ resp. $S^2 \setminus V_i$] that contains $Y$ and the component of $U \setminus V_1$ [ resp. $S^2 \setminus V_1$] that contains the other component of the frontier of $U$. Let
$W \subset U$ be the union $\cup_i W_i.$ Since the nested intersection of compact connected sets is compact and connected, $W$ is open and has two
complementary components in $S^2$ so must be an open annulus.
The boundary component $B$ of $\partial W_c$ corresponding to the end of $W$ that is disjoint from $V_1$ has arbitrarily small $f_c$-invariant neighborhoods. It follows that if $x_i \in V_i$ is periodic, then the rotation number of $x_i$ with respect to $f$ converges to the rotation number $a$ of the restriction of $f_c$ to $B$. Theorem~\ref{thm: rot num} therefore implies that the interval $\rho(V_i)$ converges to the point $a$ and so has length tending to zero.
\end{proof}
\begin{lemma}\label{lem: ends}
Suppose $\rho(U)$ is non-trivial and
$Y$ is a component of the frontier of $U$. Let $a$ be the rotation number of the component of $\partial U_c$ corresponding to $Y$.
\begin{enumerate}
\item If $q$ is any
sufficiently large prime, then $Y$ is a frontier component of a (necessarily unique) essential $V_0(q) \in
{\cal A}(q)$ with $V_0(q) \subset U$.
\item $a \in \rho(V_0(q))$.
\item If $a \ne p/q$ for some $0 < p < q$ then $\cl_U(V_0(q')) \subset V_0(q)$ for all sufficiently large $q'$.
\end{enumerate}
\end{lemma}
\begin{proof} The first step in the proof of (1) is to prove that for all sufficiently large $q$ there exists a (necessarily unique) essential $V_0(q) \in
{\cal A}(q)$ that is not separated from $Y$ by any other essential element in ${\cal A}(q)$.
By the moreover part of Lemma~\ref{lem: cl(V) in U} we may assume that there exists an essential $V_1 \in {\cal A}(q)$ whose closure is contained in $U$. Let $\rho(\partial U_c) = \{a,b\}$. We may assume that $q$ is so large that neither $a$ nor $b$ has the form $p/q$ with $0< p < q.$ Choose
$\delta < |a -p/q|, |b-p/q|$ for all $0 < p <q.$
The proof is by contradiction: assuming that no such $V_0(q)$ exists we will inductively define an infinite sequence $\{V_i\} $ of distinct essential elements of ${\cal A}(q)$ such that $V_{i+1}$ separates $V_i$ from $Y$ and such that $|\rho(V_i)| > \delta/2$ in contradiction to Lemma~\ref{lem: monotonic}. It suffices to assume that $V_1,\ldots V_{i-1}$ have been defined for $i \ge 2$, and define $V_i$. By assumption, there exists $V^*_{i} \in {\cal A}(q) $ that separates $V_{i-1}$ from $Y$.
Since $V_i^*$ is contained between two open essential annuli in $U$, each component of
its frontier is contained in the interior of $U$. Lemma~\ref{lem: not in V} implies that the rotation number of the restriction of $f_c$ to a component of $\partial {V^*_i}_c$ has the form $p/q$ with $0 <p < q.$
Choose an essential closed curve $\alpha$ in $V^*_i$ and let $W_i$ be the union of $V^*_i$ with the component of $U \setminus \alpha$ that does not contain $V_{i-1}$. Then $W_i$ is an open annulus whose frontier components are $Y$ and a component of the frontier of $V^*_i$. Theorem ~\ref{thm: rot num} implies that $W_i$ contains a periodic point $z$ whose rotation number has distance less than $\delta/2$ from $a$ and so is not of the form $p/q$.
In particular, $z \in {\cal M}_q$ and so is contained in some $V_i \in {\cal A}(q)$ by Lemma~\ref{lem: not in V}. Proposition~\ref{prop: cA_q}-(2) implies that $V_i$ is essential and hence separates $Y$ from $V_{i-1}$. The rotation number of the restriction of $f_c$ to a component of $\partial {V_i}_c$ has the form $p/q$ with $0 <p < q$ for the same reason that components of $\partial {V^*_i}_c$ satisfy this property. Since $z \in V_i$, it follows that $|\rho(V_i)| > \delta/2$. This completes the induction step and hence the proof that there is a unique essential $V_0(q) \in {\cal A}(q)$ that is not separated from $Y$ by any essential element of ${\cal A}(q)$.
For the remainder of the proof we view $V_0(q)$ as an essential open sub-annulus of $U_c$. Let $\partial_0 U_c$ be the component of $\partial U_c$ corresponding to $Y$.
Since there exists $V_1 \in {\cal A}(q)$ whose closure is contained in $U$, the component $B(q)$ of $\fr(V_0(q))$ which is separated from
$\partial_0 U_c$ by $V_0(q)$ is contained in $U$. Lemma~\ref{lem: not in V} implies that if $q \ne q'$ then $B(q) \cap B(q') = \emptyset$. Let $W(q)$ be the open sub-annulus of $U_c$ bounded by $\partial_0U_c$ and $B(q)$ and note that either $\cl_U(W(q)) \subset W(q')$ or $\cl_U(W(q')) \subset W(q)$.
Theorem ~\ref{thm: rot num} implies that $W(q)$ contains a periodic point $w$ whose rotation number is arbitrarily close to $a$ and in particular is not of the form $p/q$. Lemma~\ref{lem: not in V} and Proposition~\ref{prop: cA_q}-(2) imply that $w$ is contained in some element of $ {\cal A}(q)$ that is essential and hence must be $V_0(q)$. Theorem~\ref{thm: irr rot num} implies that $a \in \rho(V_0(q))$ which verifies (2). The same argument shows that $V_0(q')$ contains a point in $W(q)$ and hence that $B(q)$ is not contained in the component of $U_c \setminus V_0(q')$ that contains $\partial_0 U_c$. Assuming without loss that $q' > 2/\delta$, \ $\rho(V_0(q'))$ does not contain any element
of the form $p/q$ with $0 < p < q$. It follows that $B(q)$ is disjoint from $V_0(q')$ and so is contained in the component of $U_c \setminus V_0(q')$ that is separated from $\partial_0 U_c$ by $B(q')$. In other words, $\cl_UW(q') \subset W(q)$. Item (3) will follow once we
prove that $W(q) = V_0(q)$ (which is equivalent to showing that
$\partial_0 U_c$ is the boundary component $B'(q)$ of $V_0(q)$ which is not
$B(q)$).
We claim that if $P$ is an open set in $W(q)\setminus V_0(q)$ then $P \cap W(q') = \emptyset$. If this fails then the open set $P \cap W(q')$ would contain a birecurrent point $x$, which by Lemma~\ref{U covers} is contained in some, necessarily inessential, $V \in {\cal A}(q)$.
Lemma~\ref{lem: not in V} and Proposition~\ref{prop: cA_q} therefore
imply that $\rho(x)$ has
the form $p/q$ with $0<p<q$. It follows that $x$ is not contained in $V_0(q')$ but is contained in an essential element of ${\cal A}(q')$. This contradicts the assumption that $x \in W(q')$ and so verifies the claim.
One immediate consequence of the claim is that $\partial _0 U_c \subset B'(q)$. Another is that each point in $B'(q)$ has a neighborhood that does not intersect any element of ${\cal A}(q)$ other than $V_0(q)$. It follows that $\partial _0 U_c \subset B'(q)\subset \partial_0 U_c \cup
\Fix(f^q)$. Since (c.f. the proof of Lemma~\ref{lem: not in V})
rotation number is constant on each component of $\partial_0 U_c \cup
\Fix(f_c^q) = \partial_0 U_c \cup
\Fix(f^q)$, it must be zero on $B'(q)$. Since $\Fix(f) = \emptyset$
this implies that $B'(q)
\cap \Fix(f^q) = \emptyset$ and hence that $B'(q) = \partial_0 U_c$.
\end{proof}
\begin{remark} In the following definition we assume that $\rho(U)$ is non-trivial. If $\rho(U)$ is trivial then both ends of $U$ are singular by Lemma~\ref{lem: U rot int} so $\Fix(F)$ is a pair of points and $U$ is the unique element of ${\cal A}$.
\end{remark}
\begin{defns} \label{defn: hat Y} Suppose that $\rho(U)$ is non-trivial, that $Y_0$ and $Y_1$ are the frontier components
of $U \in {\cal A}$ and that $a_i$ is the rotation number of the component
of $\partial U_c$ determined by $Y_i$. Choose $Q = Q(U)$ so that:
\begin{itemize}
\item If $a_i = p/q$ for some $0 < p < q$ then $q < Q$.
\item
For every prime $q > Q$ there is an
essential element $V \in {\cal A}(q)$
such that $cl(V) \subset U$. (See Lemma~\ref{lem: cl(V) in U}).)
\item For every prime $q > Q$ there are distinct elements $V_0(q),
V_1(q) \in {\cal A}(q)$ (as in Lemma~\ref{lem: ends}) that are contained
in $U$ such that $\fr(V_i(q)) = Y_i \cup B_i(q)$ for $i=0,1$ where $B_0(q)
\cap B_1(q) = \emptyset$.
\end{itemize}
Define
\[
\hat Y_i = \bigcap_{q > Q} \cl( V_i(q)).
\]
and $\Check U = \cl(U) \setminus ( \hat Y_0 \cup \hat Y_1) \subset U$.
\end{defns}
\begin{lemma}\label{lem: Check U} Assume notation as above.
\begin{enumerate}
\item $\hat Y_i$ is well-defined, i.e., independent of the
choice of $Q$.
\item$\Check U$ and $\Check U \cup (\hat Y_i \cap U)$ are
essential open $f$-invariant annuli.
\item If $a_i = 0$ then $\hat Y_i \cap U$ has measure $0.$
\item If $a_i \ne 0$ then $ \hat Y_{i} \cap U \subset {\cal W}_0$.
\item $\rho(y) = a_i$ for each $y \in \hat Y_i \cap U$.
\end{enumerate}
\end{lemma}
\begin{remark} For any $x \in \Check U$ the $\omega$-limit set
$\omega(F, x)$ is contained in $U$ because the orbit of $x$ is
separated from the frontier of $U$ by $V_0(q) \cup V_1(q)$ for some
$q$ Lemma~\ref{lem: not in V}-(1) therefore implies that every
point in $\Check U$ has non-zero rotation number. If $a_0 = a_1 = 0$
then item (5) of Lemma~\ref{lem: Check U} implies that $\Check U$ is
the subset of $U$ with non-zero rotation number.
\end{remark}
\begin{proof} By part (3) of Lemma~\ref{lem: ends}
there is a sequence of primes $q_j \to \infty$ such that
\[
\hat Y_i = \bigcap_{q_j} \cl( V_i(q_j))
\]
and
$$
V_i(q_{j+1}) \cup B_i(q_{j+1}) )\subset V_i(q_j)
$$
for all $j$, where $B_i(q)$ denotes the component of
the frontier of $V_i(q)$ which lies in $U$.
This proves that $\hat Y_i$ does not depend on the choice of $Q$ in its definition and so is well-defined.
Let $Z_i(q)$ be the component of $S^2 \setminus B_i(q)$ that contains $Y_i$. Then $\cl(Z_i(q_{j+1})) = Z_i(q_{j+1}) \cup B_i(q_{j+1}) \subset Z_i(q_j)$ for the
sequence of primes $\{q_j\}$ chosen above. Define
$$
\hat Z_i = \bigcap_{q_j} \cl(Z_i(q_j)) = \bigcap_{q>Q} \cl(Z_i(q))
$$
Then $\hat Z_0$ and $\hat Z_1$ are disjoint, compact, connected sets and
$\Check U = \cl(U) \setminus ( \hat Y_0 \cup \hat Y_1)
= S^2 \setminus ( \hat Z_0 \cup \hat Z_1).$ Thus $\Check U$
is an open subsurface of $S^2$ with two ends and hence an
annulus. The set $\Check U$ separates $\hat Z_0$ and
$\hat Z_1$ each of which contains a point of $\Fix(F).$
Hence $\Check U$ is essential in ${\cal M}$ and therefore in $U$.
It is $f$-invariant by part (1) of Lemma~(\ref{lem: ess => inv}).
Having verified that $\Check U$ is an essential open $f$-invariant
annulus, the same is true for $\Check U \cup (\hat Y_i \cap U)$, which
is the union of $\Check U$ with the essential sub-annulus of $U$
bounded by $Y_i$ and an essential simple closed curve in $\Check U$. This
completes the proof of (2).
Item (5) follows from the fact that $a_i \in \rho(V_0(q))$ and the
fact that $|\rho(V_0(q)| \le 1/q$. Item (3) then follows from
Proposition~\ref{prop: +meas} and that fact that $\Fix(f) =
\emptyset$.
For (4) suppose that $a_i \ne 0$ and that $x \in \hat Y_i$.
If the $\omega$-limit set of $x$ contains a point in $U$ then $x \in
{\cal W}_0$ by Remark~\ref{rem: free disk}. Otherwise there is a non-fixed point
$z$ in $\omega(x, f_c)$. If $D_0$ is a disk neighborhood of $z$ then
$D_0 \cap U$ is a free disk that the orbit of $x$ intersects more than
once and again $x \in {\cal W}_0$.
\end{proof}
We are now prepared to complete the proof of Theorem~(\ref{thm: annuli}). For the definition of free disk recurrent and weakly free disk recurrent see Definition~\ref{defn: omega-free}.
\bigskip
\noindent
{\bf Theorem~\ref{thm: annuli}.} {\em Suppose $F \in \Diff_\mu(S^2)$ has
entropy zero. Let $f = F|_{{\cal M}}$ where ${\cal M} = S^2
\setminus \Fix(F)$. Then there is a countable collection ${\cal A}$ of pairwise disjoint
open $f$-invariant annuli such that
\begin{enumerate}
\item ${\cal U} = \bigcup_{U \in {\cal A}} U$ is the the set ${\cal W}$ of weakly free disk
recurrent points for $f$.
\item If $U \in {\cal A}$ and $z$ is in the frontier of $U$ in $S^2$, there
are components $F_+(z)$ and $F_-(z)$ of $\Fix(F)$ so that $\omega(F,z)
\subset F_+(z) $ and $\alpha(F,z) \subset F_-(z)$.
\item For each $U \in {\cal A}$ and each component $C_{{\cal M}}$ of the frontier of
$U$ in ${\cal M}$, $F_+(z)$ and $F_-(z)$ are independent of the choice of $z
\in C_{{\cal M}}$.
\end{enumerate}
}
\begin{proof} By items (2) and (3) of Proposition~\ref{intermediate} it suffices
to prove (1).
Given $x \not \in {\cal U}$, choose a lift $\tilde x \in H$ and let $\tilde C$ be the home domain for $\tilde x$. If there is a free disk $D$ and $n > 0$ such that $x,f^n(x) \in D$, then there is a lift $\tilde D$ of $D$ that contains $\tilde x$ and a covering translation $T$ such that $\tilde f_{\tilde C}^n(x) \in T(D)$. Thus $T$ is an $\tilde f_{\tilde C}$-near cycle for $\tilde x$. Since $\tilde x$ and $\tilde f_{\tilde C}^n(\tilde x)$ are both contained in $\tilde U$, $T$ preserves $U$ and so preserves $\tilde C$ in contradiction to Lemma~\ref{more home domains}-(2). We conclude that there is no such $n$ and $D$ which proves that $x \not \in {\cal W}_0$ . Thus ${\cal W}_0 \subset {\cal U}$ and the interior of the closure in ${\cal M}$ of any component of ${\cal W}_0$ is contained in some $U \in {\cal A}$ by Lemma~\ref{U tilde properties}-(5). This proves that ${\cal W} \subset {\cal U}$.
To prove the converse note that the $\omega$-limit set of any point in
$\Check U$ lies in $U$ and hence contains points that are not fixed by
$f$. Remark~\ref{rem: free disk} therefore implies that $\Check U
\subset {\cal W}_0$.
If both $a_0$ and $a_1$ are non-zero then $U = \Check U \cup (U
\cap (\hat Y_0 \cup \hat Y_1)) \subset {\cal W}_0 \subset {\cal W}$ by item (4) of Lemma~\ref{lem: Check U} . If both $a_0$ and $a_1$ are zero then $\Check U$ is dense in $U$ by item (3) of Lemma~\ref{lem: Check U}. Thus $U \subset \cl(\Check U) \subset {\cal W}$ since $\Check U$ is a connected (item (2) of Lemma~\ref{lem: Check U}) subset of ${\cal W}_0$. For the remaining case we may assume
that $a_0 =0$ and $a_1 \ne 0$. Then $U \subset\cl(\Check U
\cup (\hat Y_1 \cap U))\subset {\cal W}$ because $\Check U
\cup (\hat Y_1 \cap U)$ is a connected subset of ${\cal W}_0$.
\end{proof}
\begin{prop}\label{prop: C(x)-pre}
Each $x \in \Check U$ is contained in an $f$-invariant set $C(x)
\subset U$ with the following properties:
\begin{enumerate}
\item Each $C(x)$ is compact, connected and essential in $U$, i.e.
its complement in $U$ has two components and it separates the ends
of $U$.
\item If $x,y \in \Check U$ then $C(x)$ and $C(y)$ are equal or disjoint.
\item The rotation number $\rho$ is constant on
$C(x)$. In fact, each $C(x)$ is a connected component of a level set
of the function $\rho.$
\item $\rho$ is continuous on $U$ and
extends continuously to $cl(U)$ assigning (in the notation of Definition~\ref{defn: hat Y}) $a_i$ to $\hat Y_i$.
\end{enumerate}
\end{prop}
\begin{proof} The definition of $C(x)$ is very similar to the definition of $\hat Y_i$ in Definition~\ref{defn: hat Y}.
We specify $Q = Q(x) $ by a series of largeness conditions. By Lemma~\ref{lem: ends}-(3) and the assumption that $x \in \Check U$, we may assume that $x \not \in V_0(q) \cup V_1(q)$ for $q \ge Q$. In particular, $\omega(x) \subset U$. We may also assume that $\rho(x) \ne p/q$ for $q \ge Q$ and $0 < p < q$. Lemma~\ref{lem: not in V} therefore implies that $x$ is contained in some $V(q,x) \in {\cal A}(q)$ which is essential by for Lemma~\ref{prop: cA_q}. Since $x \not \in V_0(q) \cup V_1(q)$, we have $\cl(V(q,x)) \subset U$.
Define
\[
C(x) = \bigcap_{q > Q} \cl( V(q,x)).
\]
Given $q$ let $\delta$ be the minimum value of $|p/q - \rho(x)|$ for
$0 < p < q$. If $q' >1/\delta$ then $\rho(V(q',x))$ does not contain
$p/q$ for $0 < p < q$ and so does not contain any points in the
frontier of $V(x,q)$. It follows that $\cl(V(q',x)) \subset V(q,x)$.
We may therefore choose a sequence of primes $q_j \to \infty$ such
that
\[
C(x) = \bigcap_{q_j} \cl( V(q_j,x))
\]
and
$$
\cl(V(q_{j+1},x)) \subset V(q_j,x)
$$
for all $j$.
This proves that $C(x)$ does not depend on the choice of $Q$ in its definition and so is well-defined.
Items (1) and (2) are obvious from the construction. The first part of (3) follows from the fact that $|\rho(V(q,x))| \le 1/q$.
If $y \in V(q,x)$ then $V(q,x)$ is a neighborhood of $y$ on which
$\rho(x)$ is in $\rho(V_k)$. This proves
$\rho$ is continuous at $y$. Hence $\rho$ is continuous on
$\Check U.$
Assume notation as in Definition~\ref{defn: hat Y}. The sets
$V_i(q)$ form a neighborhood basis of $\hat Y_i$ and
$\hat Y_i$ contains $Y_i$.
Since for $x \in V_i(q),$ both $a_i$
and $\rho_{f}(x)$ are in $\rho(V_i(q))$ and $|\rho(V_i(q))| \le 1/q,$
it follows that the extension of $\rho$ to $\hat Y_i$ assigning
the value $a_i$ to every point of $\hat Y_i$ is also continuous. This completes the proof of (4).
Finally observe that if $y \notin C(x)$ then there exists $q > Q$ such
that $y \notin \cl(V(q,x)).$ Then the frontier of $V(q,x)$ separates
$C(x)$ and $y$ and this frontier consists of points whose rotation
number is not equal to $\rho(x)$ so $y$ is not in the same connected
component as $x$ of the level set of $\rho$. It follows that $C(x)$
is a connected component of this level set, proving part (3).
\end{proof}
If $x$ is in a component of the frontier of $\Check U$ it is natural
to define $C(x)$ to be the set $\hat Y_i$ (as in
Definition~(\ref{defn: hat Y})) which contains it. This set $C(x)$ is compact
and connected. As noted in the previous theorem the function $\rho_f$
extends continuously, but this $C(x)$ may not necessarily
separate the sphere $S^2$. In fact it may be
a single point.
We have completed most of the work need to prove the following.
\bigskip
\noindent
{\bf Theorem \ref{thm: C(x)}.}
{\em
Suppose $F \in \Diff_\mu(S^2)$ has entropy zero. Let $f = F|_{{\cal M}}$ where
${\cal M} = S^2 \setminus \Fix(F)$ and let ${\cal A}$ be as in Theorem~\ref{thm: annuli}.
Then for each $U \in {\cal A}$ there is a partition of $cl(U)\subset S^2$ into a family ${\cal C}$
of closed $F$-invariant sets
with the following properties:
\begin{itemize}
\item Each $C \in {\cal C}$ is compact and connected.
\item There are two elements of ${\cal C}$ (which may coincide), called {\em ends}
and denoted $C_0$ and $C_1,$
each of which contains a component of the frontier of $U$.
Every element of ${\cal C}$ which is not an end
is a subset of $U$ and is called {\em interior}.
Each interior $C$ is essential in $U$, i.e.
its complement (in $U$) has two components and it separates $C_0$ and $C_1$.
\item The rotation number $\rho_f(x) \in \mathbb R/\mathbb Z$ is well defined and
constant on any interior $C$. In fact, each $C$ is a connected
component of a level set of $\rho_f(x).$ Moreover, $\rho_f(x)$ is
continuous on $U$
and has a unique continuous
extension to $cl(U)$. The value of this extension on $C_i, \ i=0,1$ is
the element of $\rho(\partial U_c)$ corresponding to the component of
the frontier of $U$ contained in $C_i.$
\end{itemize}
}
\begin{proof}
We first define the two end elements of ${\cal C}$ by letting
$C_0 = \hat Y_0$ and $C_1 = \hat Y_1.$
We define the collection ${\cal C}$ to consist of the these two sets
together with the closed sets $C(x)$ for $x \in \Check U.$
Proposition~(\ref{lem: Check U}) and
Proposition~(\ref{prop: C(x)-pre}) (2) imply this is a partition of $U$.
Proposition~(\ref{prop: C(x)-pre}) (1) says that each $C$ is compact
and connected and that each interior $C$ is essential and separates $U$.
Parts (3) and (4) of Proposition~(\ref{prop: C(x)-pre}) complete
the proof.
\end{proof}
\section{The proof of Theorem~\ref{thm: ent0}.}
Recall that a group $G$ is called {\em indicable} if
there is a non-trivial homomorphism $\phi: G \to \mathbb Z.$
We say $G$ is {\em virtually indicable} if it has
a finite index subgroup which is indicable.
\begin{prop}\label{prop: +ent}
Suppose that $S$ is a surface and
$F: S \to S$ is $C^{1+\epsilon}$ and has positive
topological entropy. Then every finitely generated
infinite subgroup $H$ of the centralizer $Z(F)$of $F$
is virtually indicable and has a finite index subgroup that
has a global fixed point.
\end{prop}
\begin{proof}
A result of Katok \cite{katok:horseshoe} asserts that $F^q$ has a
hyperbolic fixed point $p$ for some $q\ge 1.$ The orbit of $p$ under
$H$ consists of hyperbolic fixed points of $F^q$ at which the
derivative of $DF^q$ has the same eigenvalues as $DF_p^q.$ If the
$H$ orbit of $p$ were infinite, continuity of the derivative would
imply that at any limit point of this orbit $DF^q$ would have the
same eigenvalues and in particular would be hyperbolic. But this is
impossible since hyperbolic fixed points are isolated. We conclude
the orbit of $p$ under $H$ is finite and hence that the subgroup
$H_0$ of $H$ that fixes $p$ has finite index.
After passing to a further finite index subgroup we may assume that
$Dh_p$ has positive eigenvalues and the same eigenspaces as $DF_p$
for each $h \in H_0$. For each eigenspace the function which
assigns to $h$ the log of the eigenvalue of $Dh_p$ on that
eigenspace is a homomorphism from $H_0$ to $\mathbb R$. If this is
non-trivial we are done. Otherwise both eigenvalues are $1$ for
each $Dh_p$. Hence in the appropriate basis
\[
Dh_p =
\begin{pmatrix}
1 & r_h\\
0 & 1
\end{pmatrix}
\]
for some $r_h \in \mathbb R$. The function $h \mapsto r_h$ defines a
homomorphism from $H_0$ to $\mathbb R$, so we are done unless $r_h = 0$ for
all $h\in H_0.$ But in this latter case $Dh_p = I$ for all $h \in H_0$
so we may apply the Thurston stability theorem
(\cite{thurston:stability}; see also Theorem 3.4 of
\cite{franks:distortion}) to conclude there is a non-trivial
homomorphism from $H_0$ to $\mathbb R.$
\end{proof}
In the case of an entropy zero diffeomorphism $F$ the analogous
result does not hold, even when we restrict to the group
of area preserving diffeomorphisms.
\begin{exs}\label{example}
{\rm Let $S = S^2$ be the unit sphere in $\mathbb R^3$. Let $F: S \to S$
be a diffeomorphism whose restriction to each of the level sets
$z =c$ is a rotation of that circle and with the property that
$F = id$ for all points $(x,y,z)$ with $|z| \ge 3/4.$ We assume
that $F$ is not the identity on the equator $z = 0.$ Let
$g: S \to S$ be rotation about the $z$-axis by an angle which
is an irrational multiple of $\pi.$ Let $h:S \to S$ be a
diffeomorphism supported in the interior of the disks $|z| > 3/4$
with the property that $h$ preserves area and the $h$-orbits of $(0,0,1)$
and $(0,0,-1)$ are infinite. Let $G$ be the group of all rotations
about the $z$-axis through angles which are rational multiples
of $\pi.$}
\begin{enumerate}
\item The group $H$ generated by $g$ and $h$ lies in the
centralizer $Z(F)$ of $F$ but has no finite index subgroup
with a global fixed point.
\item The group $G$ is a subgroup of $Z(g)$. Every element
of $G$ has finite order so there are no
non-trivial homomorphisms from any subgroup of
$G$ to $\mathbb R$ and hence $G$ is not virtually indicable.
\end{enumerate}
\end{exs}
\trycomment{ Is there an example which is not virtually indicable
and which has no finite index subgroup with a global fixed point?}
The first example above shows that we cannot generalize
Proposition~(\ref{prop: +ent}) to the centralizer
of a diffeomorphism $F$ with zero entropy, even in the
group of area preserving diffeomorphisms. The second example
shows the necessity of the hypothesis of finitely generated in
the following.
\bigskip
\noindent
{\bf Theorem~\ref{thm: ent0}.}
{\em If $F \in \Diff_{\mu}(S^2)$ has infinite order then each finitely
generated infinite subgroup $H$ of $Z(F)$ is virtually indicable.}
\begin{proof}
The case that $F$ has positive entropy is covered
by Proposition~(\ref{prop: +ent}) so we need only
consider the case when $F$ has entropy zero.
We assume that every finite index subgroup of $H$
admits only the trivial homomorphism to $\mathbb R$ and
show this leads to a contradiction.
Recall that ${\cal A}$ denotes the family of $F$-invariant
annuli guaranteed by Theorem~(\ref{thm: annuli}).
There is a special case when every $U \in {\cal A}$
has trivial rotation interval $\rho(U).$
We postpone this until after the more general case.
Let $U \in {\cal A}$ be an element of ${\cal A}$
with the rotation interval $\rho(U)$ non-trivial.
By Remark~\ref{first centralizer invariance}, each $h \in H$ permutes the open annuli in the family
${\cal A}.$ In particular,
for any $h \in H$ the open annuli $U$ and $h(U)$ must be disjoint or
equal.
Since elements of $H$ preserve area the $H$ orbit of the open set $U$
must be finite. We let $H'$ be the finite index subgroup of $H$ which
leaves $U$ invariant. Let $x \in U$ be a point with $\rho_F(x) =
\lambda$ which is irrational. Such a point exists by Theorem~
(\ref{thm: irr rot num}) since the rotation interval of $F_c:U_c \to
U_c$ is non-trivial. The set $C(x)$ was defined in
Proposition~(\ref{prop: C(x)-pre}) to be the intersection of a
family $\{V_n\}$ of essential $F$-invariant annuli containing $x$ with
the property that the rotation interval (under $F|_U$) is contained in
an interval $I_n$ with $\bigcap I_n = \{\lambda\}$. The annulus $V_n$
is an element of ${\cal A}(q)$ for some $q,$ and, as above,
$h(V_n)$ is either equal to $V_n$ or disjoint from it.
However, $h$ cannot map an essential open subannulus of $U$ to
a disjoint subannulus since that would require that it map an open
subset of $U$ onto a proper subset contradicting the property that
$h$ preserves area.
Choose one component, $V_n^+,$ of the complement of $C(x)$
in $V_n$ in such
a way that $V_{n+1}^+ \subset V_n^+$, i.e., always choose the component
on the same side of $C(x).$ Let
$\bar A_n$ denote $V_n^+$ with its ends compactified
by the prime end compactification.
Let $\partial^+ A$ denote the
circle of prime ends added to the end corresponding to $C(x).$ The
natural identification of these circles for different
$n$ is reflected in the notation which is independent of $n.$
Let $A_n = V_nA^+ \cup \partial^+ A,$ i.e. $V_n^+$ with only
one end (the one corresponding to $C(x)$) compactified. Then
$A_{n+1} \subset A_n$ and
\[
\bigcap_{n>0} A_n = \partial^+ A.
\]
Let $\bar f: \bar A_n \to \bar A_n$
and $\bar h: \bar A_n \to \bar A_n$ denote the natural extensions of
$F$ and $h \in H'$ to $\bar A_n.$
The rotation number $\rho(\bar f|_{\partial^+ A})$ of the
restriction of $\bar f$ to
$\partial^+ A$ must be $\lambda.$ This is because if it were not
and $p/q$ is between $\rho(\bar f|_{\partial^+ A})$ and $\lambda$ then
by Theorem~(\ref{thm: rot num}) applied to $ \bar A_n$
there would be periodic points in the
interior of $\bar A_n \subset V_n$ with rotation number $p/q$ for all $n$, a
contradiction.
For each $n$ there is a homomorphism $\phi_n: H' \to S^1 = \mathbb R/\mathbb Z$
given by $h \mapsto \rho_\mu(h|_{\bar A_n})$ where $\rho_\mu(h|_{\bar A_n})$
denotes the mean rotation number of $h$ on the annulus $\bar A_n$
(see (1.1) of \cite{franks:openann}).
Let $H''$ denote the subgroup of $H'$ which is the kernel of the
canonical homomorphism from $H'$ to its abelianization. Then
$\rho_\mu(h|_{\bar A_n}) = 0$
for all $h \in H''.$ Also the abelianization
of $H'$ must be finite since this is one of the equivalent conditions
for $H'$ not to be indicable. Therefore $H''$ has finite index in $H'$ (and
hence in $H$).
Since $\rho_\mu(\bar h|_{\bar A_n})=0$ for each $n$ and
each $\bar h \in H''$
we conclude from Theorem~(2.1) of \cite{franks:openann}
that it has a fixed point $y_n$ in
$\bar A_n$ for all $n$. Let $B$ be the closed disk which is the union
of $\partial^+ A$ and the component of the complement
of $C(x)$ in $S^2$ which contains $V_n^+.$
Lemma~(\ref{lem: frontier fp}) applied to the open annulus
$V_n^+$ as a subset of $B$ implies that $cl_B(V_n^+)$ contains
a fixed point $x_n$ of $\bar h.$
Taking the limit of a subsequence we note that each $\bar h \in
H''$ has a fixed point in $\partial^+ A.$
But the rotation number of $\bar f$ on $\partial^+ A$ is irrational
so $\bar f$ has a unique invariant minimal set which is the omega
limit set $\omega(x, \bar f)$ for each $x \in \partial^+ A$. Since
$\bar f$ preserves $\Fix(\bar h)$ we conclude this minimal set is in
$\Fix(\bar h)$ for every $\bar h \in H''$.
We have found a prime end (in fact infinitely many ) in
$\partial^+ A$ which is fixed for
each $\bar h \in H''$. It follows from
Corollary~(\ref{cor: frontier gfp}) that there is a point of $\Fix(H'')$
in $cl(V_n^+)$ for each $n$. Taking the limit of a subsequence again
we find a point of $\Fix(H'')$ which lies in $\bigcap_n cl(V_n^+) =C(x).$
Choosing an infinite collection $\{\lambda_i\}$ of distinct irrationals in
the rotation interval of $F|_U$ and repeating the construction
we obtain an infinite collection of global fixed points for
$H''$ with distinct rotation numbers for $F|_U.$ They must
possess a limit point in $\Fix(H'').$
Proposition~(3.1) of \cite{fh:morita} asserts that if there is an
accumulation point of $\Fix(H'')$ then there is a homomorphism from
$H''$ to $\mathbb R$. So $H''$ is indicable.
We are left with addressing the special case that every $U
\in {\cal A}$ has trivial $\rho(U).$ Note that when
$\rho(U)$ is trivial, by Proposition~(\ref{prop: +meas})
it cannot consist of the single element $0 \in \mathbb R/\mathbb Z.$ Note also
that if every $\rho(U)$ is trivial then $\Fix(F)$ has precisely
two components and ${\cal A}$ contains only one element. This is
because there is at least one $U \in {\cal A}$ and its complement has two
components, and if $\Fix(F)$ has three components, two of them must
be in one component of the complement of $U$ which implies the
corresponding end of $U$ is non-singular. Non-singular ends
always imply $\rho(U)$ is non-trivial by Lemma~(\ref{lem: U rot int}).
Moreover, the two components of $\Fix(F)$ must be single points
since for an end with more than one fixed
point in its frontier, the compactification of that end will be
the prime end compactification and $F_c: U_c \to U_c$ will be the identity
on that boundary component. This would imply that $\rho(U) =\{0\}$,
a contradiction to the remarks above.
It follows that there is only one $U \in {\cal A}$
and it is the complement of the two fixed points of $F.$
As a consequence of Theorem~(\ref{thm: rot num}) and
Proposition~(\ref{prop: +meas}) we conclude that almost all points of
$S^2$ have the same rotation number (with respect to the two fixed
points). First assume that number is rational, say $p/q$ in lowest
terms. In that case $g = F^q$ necessarily has more than
$2$ components to its fixed point set (by Theorem~(\ref{thm: rot num})
again). We consider $H \subset Z(g)$
and the argument from the first part of this theorem can be applied.
Finally suppose that $\rho(U)$ is a single irrational $\lambda.$
Passing to a subgroup of index
$2$ if necessary we may assume that $H$ fixes the two fixed points of $F$.
Blowing up the two points we obtain a homeomorphism (also called $F$)
on an annulus. The restriction to the boundary component corresponding
to the fixed point $x$ is conjugate to the projectivization of $DF_x$.
It must have rotation number $\lambda$ since otherwise there would be
an additional periodic point by Theorem~(\ref{thm: rot num}).
This map on the boundary circle is the projectivization of an element
of $SL(2,\mathbb R)$, i.e. a fractional linear transformation. Since its
rotation number is irrational, in appropriately chosen coordinates it
is an irrational rotation. It follows that the restrictions of
blow-ups of elements of $H$ to this circle are rotations, since the
centralizer of an irrational rotation consists of rotations.
Therefore this group of restrictions, which is finitely generated
because it is the image under a homomorphism of a finitely generated
group, is abelian. Since since it admits no non-trivial homomorphisms
to $\mathbb R$ it must be finite. We conclude there is a finite index
subgroup $H'$ of $H$ whose restrictions to the boundary circle are all
the identity. It follows that for any $h \in H'$ we have $Dh_x = I$
(since the projectivization of $Dh_x$ is the identity and $det(Dh_x) =
1$). We can then apply the Thurston stability theorem
(\cite{thurston:stability}) to arrive at a contradiction to the
assumption that there are no non-trivial homomorphisms from $H'$ to
$\mathbb R.$
\end{proof}
We now provide the proof of Corollary~(\ref{no actions}).
\bigskip
\noindent
{\bf Corollary~\ref{no actions}.}
{\em
If $\Sigma_g$ is the closed orientable surface of genus $g \ge 2$ then at least one of the following holds.
\begin{enumerate}
\item No finite index subgroup of $\mcg(\Sigma_g)$ acts faithfully on $S^2$ by area preserving diffeomorphisms.
\item For all \ $1 \le k \le g-1$, there is an indicable finite index subgroup $\Gamma$ of the bounded mapping class group $\mcg(S_k,\partial S_k)$ where $S_k$ is the surface with genus $k$ and connected non-empty boundary.
\end{enumerate}
}
\begin{proof}
We assume that (1) fails, i.e., that there is a finite index subgroup
$G$ of $\mcg(\Sigma_g)$ which acts faithfully on $S^2$ by area
preserving diffeomorphisms, and show that this implies (2).
Suppose \ $1 \le k \le g-1$ and $S$ is the compact surface with
genus $k$ and a connected non-empty boundary, $\partial S.$
We assume $S$ is embedded in $\Sigma_g$ with $\partial S$ a separating
closed curve and let $S'$ be the closure of complement of $S$,
a surface with genus $g -k$ and boundary $\partial S.$
There is a natural embedding of $\mcg(S, \partial S)$
into $\mcg(\Sigma_g)$ obtained
by extending a representative of an element of $\mcg(S, \partial S)$ to
all of $\Sigma_g$ by letting it be the identity on the complement
of $S$. Similarly there is a natural embedding of $\mcg(S', \partial S')$
into $\mcg(\Sigma_g)$. If $\Gamma_0$ and $\Gamma_0'$ are the images of
these two embeddings it is clear that every element of $\Gamma_0$ commutes
with every element of $\Gamma_0'$ since they have representatives in
$\Diff(\Sigma_g)$ which commute.
We let $\Gamma_1 = \Gamma_0 \cap G$ and $\Gamma_1' = \Gamma_0' \cap
G$. Since $\Gamma_1'$ has finite index in $\mcg(S', \partial S)$ it
contains an element $\gamma$ of infinite order. Suppose $\phi: G \to
\Diff_\mu(S^2)$ is the injective homomorphism defining the action of
$G$. Let $F = \phi(\gamma)$. Then $\phi(\Gamma_1)$ is in the
centralizer $Z(F).$ According to Theorem~(\ref{thm: ent0}) $\Gamma_1$
is virtually indicable. Therefore there is an indicable $\Gamma$ of
finite index in $\Gamma_1$.
\end{proof}
\bibliographystyle{plain}
|
\section{Introduction}\label{s1}
CMV operators are a special class of unitary semi-infinite or doubly-infinite five-diagonal matrices which received enormous attention in recent years. We refer to \eqref{2.8} and \eqref{3.18} for the explicit form of doubly infinite CMV operators on ${\mathbb{Z}}$ in the case of scalar, respectively, matrix-valued Verblunsky coefficients. For the corresponding half-lattice CMV operators we refer to \eqref{2.14} and \eqref{3.26}.
The actual history of CMV operators (with scalar Verblunsky coefficients) is somewhat intriguing: The corresponding unitary semi-infinite five-diagonal matrices were first introduced in 1991 by Bunse--Gerstner and Elsner \cite{BGE91}, and subsequently discussed in detail by Watkins \cite{Wa93} in 1993 (cf.\ the discussion in Simon
\cite{Si06}). They were subsequently rediscovered by Cantero, Moral, and Vel\'azquez (CMV) in \cite{CMV03}. In \cite[Sects.\ 4.5, 10.5]{Si04}, Simon introduced the corresponding notion of unitary doubly infinite five-diagonal matrices and coined the term ``extended'' CMV matrices. For simplicity, we will just speak of CMV operators, irrespective of whether or not they are half-lattice or full-lattice operators. We also note that in a context different from orthogonal polynomials on the unit circle, Bourget, Howland, and Joye \cite{BHJ03} introduced a family of doubly infinite matrices with three sets of parameters which, for special choices of the parameters, reduces to two-sided CMV matrices on ${\mathbb{Z}}$. Moreover, it is possible to connect unitary block Jacobi matrices to the trigonometric moment problem (and hence to CMV matrices) as discussed by Berezansky and Dudkin \cite{BD05}, \cite{BD06}.
The relevance of this unitary operator ${\mathbb{U}}$ on $\ell^2({\mathbb{Z}})^m$, more precisely, the relevance of the corresponding half-lattice CMV operator ${\mathbb{U}}_{+,0}$ in
$\ell^2({\mathbb{N}}_0)^m$ is derived from its intimate relationship with the trigonometric moment problem and hence with finite measures on the unit circle ${\partial\hspace*{.2mm}\mathbb{D}}$. (Here
${\mathbb{N}}_0={\mathbb{N}}\cup\{0\}$.) Following \cite{CGZ07}, \cite{CGZ08}, \cite{GZ06},
\cite{GZ06a}, and \cite{Zi08}, this will be reviewed in some detail, and also extended in certain respects, in Sections \ref{s2} and \ref{s3}, as this material is of fundamental importance to the principal topics (such as decoupling of full-lattice CMV operators into direct sums of left and right half-lattice CMV operators and a similar result for associated Green's functions) discussed in this paper, but we also refer to the monumental two-volume treatise by Simon \cite{Si04} (see also \cite{Si04b} and \cite{Si05}) and the exhaustive bibliography therein. For classical results on orthogonal polynomials on the unit circle we refer, for instance, to \cite{Ak65}, \cite{Ge46}--\cite{Ge61}, \cite{Kr45},
\cite{Sz20}--\cite{Sz78}, \cite{Ve35}, \cite{Ve36}. More recent references relevant to the spectral theoretic content of this paper are \cite{De07}, \cite{GJ96}--\cite{GT94},
\cite{GZ06}, \cite{GZ06a}, \cite{GN01}, \cite{PY04}, \cite{Si04a}, and \cite{Zi08}. The
full-lattice CMV operators ${\mathbb{U}}$ on ${\mathbb{Z}}$ are closely related to an important, and only recently intensively studied, completely integrable nonabelian version of the defocusing nonlinear Schr\"odinger equation (continuous in time but discrete in space), a special case of the Ablowitz--Ladik system. Relevant references in this context are, for instance,
\cite{AL75}--\cite{APT04}, \cite{GGH05}, \cite{GH05}--\cite{GHMT07b}, \cite{Li05},
\cite{MEKL95}--\cite{Ne06}, \cite{Sc89}, \cite{Ve99}, and the literature cited therein. We emphasize that the case of matrix-valued coefficients $\alpha_k$ is considerably less studied than the case of scalar coefficients.
We should also emphasize that while there is an extensive literature on orthogonal matrix-valued polynomials on the real line and on the unit circle, we refer, for instance, to
\cite{AN84}, \cite{BC92}, \cite[Ch.\ VII]{Be68}, \cite{BG90}, \cite{CFMV03}, \cite{CG06},
\cite{DG92}--\cite{DV95}, \cite{Ge81}, \cite{Ge82}, \cite{Kr49}, \cite{Kr71}, \cite{Le47},
\cite{Lo99}, \cite{Os97}--\cite{Os02}, \cite{Ro90}, \cite{Si06a}, \cite{YM01}--\cite{YK78}, and the large body of literature therein, the case of CMV operators with matrix-valued Verblunsky coefficients appears to be a much less explored frontier. The only references we are aware of in this context are Simon's treatise \cite[Part 1, Sect.\ 2.13]{Si04} and the recent papers \cite{Ar08}, \cite{CGZ07}, \cite{DPS08}, and \cite{Si06}.
Finally, a brief description of the content of each section in this paper: In Section \ref{s2} we review, and in part, extend the basic Weyl--Titchmarsh theory for half-lattice CMV operators with scalar Verblunsky coefficients originally derived in \cite{GZ06}, and recall its intimate connections with transfer matrices and orthogonal Laurent polynomials. The principal result of this section, Theorem \ref{t2.3}, then provides a necessary and sufficient condition for the difference between the full-lattice CMV operator $U$ and its ``decoupling'' into a direct sum of appropriate left and right half-lattice CMV operators to be of rank one. The same result is also derived for the resolvent differences of $U$ and the resolvent of its decoupling into a direct sum of left and right half-lattice CMV operators. Theorem \ref{t2.3} is in sharp contrast to the familiar Jacobi case, since decoupling a full-lattice CMV matrix by changing one of the Verblunsky coefficients
results in a perturbation of rank two. While this difference compared to Jacobi operators was noticed first by Simon \cite[Sect.\ 4.5]{Si04}, we explore it further here and provide a complete discussion of this decoupling phenomenon, including its extension to the matrix-valued case, which represents a new result. We conclude this section with a discussion of half-lattice Green's functions in Lemma \ref{l2.6}, extending a result in
\cite{GZ06}.
In Section \ref{s3} we develop all these results for CMV operators with
$m\times m$, $m\in{\mathbb{N}}$, matrix-valued Verblunsky coefficients. In particular, in
Theorem \ref{t3.6}, the principal result of this section, we provide
a necessary and sufficient condition for the difference between the
full-lattice CMV operator $U$ and its decoupling into a direct sum of
appropriate left and right half-lattice CMV operators to be of minimal
rank $m$.
Finally, Appendix \ref{sA} summarizes basic facts on matrix-valued
Caratheodory and Schur functions relevant to this paper.
\section{CMV operators with scalar coefficients} \label{s2}
This section is devoted to a study of CMV operators associated with
scalar Verblunsky coefficients. We derive a criterion under which
a difference of a full-lattice CMV operator and a direct sum of two
half-lattice CMV operators is of rank one. The same condition will
also imply a similar result for the resolvents of these operators. At
the end of the section we establish relations that hold between
Weyl--Titchmarsh $m$-functions associated with the above operators
and derive explicit expressions for half-lattice Green's matrices.
We start by introducing basic notations used throughout this paper.
Let $\s{{\mathbb{Z}}}$ be the space of complex-valued sequences and
$\lt{{\mathbb{Z}}}\subset \s{{\mathbb{Z}}}$ be the usual Hilbert space of all square
summable complex-valued sequences with scalar product
$(\cdot,\cdot)_{\lt{{\mathbb{Z}}}}$ linear in the second argument. The {\it
standard basis} in $\lt{{\mathbb{Z}}}$ is denoted by
\begin{equation}
\{\delta_k\}_{k\in{\mathbb{Z}}}, \quad
\delta_k=(\dots,0,\dots,0,\underbrace{1}_{k},0,\dots,0,\dots)^\top,
\; k\in{\mathbb{Z}}.
\end{equation}
For $m\in{\mathbb{N}}$ and $J\subseteq{\mathbb{R}}$ an interval, we will identify
$\oplus_{j=1}^m\lt{J\cap{\mathbb{Z}}}$ and $\lt{J\cap{\mathbb{Z}}}\otimes\mathbb{C}^m$ and then
use the simplified notation $\ltm{J\cap{\mathbb{Z}}}$. For simplicity, the
identity operators on $\lt{J\cap{\mathbb{Z}}}$ and $\ltm{J\cap{\mathbb{Z}}}$ are
abbreviated by $I$ without separately indicating its dependence on
$m$ and $J$. The identity $m\times m$ matrix is denoted by $I_m$.
Throughout this section we make the following basic assumption:
\begin{hypothesis} \label{h2.1}
Let $\alpha=\{\alpha_k\}_{k\in{\mathbb{Z}}}\in \s{{\mathbb{Z}}}$ be a sequence of complex
numbers such that
\begin{equation} \label{2.2}
\alpha_k\in\mathbb{D}, \quad k \in {\mathbb{Z}}.
\end{equation}
\end{hypothesis}
Given a sequence $\alpha$ satisfying \eqref{2.2}, we define the
following sequence of positive real numbers $\{\rho_k\}_{k\in{\mathbb{Z}}}$ by
\begin{equation}
\rho_k = \big[1-\abs{\alpha_k}^2\big]^{1/2}, \quad k \in {\mathbb{Z}}. \label{2.3}
\end{equation}
Following Simon \cite{Si04}, we call $\{\alpha_k\}_{k\in{\mathbb{Z}}}$ the
Verblunsky coefficients in honor of Verblunsky's pioneering work in
the theory of orthogonal polynomials on the unit circle \cite{Ve35},
\cite{Ve36}.
Next, we also introduce a sequence of $2\times 2$ unitary matrices
$\Theta_k$ by
\begin{equation} \label{2.4}
\Theta_k = \begin{pmatrix} -\alpha_k & \rho_k \\ \rho_k & \overline{\alpha_k}
\end{pmatrix},
\quad k \in {\mathbb{Z}},
\end{equation}
and two unitary operators $V$ and $W$ on $\lt{{\mathbb{Z}}}$ by their matrix
representations in the standard basis of $\lt{{\mathbb{Z}}}$ as follows,
\begin{align} \label{2.5}
V &= \begin{pmatrix} \ddots & & &
\raisebox{-3mm}[0mm][0mm]{\hspace*{-5mm}\Huge $0$} \\ & \Theta_{2k-2} &
& \\ & & \Theta_{2k} & & \\ &
\raisebox{0mm}[0mm][0mm]{\hspace*{-10mm}\Huge $0$} & & \ddots
\end{pmatrix}, \quad
W = \begin{pmatrix} \ddots & & &
\raisebox{-3mm}[0mm][0mm]{\hspace*{-5mm}\Huge $0$}
\\ & \Theta_{2k-1} & & \\ & & \Theta_{2k+1} & & \\ &
\raisebox{0mm}[0mm][0mm]{\hspace*{-10mm}\Huge $0$} & & \ddots
\end{pmatrix},
\end{align}
where
\begin{align}
\begin{pmatrix}
V_{2k-1,2k-1} & V_{2k-1,2k} \\
V_{2k,2k-1} & V_{2k,2k}
\end{pmatrix} = \Theta_{2k},
\quad
\begin{pmatrix}
W_{2k,2k} & W_{2k,2k+1} \\ W_{2k+1,2k} & W_{2k+1,2k+1}
\end{pmatrix} = \Theta_{2k+1},
\quad k\in{\mathbb{Z}}.
\end{align}
Moreover, we introduce the unitary operator $U$ on $\lt{{\mathbb{Z}}}$ by
\begin{equation} \label{2.7}
U = VW,
\end{equation}
or in matrix form, in the standard basis of $\lt{{\mathbb{Z}}}$, by
\begin{align}
U &= \begin{pmatrix} \ddots &&\hspace*{-8mm}\ddots
&\hspace*{-10mm}\ddots &\hspace*{-12mm}\ddots &\hspace*{-14mm}\ddots
&&& \raisebox{-3mm}[0mm][0mm]{\hspace*{-6mm}{\Huge $0$}}
\\
&0& -\alpha_{0}\rho_{-1} & -\overline{\alpha_{-1}}\alpha_{0} & -\alpha_{1}\rho_{0} &
\rho_{0}\rho_{1}
\\
&& \rho_{-1}\rho_{0} &\overline{\alpha_{-1}}\rho_{0} & -\overline{\alpha_{0}}\alpha_{1} &
\overline{\alpha_{0}}\rho_{1} & 0
\\
&&&0& -\alpha_{2}\rho_{1} & -\overline{\alpha_{1}}\alpha_{2} & -\alpha_{3}\rho_{2} &
\rho_{2}\rho_{3}
\\
&&\raisebox{-4mm}[0mm][0mm]{\hspace*{-6mm}{\Huge $0$}} &&
\rho_{1}\rho_{2} & \overline{\alpha_{1}}\rho_{2} & -\overline{\alpha_{2}}\alpha_{3} &
\overline{\alpha_{2}}\rho_{3}&0
\\
&&&&&\hspace*{-14mm}\ddots &\hspace*{-14mm}\ddots
&\hspace*{-14mm}\ddots &\hspace*{-8mm}\ddots &\ddots
\end{pmatrix} \label{2.8}
\\
&= \rho^- \rho \, \delta_{\rm even} \, S^{--} + ({\overline \alpha}^-\rho \, \delta_{\rm even} -
\alpha^+\rho \, \delta_{\rm odd}) S^- - {\overline \alpha}\alpha^+ \notag
\\
& \quad + ({\overline \alpha} \rho^+ \, \delta_{\rm even} - \alpha^{++} \rho^+ \,
\delta_{\rm odd}) S^+ + \rho^+ \rho^{++} \, \delta_{\rm odd} \, S^{++},\label{2.9}
\end{align}
where $\delta_{\rm even}$ and $\delta_{\rm odd}$ denote the characteristic functions of the
even and odd integers,
\begin{equation}\label{2.10}
\delta_{\rm even} = \chi_{_{2{\mathbb{Z}}}}, \quad \delta_{\rm odd} = 1 - \delta_{\rm even} = \chi_{_{2{\mathbb{Z}} +1}}
\end{equation}
and $S^\pm$, $S^{++}$, $S^{--}$ denote the shift operators acting
upon $\s{{\mathbb{Z}}}$, that is, $S^{\pm}f(\cdot)=f^{\pm}(\cdot)=f
(\cdot\pm1)$ for $f\in \s{{\mathbb{Z}}}$, $S^{++}=S^+S^+$, and $S^{--}=S^-S^-$.
Here the diagonal entries in the infinite matrix \eqref{2.8} are
given by $U_{k,k}=-\overline{\alpha_k}\alpha_{k+1}$, $k\in{\mathbb{Z}}$.
As explained in the introduction, in the recent literature on
orthogonal polynomials on the unit circle, such operators $U$ are
frequently called CMV operators.
Next we recall some of the principal results of \cite{GZ06} needed in
this paper.
\begin{lemma} \label{l2.2}
Let $z\in\mathbb{C}\backslash\{0\}$ and suppose $\{u(z,k)\}_{k\in{\mathbb{Z}}}$,
$\{v(z,k)\}_{k\in{\mathbb{Z}}}\in\s{{\mathbb{Z}}}$. Then the following items
$(i)$--$(iii)$ are equivalent:
\begin{align}
(i) &\quad (U u(z,\cdot))(k) = z u(z,k), \quad (W u(z,\cdot))(k)=z
v(z,k), \quad k\in{\mathbb{Z}}.
\\
(ii) &\quad (W u(z,\cdot))(k)=z v(z,k), \quad (V v(z,\cdot))(k) =
u(z,k), \quad k\in{\mathbb{Z}}.
\\
(iii) &\quad \binom{u(z,k)}{v(z,k)} = T(z,k)
\binom{u(z,k-1)}{v(z,k-1)}, \quad k\in{\mathbb{Z}}, \label{2.11}
\end{align}
where the transfer matrices $T(z,k)$, $z\in{\C\backslash\{0\}}$, $k\in{\mathbb{Z}}$, are given
by
\begin{equation}
T(z,k) = \begin{cases} \frac{1}{\rho_{k}} \begin{pmatrix} \alpha_{k} & z
\\ 1/z & \overline{\alpha_{k}} \end{pmatrix}, & \text{$k$
odd,} \\[20pt] \frac{1}{\rho_{k}} \begin{pmatrix} \overline{\alpha_{k}} & 1 \\
1 & \alpha_{k} \end{pmatrix}, & \text{$k$ even.} \end{cases} \label{2.12}
\end{equation}
Here $U$, $V$, and $W$ are understood in the sense of difference
expressions on $\s{{\mathbb{Z}}}$ rather than difference operators on
$\lt{{\mathbb{Z}}}$.
\end{lemma}
If one sets $\alpha_{k_0} = e^{it}$, $t\in [0,2\pi)$, for some reference
point $k_0\in{\mathbb{Z}}$, then the CMV operator (denoted in this case by
$U^{(t)}_{k_0}$) splits into a direct sum of two half-lattice
operators $U_{-,k_0-1}^{(t)}$ and $U_{+,k_0}^{(t)}$ acting on
$\lt{(-\infty,k_0-1]\cap{\mathbb{Z}}}$ and on $\lt{[k_0,\infty)\cap{\mathbb{Z}}}$,
respectively. Explicitly, one obtains
\begin{align}
\begin{split}
& U^{(t)}_{k_0}=U^{(t)}_{-,k_0-1} \oplus U^{(t)}_{+,k_0} \, \text{ on
} \, \lt{(-\infty,k_0-1]\cap{\mathbb{Z}}} \oplus \lt{[k_0,\infty)\cap{\mathbb{Z}}}
\\ & \text{if } \, \alpha_{k_0} = e^{it}, \; t\in [0,2\pi).
\label{2.13}
\end{split}
\end{align}
(Strictly, speaking, setting $\alpha_{k_0} = e^{it}$, $t\in[0,2\pi)$,
for some reference point $k_0\in{\mathbb{Z}}$ contradicts our basic Hypothesis
\ref{h2.1}. However, as long as the exception to Hypothesis
\ref{h2.1} refers to only one site, we will safely ignore this
inconsistency in favor of the notational simplicity it provides by
avoiding the introduction of a properly modified hypothesis on
$\{\alpha_k\}_{k\in{\mathbb{Z}}}$.) Similarly, one obtains $V^{(t)}_{k_0}$,
$W^{(t)}_{k_0}$, $V^{(t)}_{\pm,k_0}$, and $W^{(t)}_{\pm,k_0}$, so
that
\begin{align}
& V^{(t)}_{k_0}=V^{(t)}_{-,k_0-1} \oplus V^{(t)}_{+,k_0}, &&
W^{(t)}_{k_0}=W^{(t)}_{-,k_0-1} \oplus W^{(t)}_{+,k_0}, \notag
\\
& U^{(t)}_{k_0}=V^{(t)}_{k_0} W^{(t)}_{k_0}, && U^{(t)}_{\pm,k_0} =
V^{(t)}_{\pm,k_0} W^{(t)}_{\pm,k_0}. \label{2.14}
\end{align}
For simplicity we will abbreviate
\begin{equation}
U_{\pm,k_0} =
U_{\pm,k_0}^{(t=0)}=V_{\pm,k_0}^{(t=0)}W_{\pm,k_0}^{(t=0)}
=V_{\pm,k_0} W_{\pm,k_0}. \label{2.15}
\end{equation}
It is instructive to introduce one more sequence of Verblunsky
coefficients $\beta=\{\beta_k\}_{k\in{\mathbb{Z}}}\in\s{{\mathbb{Z}}}$ by
\begin{align}
\beta_k=e^{-it}\alpha_k, \quad k\in{\mathbb{Z}}, \label{2.17}
\end{align}
so that the sequence $\{\rho_k\}_{k\in{\mathbb{Z}}}$ is unchanged and
$\alpha_{k_0}=e^{it}$ corresponds to $\beta_{k_0}=1$. Then the CMV
operators $U_\beta$ and $U_{\pm,k_0;\beta}$ associated with $\beta$ are
unitarily equivalent to the corresponding CMV operators $U_\alpha$ and
$U^{(t)}_{\pm,k_0;\alpha}$ associated with $\alpha$. Indeed, one verifies
that
\begin{align}
\begin{pmatrix}e^{-it/2}&0\\0&e^{it/2}\end{pmatrix}
\begin{pmatrix} -\alpha_k & \rho_k \\ \rho_k & \overline{\alpha_k}\end{pmatrix}
\begin{pmatrix}e^{-it/2}&0\\0&e^{it/2}\end{pmatrix} =
\begin{pmatrix} -\beta_k & \rho_k \\ \rho_k & \overline{\beta_k}\end{pmatrix},
\quad k\in{\mathbb{Z}}, \label{2.17a}
\end{align}
and hence, setting $A$ to be the following diagonal unitary operator
on $\lt{{\mathbb{Z}}}$,
\begin{align}
A = e^{-it/2}\delta_{\rm odd} + e^{it/2}\delta_{\rm even},
\end{align}
one obtains for the full-lattice and the direct sum of half-lattice
CMV operators,
\begin{align}
& AU_{\alpha}A^* = [AV_{\alpha}A] [A^*W_{\alpha}A^*] = V_{\beta}W_{\beta} =
U_{\beta}, \label{2.17b}
\\
& AU^{(t)}_{k_0;\alpha}A^* = \big[AV^{(t)}_{k_0;\alpha}A\big]
\big[A^*W^{(t)}_{k_0;\alpha}A^*\big] = V_{k_0;\beta}W_{k_0;\beta} = U_{k_0;\beta}.
\label{2.17c}
\end{align}
We refer to \cite[Sect. 3]{GZ06a} for additional results on CMV
operators with Verblunsky coefficients related via \eqref{2.17}.
Now we turn to our principal result of this section.
\begin{theorem} \label{t2.3}
Fix $t_1, t_2\in[0,2\pi)$, $k_0\in{\mathbb{Z}}$, $z\in{\C\backslash{\partial\hspace*{.2mm}\mathbb{D}}}$, and let
$U^{(t_1,t_2)}_{k_0}$ denote the following unitary operator on
$\lt{{\mathbb{Z}}}$,
\begin{align}
U^{(t_1,t_2)}_{k_0} = U^{(t_1)}_{-,k_0-1} \oplus U^{(t_2)}_{+,k_0}.
\label{2.18}
\end{align}
Then $U-U^{(t_1,t_2)}_{k_0}$ and
$(U-zI)^{-1}-\big(U^{(t_1,t_2)}_{k_0}-zI\big)^{-1}$ are of rank one if and
only if the relation $t_1=2\arg\big[i(\alpha_{k_0}e^{-it_2/2}-e^{it_2/2})\big]$
holds. Otherwise, these differences are of rank two. In particular,
$U-U^{(t)}_{k_0}$ and $(U-zI)^{-1}-\big(U^{(t)}_{k_0}-zI\big)^{-1}$ are of
rank two for any $t\in[0,2\pi)$.
\end{theorem}
\begin{proof}
Similar to \eqref{2.18} we introduce unitary operators
$V^{(t_1,t_2)}_{k_0}$ and $W^{(t_1,t_2)}_{k_0}$ by
\begin{align}
V^{(t_1,t_2)}_{k_0} = V^{(t_1)}_{-,k_0-1} \oplus V^{(t_2)}_{+,k_0}
\,\text{ and }\, W^{(t_1,t_2)}_{k_0} = W^{(t_1)}_{-,k_0-1} \oplus
W^{(t_2)}_{+,k_0} \,\text{ on }\, \lt{{\mathbb{Z}}}. \label{2.19}
\end{align}
Then $U^{(t_1,t_2)}_{k_0}=V^{(t_1,t_2)}_{k_0}W^{(t_1,t_2)}_{k_0}$ by
\eqref{2.14}, and hence, it follows from \eqref{2.5} that
\begin{align}
U-U^{(t_1,t_2)}_{k_0} =
\begin{cases}
V\big(W-W^{(t_1,t_2)}_{k_0}\big), & \text{$k_0$ odd,} \\
\big(V-V^{(t_1,t_2)}_{k_0}\big)W, & \text{$k_0$ even.}
\end{cases} \label{2.20}
\end{align}
For $k_0$ odd, $D=W-W^{(t_1,t_2)}_{k_0}$ is block-diagonal with all
its $2\times 2$ blocks on the diagonal being zero except for one
which has the following form
\begin{align}
\begin{pmatrix}
D_{k_0-1,k_0-1} & D_{k_0-1,k_0} \\ D_{k_0,k_0-1} & D_{k_0,k_0}
\end{pmatrix}
=
\begin{pmatrix}
-\alpha_{k_0} & \rho_0 \\ \rho_{k_0} & \overline{\alpha_{k_0}}
\end{pmatrix}
-
\begin{pmatrix}
-e^{it_1} & 0 \\ 0 & e^{-it_2}
\end{pmatrix}. \label{2.21}
\end{align}
Thus, the difference $U-U^{(t_1,t_2)}_{k_0}$ in \eqref{2.20} is
always of rank one or two and it is precisely of rank one if and only
if the $2\times 2$ matrix in \eqref{2.21} is of rank 1. The latter
case is equivalent to
\begin{align}
0&=\det\begin{pmatrix} D_{k_0-1,k_0-1} & D_{k_0-1,k_0} \\
D_{k_0,k_0-1} & D_{k_0,k_0}
\end{pmatrix}
= e^{it_1}\overline{\alpha_{k_0}} + e^{-it_2}\alpha_{k_0} - e^{i(t_1-t_2)} - 1
\notag
\\
&= (\overline{\alpha_{k_0}}-e^{-it_2})\left(e^{it_1} + e^{-it_2}
\frac{\alpha_{k_0}-e^{it_2}}{\overline{\alpha_{k_0}}-e^{-it_2}}\right) \notag
\\
&= (\overline{\alpha_{k_0}}-e^{-it_2})\left(e^{it_1} -
\frac{i\alpha_{k_0}e^{-it_2/2}-e^{it_2/2}}
{\overline{i\alpha_{k_0}e^{-it_2/2}-e^{it_2/2}}}\right),\label{2.22a}
\end{align}
which holds if and only if
$t_1=2\arg\big[i(\alpha_{k_0}e^{-it_2/2}-e^{it_2/2})\big]$. The case of
even $k_0$ follows similarly.
Finally, the statement for the resolvents follows from the result
for $U-U^{(t_1,t_2)}_{k_0}$ and the following identity,
\begin{align}
(U-zI)^{-1}-(U^{(t_1,t_2)}_{k_0}-zI)^{-1} = -(U-zI)^{-1}
\Big[U-U^{(t_1,t_2)}_{k_0}\Big] (U^{(t_1,t_2)}_{k_0}-zI)^{-1}.
\end{align}
\end{proof}
Next we present formulas that link various spectral theoretic objects
associated with half-lattice CMV operators $U^{(t)}_{\pm,k_0}$ for
different values of $t\in[0,2\pi)$. We start with an analog of Lemma
\ref{l2.2} for difference expressions $U^{(t)}_{\pm,k_0}$,
$V^{(t)}_{\pm,k_0}$, and $W^{(t)}_{\pm,k_0}$. In the special case
$t=0$ it is proven in \cite[Lem. 2.3]{GZ06} and the general case
below follows immediately from the special case and the observation
of unitary equivalence in \eqref{2.17c}.
\begin{lemma} \label{l2.3}
Fix $t\in[0,2\pi)$, $k_0\in{\mathbb{Z}}$, $z\in\mathbb{C}\backslash\{0\}$, and let
$\big\{\hat p^{(t)}_+(z,k,k_0)\big\}_{k\geq k_0}$,
$\big\{\hat r^{(t)}_+(z,k,k_0)\big\}_{k\geq k_0}\in\s{[k_0,\infty)\cap{\mathbb{Z}}}$. Then the
following items $(i)$--$(iii)$ are equivalent:
\begin{align}
(i) &\quad \big(U^{(t)}_{+,k_0} \hat p^{(t)}_+(z,\cdot,k_0)\big)(k) = z \hat
p^{(t)}_+(z,k,k_0), \notag \\
&\quad \big(W^{(t)}_{+,k_0} \hat p^{(t)}_+(z,\cdot,k_0)\big)(k) = z \hat
r^{(t)}_+(z,k,k_0), \quad k\geq k_0.
\\
(ii) &\quad \big(W^{(t)}_{+,k_0} \hat p^{(t)}_+(z,\cdot,k_0)\big)(k) = z
\hat r^{(t)}_+(z,k,k_0), \notag \\
&\quad \big(V^{(t)}_{+,k_0} \hat r^{(t)}_+(z,\cdot,k_0)\big)(k) = \hat
p^{(t)}_+(z,k,k_0), \quad k\geq k_0.
\\
(iii) &\quad \binom{\hat p^{(t)}_+(z,k,k_0)}{\hat r^{(t)}_+(z,k,k_0)}
= T(z,k) \binom{\hat p^{(t)}_+(z,k-1,k_0)}{\hat
r^{(t)}_+(z,k-1,k_0)}, \quad k > k_0, \notag \\
&\quad \hat p^{(t)}_+(z,k_0,k_0) =
\begin{cases}
ze^{it} \hat r^{(t)}_+(z,k_0,k_0), & \text{$k_0$ odd}, \\
e^{-it} \hat r^{(t)}_+(z,k_0,k_0), & \text{$k_0$ even}.
\end{cases}
\end{align}
Similarly, let $\big\{\hat p^{(t)}_-(z,k,k_0)\big\}_{k\leq k_0}$,
$\big\{\hat r^{(t)}_-(z,k,k_0)\big\}_{k\leq k_0}\in\s{(-\infty,k_0]\cap{\mathbb{Z}}}$. Then the
following items $(iv)$--$(vi)$are equivalent:
\begin{align}
(iv) &\quad \big(U^{(t)}_{-,k_0} \hat p^{(t)}_-(z,\cdot,k_0)\big)(k) = z
\hat p^{(t)}_-(z,k,k_0), \notag \\
&\quad \big(W^{(t)}_{-,k_0} \hat p^{(t)}_-(z,\cdot,k_0)\big)(k) = z \hat
r^{(t)}_-(z,k,k_0), \quad k\leq k_0.
\\
(v) &\quad \big(W^{(t)}_{-,k_0} \hat p^{(t)}_-(z,\cdot,k_0)\big)(k) = z
\hat r^{(t)}_-(z,k,k_0), \notag \\
&\quad \big(V^{(t)}_{-,k_0} \hat r^{(t)}_-(z,\cdot,k_0)\big)(k) = \hat
p^{(t)}_-(z,k,k_0), \quad k\leq k_0.
\\
(vi) &\quad \binom{\hat p^{(t)}_-(z,k-1,k_0)}{\hat
r^{(t)}_-(z,k-1,k_0)}=T(z,k)^{-1} \binom{\hat
p^{(t)}_-(z,k,k_0)}{\hat r^{(t)}_-(z,k,k_0)}, \quad k \leq k_0,\notag
\\
&\quad \hat p^{(t)}_-(z,k_0,k_0) =
\begin{cases}
-e^{it}\hat r^{(t)}_-(z,k_0,k_0), & \text{$k_0$ odd,}
\\
-ze^{-it} \hat r^{(t)}_-(z,k_0,k_0), & \text{$k_0$ even.}
\end{cases}
\end{align}
\end{lemma}
In the following, we denote by
$\Big(\begin{smallmatrix}p^{(t)}_\pm(z,k,k_0)\\
r^{(t)}_\pm(z,k,k_0)\end{smallmatrix}\Big)_{k\in{\mathbb{Z}}}$ and
$\Big(\begin{smallmatrix}q^{(t)}_\pm(z,k,k_0)\\
s^{(t)}_\pm(z,k,k_0)\end{smallmatrix}\Big)_{k\in{\mathbb{Z}}}$,
$z\in\mathbb{C}\backslash\{0\}$, four linearly independent solutions of
\eqref{2.11} with the initial conditions:
\begin{align}
\binom{p^{(t)}_+(z,k_0,k_0)}{r^{(t)}_+(z,k_0,k_0)} = \begin{cases}
\binom{z\overline{\gamma}}{\gamma}, &
\text{$k_0$ odd,} \\[1mm]
\binom{\gamma}{\overline{\gamma}}, & \text{$k_0$ even,} \end{cases} \quad
\binom{q^{(t)}_+(z,k_0,k_0)}{s^{(t)}_+(z,k_0,k_0)} = \begin{cases}
\binom{z\overline{\gamma}}{-\gamma}, &
\text{$k_0$ odd,} \\[1mm]
\binom{-\gamma}{\overline{\gamma}}, & \text{$k_0$ even.} \end{cases} \label{2.22}
\\
\binom{p^{(t)}_-(z,k_0,k_0)}{r^{(t)}_-(z,k_0,k_0)} = \begin{cases}
\binom{\overline{\gamma}}{-\gamma}, &
\text{$k_0$ odd,} \\[1mm]
\binom{-z\gamma}{\overline{\gamma}}, & \text{$k_0$ even,} \end{cases} \quad
\binom{q^{(t)}_-(z,k_0,k_0)}{s^{(t)}_-(z,k_0,k_0)} =
\begin{cases} \binom{\overline{\gamma}}{\gamma}, &
\text{$k_0$ odd,} \\[1mm]
\binom{z\gamma}{\overline{\gamma}}, & \text{$k_0$ even,} \end{cases} \label{2.23}
\end{align}
where $\gamma=e^{-it/2}$. Then it follows that $p^{(t)}_\pm(z,k,k_0)$,
$q^{(t)}_\pm(z,k,k_0)$, $r^{(t)}_\pm(z,k,k_0)$, and
$s^{(t)}_\pm(z,k,k_0)$, $k,k_0\in{\mathbb{Z}}$, are Laurent polynomials in $z$,
that is, finite linear combinations of terms $z^k$, k$\in{\mathbb{Z}}$, with
complex-valued coefficients.
Since all of the above sequences satisfy the same recursion relation
\eqref{2.11} which can have at most two linearly independent
solutions, these sequences satisfy various identities. Some of them
we state in the following lemma.
\begin{lemma} \label{l2.4}
Let $t_1,t_2\in[0,2\pi)$ and $\gamma_j=e^{-it_j/2}$, $j=1,2$. Then
\begin{align}
&\binom{q^{(t_2)}_\pm(z,\cdot,k_0)}{s^{(t_2)}_\pm(z,\cdot,k_0)} =
\Re(\gamma_1\overline{\gamma_2})
\binom{q^{(t_1)}_\pm(z,\cdot,k_0)}{s^{(t_1)}_\pm(z,\cdot,k_0)} +
i\Im(\gamma_1\overline{\gamma_2})
\binom{p^{(t_1)}_\pm(z,\cdot,k_0)}{r^{(t_1)}_\pm(z,\cdot,k_0)},
\label{2.24}
\\
&\binom{p^{(t_2)}_\pm(z,\cdot,k_0)}{r^{(t_2)}_\pm(z,\cdot,k_0)} =
i\Im(\gamma_1\overline{\gamma_2})
\binom{q^{(t_1)}_\pm(z,\cdot,k_0)}{s^{(t_1)}_\pm(z,\cdot,k_0)} +
\Re(\gamma_1\overline{\gamma_2})
\binom{p^{(t_1)}_\pm(z,\cdot,k_0)}{r^{(t_1)}_\pm(z,\cdot,k_0)},
\label{2.25}
\\
&\binom{q^{(t_2)}_-(z,\cdot,k_0)}{s^{(t_2)}_-(z,\cdot,k_0)} =
\frac{\gamma_1\overline{\gamma_2}-\overline{\gamma_1}\gamma_2z}{2z^{k_0 \, ({\rm mod}\,2)}}
\binom{q^{(t_1)}_+(z,\cdot,k_0)}{s^{(t_1)}_+(z,\cdot,k_0)} +
\frac{\gamma_1\overline{\gamma_2}+\overline{\gamma_1}\gamma_2z}{2z^{k_0 \, ({\rm mod}\,2)}}
\binom{p^{(t_1)}_+(z,\cdot,k_0)}{r^{(t_1)}_+(z,\cdot,k_0)},
\label{2.26}
\\
&\binom{p^{(t_2)}_-(z,\cdot,k_0)}{r^{(t_2)}_-(z,\cdot,k_0)} =
\frac{\gamma_1\overline{\gamma_2}+\overline{\gamma_1}\gamma_2z}{2z^{k_0 \, ({\rm mod}\,2)}}
\binom{q^{(t_1)}_+(z,\cdot,k_0)}{s^{(t_1)}_+(z,\cdot,k_0)} +
\frac{\gamma_1\overline{\gamma_2}-\overline{\gamma_1}\gamma_2z}{2z^{k_0 \, ({\rm mod}\,2)}}
\binom{p^{(t_1)}_+(z,\cdot,k_0)}{r^{(t_1)}_+(z,\cdot,k_0)},
\label{2.26a}
\\
&\binom{q^{(t_2)}_-(z,\cdot,k_0-1)}{s^{(t_2)}_-(z,\cdot,k_0-1)} =
\frac{i\Im(\gamma_1\overline{\gamma_2}+\alpha_{k_0}\gamma_1\gamma_2)}{\rho_{k_0}}
\binom{q^{(t_1)}_+(z,\cdot,k_0)}{s^{(t_1)}_+(z,\cdot,k_0)} \notag
\\ &\hspace{35mm}
+ \frac{\Re(\gamma_1\overline{\gamma_2}+\alpha_{k_0}\gamma_1\gamma_2)}{\rho_{k_0}}
\binom{p^{(t_1)}_+(z,\cdot,k_0)}{r^{(t_1)}_+(z,\cdot,k_0)},
\label{2.27}
\\
&\binom{p^{(t_2)}_-(z,\cdot,k_0-1)}{r^{(t_2)}_-(z,\cdot,k_0-1)} =
\frac{\Re(\gamma_1\overline{\gamma_2}-\alpha_{k_0}\gamma_1\gamma_2)}{\rho_{k_0}}
\binom{q^{(t_1)}_+(z,\cdot,k_0)}{s^{(t_1)}_+(z,\cdot,k_0)} \notag
\\ &\hspace{35mm}
+ \frac{i\Im(\gamma_1\overline{\gamma_2}-\alpha_{k_0}\gamma_1\gamma_2)}{\rho_{k_0}}
\binom{p^{(t_1)}_+(z,\cdot,k_0)}{r^{(t_1)}_+(z,\cdot,k_0)}.
\label{2.27a}
\end{align}
In particular, whenever
$t_1=2\arg\big[i(\alpha_{k_0}e^{-it_2/2}-e^{it_2/2})\big]$ identity
\eqref{2.27a} simplifies to
\begin{align}
\binom{p^{(t_2)}_-(z,\cdot,k_0-1)}{r^{(t_2)}_-(z,\cdot,k_0-1)} &=
\frac{i\abs{e^{it_2}-\alpha_{k_0}}}{\rho_{k_0}}
\binom{p^{(t_1)}_+(z,\cdot,k_0)}{r^{(t_1)}_+(z,\cdot,k_0)}. \label{2.28}
\end{align}
\end{lemma}
\begin{proof}
Since both sides of \eqref{2.24}--\eqref{2.27a} satisfy the same
recursion relation \eqref{2.11} it suffices to check these equalities
only at one point, say at point $k=k_0$. Substituting \eqref{2.22}
and \eqref{2.23} into \eqref{2.24}--\eqref{2.26a} one verifies the
first four identities. Using \eqref{2.22} and \eqref{2.23} once again
and applying transfer matrix $T(z,k_0)$ to the left hand-sides of
\eqref{2.27}--\eqref{2.28} one verifies the last three identities.
\end{proof}
Next, following \cite{GZ06} we introduce half-lattice
Weyl--Titchmarsh $m$-functions associated with the CMV operators
$U^{(t)}_{\pm,k_0}$ by
\begin{align}
\begin{split}
m^{(t)}_\pm(z,k_0) &= \pm
(\delta_{k_0},(U^{(t)}_{\pm,k_0}+zI)(U^{(t)}_{\pm,k_0}-zI)^{-1}
\delta_{k_0})_{\lt{{\mathbb{Z}}\cap[k_0,\pm\infty)}} \\
& =\pm \oint_{\partial\hspace*{.2mm}\mathbb{D}} d\mu^{(t)}_{\pm}(\zeta,k_0)\,
\frac{\zeta+z}{\zeta-z}, \quad z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}},
\end{split} \label{2.30}
\end{align}
where
\begin{equation}
d\mu^{(t)}_\pm(\zeta,k_0) = d(\delta_{k_0},E_{U^{(t)}_{\pm,k_0}}(\zeta)
\delta_{k_0})_{\lt{{\mathbb{Z}}\cap[k_0,\pm\infty)}}, \quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}}, \label{2.31}
\end{equation}
and $dE_{U^{(t)}_{\pm,k_0}}(\cdot)$ denote the operator-valued
spectral measures of the operators $U^{(t)}_{\pm,k_0}$,
\begin{equation}
U^{(t)}_{\pm,k_0}=\oint_{{\partial\hspace*{.2mm}\mathbb{D}}} dE_{U^{(t)}_{\pm,k_0}}(\zeta)\,\zeta.
\end{equation}
Then following the steps of \cite[Cor. 2.14]{GZ06} one verifies that
\begin{equation}
\binom{q^{(t)}_\pm(z,\cdot,k_0)}{s^{(t)}_\pm(z,\cdot,k_0)} +
m^{(t)}_\pm(z,k_0)
\binom{p^{(t)}_\pm(z,\cdot,k_0)}{r^{(t)}_\pm(z,\cdot,k_0)} \in
\lt{[k_0,\pm\infty)\cap{\mathbb{Z}}}^2,
\;\; z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\}). \label{2.32}
\end{equation}
The special case $t=0$ of the next result is proven in \cite[Cor.
2.16 and Thm. 2.18]{GZ06}. The general case of $t\in[0,2\pi)$ stated
below follows along the same lines and hence we omit the details for
brevity.
\begin{theorem} \label{t2.5}
Let $t\in[0,2\pi)$ and $k_0\in{\mathbb{Z}}$. Then there exist unique
Caratheodory $($resp. anti-Caratheodory$)$ functions
$M^{(t)}_\pm(\cdot,k_0)$ such that
\begin{align}
&\binom{u^{(t)}_\pm(z,\cdot,k_0)}{v^{(t)}_\pm(z,\cdot,k_0)} =
\binom{q^{(t)}_+(z,\cdot,k_0)}{s^{(t)}_+(z,\cdot,k_0)} +
M^{(t)}_\pm(z,k_0)
\binom{p^{(t)}_+(z,\cdot,k_0)}{r^{(t)}_+(z,\cdot,k_0)} \in
\lt{[k_0,\pm\infty)\cap{\mathbb{Z}}}^2, \notag \\
& \hspace*{9.5cm} z\in\mathbb{C}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\}). \label{2.37}
\end{align}
In addition, sequence
$\Big(\begin{smallmatrix}u^{(t)}_\pm(z,k,k_0)\\
v^{(t)}_\pm(z,k,k_0)\end{smallmatrix}\Big)_{k\in{\mathbb{Z}}}$ satisfies
\eqref{2.11} and is unique $($up to constant scalar multiples$)$
among all sequence that satisfy \eqref{2.11} and are square summable
near $\pm\infty$.
\end{theorem}
We will call $u^{(t)}_\pm(z,\cdot,k_0)$ and
$v^{(t)}_\pm(z,\cdot,k_0)$ Weyl--Titchmarsh solutions of $U$.
Similarly, we will call $m^{(t)}_\pm(z,k_0)$ as well as
$M^{(t)}_\pm(z,k_0)$ the half-lattice Weyl--Titchmarsh
$m$-functions associated with $U^{(t)}_{\pm,k_0}$. (See also
\cite{Si04a} for a comparison of various alternative notions of
Weyl--Titchmarsh $m$-functions for $U_{+,k_0}$.)
Using \eqref{2.26}--\eqref{2.27a}, \eqref{2.32}, and Theorem
\ref{t2.5} one also verifies that
\begin{align}
M^{(t)}_+(z,k_0) &= m^{(t)}_+(z,k_0), \quad z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}},
\label{2.38}
\\
M^{(t)}_+(0,k_0) &=1, \label{2.39}
\\
M^{(t)}_-(z,k_0) &= \frac{\Re(a_{k_0}) +
i\Im(b_{k_0})m^{(t)}_-(z,k_0-1)}{i\Im(a_{k_0}) +
\Re(b_{k_0})m^{(t)}_-(z,k_0-1)} =
\frac{(1-z)m^{(t)}_-(z,k_0)+(1+z)}{(1+z)m^{(t)}_-(z,k_0)+(1-z)} \notag
\\
&= \frac{\big(m^{(t)}_-(z,k_0)+1\big)-z\big(m^{(t)}_-(z,k_0)-1\big)}
{\big(m^{(t)}_-(z,k_0)+1\big)+z\big(m^{(t)}_-(z,k_0)-1\big)}, \quad
z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}}, \label{2.40}
\\
M^{(t)}_-(0,k_0) &=\frac{\alpha_{k_0}+e^{it}}{\alpha_{k_0}-e^{it}},
\label{2.41}
\\
m^{(t)}_-(z,k_0) &=
\frac{\Re(a_{k_0+1})-i\Im(a_{k_0+1})M^{(t)}_-(z,k_0+1)}
{\Re(b_{k_0+1})M^{(t)}_-(z,k_0+1) - i\Im(b_{k_0+1})} \notag
\\
&= \frac{z\big(M^{(t)}_-(z,k_0)+1\big)-\big(M^{(t)}_-(z,k_0)-1\big)}
{z\big(M^{(t)}_-(z,k_0)+1\big)+\big(M^{(t)}_-(z,k_0)-1\big)}, \quad z\in{\C\backslash{\partial\hspace*{.2mm}\mathbb{D}}},
\label{2.42}
\end{align}
where $a_k=1+e^{-it}\alpha_k$ and $b_k=1-e^{-it}\alpha_k$, $k\in{\mathbb{Z}}$. In
particular, one infers that $M^{(t)}_\pm$ are analytic at $z=0$.
Next, we introduce the Schur (resp. anti-Schur) functions
$\Phi^{(t)}_\pm(\cdot,k)$, $k\in{\mathbb{Z}}$, by
\begin{align}
\Phi^{(t)}_\pm(z,k) = \frac{M^{(t)}_\pm(z,k)-1}{M^{(t)}_\pm(z,k)+1},
\quad z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}}. \label{2.44}
\end{align}
Then by \eqref{2.42} and \eqref{2.44},
\begin{align}
&M^{(t)}_\pm(z,k) = \frac{1+\Phi^{(t)}_\pm(z,k)}{1-\Phi^{(t)}_\pm(z,k)},
\quad m^{(t)}_-(z,k) = \frac{z-\Phi^{(t)}_-(z,k)}{z+\Phi^{(t)}_-(z,k)},
\quad z\in{\C\backslash{\partial\hspace*{.2mm}\mathbb{D}}}. \label{2.45}
\end{align}
Moreover, it follows from \eqref{2.22}, \eqref{2.44}, and Theorem
\ref{t2.5} that
\begin{align}
&\Phi^{(t)}_\pm(z,k) = \begin{cases}
ze^{it}\frac{v^{(t)}_\pm(z,k,k_0)}{u^{(t)}_\pm(z,k,k_0)}, &\text{$k$
odd,}
\\
e^{it}\frac{u^{(t)}_\pm(z,k,k_0)}{v^{(t)}_\pm(z,k,k_0)}, & \text{$k$
even,}
\end{cases} \quad k\in{\mathbb{Z}}, \; z\in{\C\backslash{\partial\hspace*{.2mm}\mathbb{D}}}, \label{2.46}
\end{align}
where $u^{(t)}_\pm(\cdot,k,k_0)$ and $v^{(t)}_\pm(\cdot,k,k_0)$ are
the sequences defined in \eqref{2.37}. Since the Weyl--Titchmarsh
solution $\Big(\begin{smallmatrix}u^{(t)}_\pm(z,k,k_0)\\
v^{(t)}_\pm(z,k,k_0)\end{smallmatrix}\Big)_{k\in{\mathbb{Z}}}$ is unique up to
a multiplicative constant, we conclude from \eqref{2.46} that
$e^{-it}\Phi^{(t)}_\pm(\cdot,k)$ is actually $t$-independent. Thus,
fixing $t_1, t_2\in[0,2\pi)$, one computes
\begin{align}
\Phi^{(t_2)}_\pm(\cdot,k) &= e^{i(t_2-t_1)}\Phi^{(t_1)}_\pm(\cdot,k),
\\
M^{(t_2)}_\pm(\cdot,k) &=
\frac{i\Im(e^{i(t_2-t_1)/2})+\Re(e^{i(t_2-t_1)/2})M^{(t_1)}_\pm(\cdot,k)}
{\Re(e^{i(t_2-t_1)/2})+i\Im(e^{i(t_2-t_1)/2})M^{(t_1)}_\pm(\cdot,k)},
\quad k\in{\mathbb{Z}}.
\end{align}
Finally, following \cite{GZ06}, we obtain the following identities,
\begin{align}
r^{(t)}_+(z,k,k_0) &= z^{k_0 \, ({\rm mod}\,2)}\overline{p^{(t)}_+(1/\overline{z},k,k_0)},
\\
s^{(t)}_+(z,k,k_0) &= -z^{k_0 \, ({\rm mod}\,2)}\overline{q^{(t)}_+(1/\overline{z},k,k_0)},
\\
r^{(t)}_-(z,k,k_0) &= -z^{(k_0 +1) \, ({\rm mod}\,2)}\overline{p^{(t)}_-(1/\overline{z},k,k_0)},
\\
s^{(t)}_-(z,k,k_0) &= z^{(k_0 +1) \, ({\rm mod}\,2)}\overline{q^{(t)}_-(1/\overline{z},k,k_0)},
\end{align}
and provide formulas for the resolvents of the half-lattice CMV operators
$U^{(t)}_{\pm,k_0}$ and the full-lattice CMV operator $U$. In the
special case $t=0$ these formulas were obtained in (2.63)--(2.66),
(2.171), (2.172), and (3.7) of \cite{GZ06}.
\begin{lemma} \label{l2.6}
Let $t\in[0,2\pi)$, $k_0\in{\mathbb{Z}}$, and $z\in\mathbb{C}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\})$. Then
the resolvent $\big(U^{(t)}_{\pm,k_0}-zI\big)^{-1}$ is given in terms of its
matrix representation in the standard basis of
$\lt{[k_0,\pm\infty)\cap{\mathbb{Z}}}$ by
\begin{align}
& \big(U^{(t)}_{+,k_0}-zI\big)^{-1}(k,k') = \frac{z^{- k_0 \, ({\rm mod}\,2)}}{2z} \notag \\
& \hspace*{1cm} \times
\begin{cases}
p^{(t)}_+(z,k,k_0)\widehat v^{(t)}_+(z,k',k_0), & \text{$k<k'$ and $k=k'$ odd,} \\
\widehat u^{(t)}_+(z,k,k_0)r^{(t)}_+(z,k',k_0), & \text{$k>k'$ and $k=k'$ even}
\end{cases} \label{2.85}
\\
& \quad = \frac{1}{2z}
\begin{cases}
-p^{(t)}_+(z,k,k_0)\overline{\widehat u^{(t)}_+(1/\overline{z},k',k_0)}, & \text{$k<k'$ and $k=k'$ odd,} \\
\widehat u^{(t)}_+(z,k,k_0)\overline{p^{(t)}_+(1/\overline{z},k',k_0)}, & \text{$k>k'$ and $k=k'$ even,}
\end{cases} \label{2.85a}
\\
&\hspace*{8.1cm} k,k'\in [k_0,\infty)\cap{\mathbb{Z}}, \notag
\\
& \big(U^{(t)}_{-,k_0}-zI\big)^{-1}(k,k') = \frac{z^{-(k_0 +1) \, ({\rm mod}\,2)}}{2z} \notag \\
& \hspace*{1cm} \times
\begin{cases}
\widehat u^{(t)}_-(z,k,k_0)r^{(t)}_-(z,k',k_0), & \text{$k<k'$ and $k=k'$ odd,}\\
p^{(t)}_-(z,k,k_0)\widehat v^{(t)}_-(z,k',k_0), & \text{$k>k'$ and $k=k'$ even}
\end{cases} \label{2.86}
\\
& \quad = \frac{1}{2z}
\begin{cases}
-\widehat u^{(t)}_-(z,k,k_0)\overline{p^{(t)}_-(1/\overline{z},k',k_0)}, & \text{$k<k'$ and $k=k'$ odd,}\\
p^{(t)}_-(z,k,k_0)\overline{\widehat u^{(t)}_-(1/\overline{z},k',k_0)}, & \text{$k>k'$ and $k=k'$ even,}
\end{cases} \label{2.86a}
\\
&\hspace*{7.8cm} k,k'\in (-\infty,k_0]\cap{\mathbb{Z}}, \notag
\end{align}
where the sequences $\widehat u^{(t)}_\pm$ and $\widehat v^{(t)}_\pm$ are defined by
\begin{align}
&\binom{\widehat u^{(t)}_\pm(z,\cdot,k_0)}{\widehat v^{(t)}_\pm(z,\cdot,k_0)} =
\binom{q^{(t)}_-(z,\cdot,k_0)}{s^{(t)}_-(z,\cdot,k_0)} + m^{(t)}_\pm(z,k_0)
\binom{p^{(t)}_-(z,\cdot,k_0)}{r^{(t)}_-(z,\cdot,k_0)} \in
\ell^2([k_0,\pm\infty)\cap{\mathbb{Z}})^2, \notag \\
& \hspace*{9.8cm} z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\}).
\end{align}
The corresponding result for the full-lattice resolvent of $U$ then reads
\begin{align}
& (U-zI)^{-1}(k,k') =
\frac{-z^{-k_0 \, ({\rm mod}\,2)}}{2z[M^{(t)}_+(z,k_0)-M^{(t)}_-(z,k_0)]} \notag \\
& \hspace*{1cm} \times
\begin{cases}
u^{(t)}_-(z,k,k_0)v^{(t)}_+(z,k',k_0), & \text{$k<k'$ and $k=k'$ odd,} \\
u^{(t)}_+(z,k,k_0)v^{(t)}_-(z,k',k_0), & \text{$k>k'$ and $k=k'$ even}
\end{cases} \label{2.87}
\\
& \quad = \frac{
\begin{cases}
u^{(t)}_-(z,k,k_0)\overline{u^{(t)}_+(1/\overline{z},k',k_0)}, & \text{$k<k'$ and $k=k'$ odd,} \\
u^{(t)}_+(z,k,k_0)\overline{u^{(t)}_-(1/\overline{z},k',k_0)}, & \text{$k>k'$ and $k=k'$
even,}
\end{cases}}
{2z[M^{(t)}_+(z,k_0)-M^{(t)}_-(z,k_0)]} \quad k,k' \in{\mathbb{Z}}. \label{2.87a}
\end{align}
Moreover, since $\,U^{(t)}_{\pm,k_0}$ and $U$ are unitary and hence zero is in
the resolvent set, \eqref{2.85}--\eqref{2.87a} analytically extend to $z=0$.
\end{lemma}
\section{CMV operators with matrix-valued coefficients} \label{s3}
In the remainder of this paper, ${\mathbb{C}^{m\times m}}$ denotes the space of $m\times m$ matrices with complex-valued
entries endowed with the operator norm $\norm{\cdot}_{{\mathbb{C}^{m\times m}}}$ (we use the standard Euclidean norm in ${\mathbb{C}}^m$). The adjoint of an element $\gamma\in{\mathbb{C}^{m\times m}}$ is denoted by
$\gamma^*$, $I_m$ denotes the identity matrix in ${\mathbb{C}}^m$, and the real and imaginary
parts of $\gamma$ are defined as usual by $\Re(\gamma)=(\gamma+\gamma^*)/2$ and $\Im(\gamma) =(\gamma-\gamma^*)/(2i)$.
\begin{remark} \label{r3.1}
For simplicity of exposition, we find it convenient to use the
following conventions: We denote by $\s{{\mathbb{Z}}}$ the vector space of all
$\mathbb{C}$-valued sequences, and by $\sm{{\mathbb{Z}}}=\s{{\mathbb{Z}}}\otimes\mathbb{C}^m$ the vector
space of all $\mathbb{C}^m$-valued sequences; that is,
\begin{align} \label{3.1}
\phi=\{\phi(k)\}_{k\in{\mathbb{Z}}}=
\begin{pmatrix}
\vdots\\\phi(-1)\\\phi(0)\\\phi(1)\\\vdots
\end{pmatrix}\in\sm{{\mathbb{Z}}}, \quad
\phi(k)=
\begin{pmatrix}
(\phi(k))_1\\(\phi(k))_2\\\vdots\\(\phi(k))_m
\end{pmatrix}\in\mathbb{C}^m, {k\in{\mathbb{Z}}}.
\end{align}
Moreover, we introduce $\smn{{\mathbb{Z}}}=\sm{{\mathbb{Z}}}\otimes\mathbb{C}^n$, $m,n\in{\mathbb{N}}$, that is,
$\Phi=(\phi_1,\dots,\phi_n)\in\smn{{\mathbb{Z}}}$, where $\phi_j\in\sm{{\mathbb{Z}}}$
for all $j=1,\dots,n$.
We also note that $\smn{{\mathbb{Z}}}=\s{{\mathbb{Z}}}\otimes\mathbb{C}^{m\times n}$, $m,n\in{\mathbb{N}}$;
which is to say that the elements of $\smn{{\mathbb{Z}}}$ can be
identified with the $\mathbb{C}^{m\times n}$-valued sequences,
\begin{align} \label{3.2}
\Phi=\{\Phi(k)\}_{k\in{\mathbb{Z}}}=
\begin{pmatrix}
\vdots\\\Phi(-1)\\\Phi(0)\\\Phi(1)\\\vdots
\end{pmatrix},
\Phi(k)=
\begin{pmatrix}
(\Phi(k))_{1,1} & \hdots & (\Phi(k))_{1,n}\\
\vdots & & \vdots \\
(\Phi(k))_{m,1} & \hdots & (\Phi(k))_{m,n}
\end{pmatrix}\in\mathbb{C}^{m\times n}, {k\in{\mathbb{Z}}},
\end{align}
by setting $\Phi=(\phi_1,\dots,\phi_n)$, where
\begin{align} \label{3.3}
\phi_j=
\begin{pmatrix}
\vdots\\\phi_j(-1)\\\phi_j(0)\\\phi_j(1)\\\vdots
\end{pmatrix}\in\sm{{\mathbb{Z}}}, \quad \phi_j(k)=
\begin{pmatrix}
(\Phi(k))_{1,j}\\\vdots\\(\Phi(k))_{m,j}
\end{pmatrix}\in\mathbb{C}^m, \; j=1,\dots,n,\; k\in{\mathbb{Z}}.
\end{align}
For the elements of $\smn{{\mathbb{Z}}}$ we define the right-multiplication by
$n\times n$ matrices with complex-valued entries by
\begin{align} \label{3.4}
\Phi C = (\phi_1,\dots,\phi_n)
\begin{pmatrix}c_{1,1} & \dots & c_{1,n}\\
\vdots&&\vdots\\c_{n,1} & \dots & c_{n,n}\end{pmatrix} =
\left(\sum_{j=1}^n \phi_j c_{j,1},\dots,\sum_{j=1}^n \phi_j
c_{j,n}\right)\in\smn{{\mathbb{Z}}}
\end{align}
for all $\Phi\in\smn{{\mathbb{Z}}}$ and $C\in\mathbb{C}^{n\times n}$. In addition, for
any linear transformation ${\mathbb{A}}: \sm{{\mathbb{Z}}}\to\sm{{\mathbb{Z}}}$, we define
${\mathbb{A}}\Phi$ for all $\Phi=(\phi_1,\dots,\phi_n)\in\smn{{\mathbb{Z}}}$ by
\begin{align} \label{3.5}
{\mathbb{A}}\Phi = ({\mathbb{A}}\phi_1,\dots,{\mathbb{A}}\phi_n)\in\smn{{\mathbb{Z}}}.
\end{align}
Given the above conventions, we note the subspace
containment: $\ltm{{\mathbb{Z}}}=\ell^2({\mathbb{Z}})\otimes\mathbb{C}^m\subset\sm{{\mathbb{Z}}}$ and
$\ltmn{{\mathbb{Z}}}=\ell^2({\mathbb{Z}})\otimes\mathbb{C}^{m\times n}\subset\smn{{\mathbb{Z}}}$. We also
note that $\ltm{{\mathbb{Z}}}$ represents a Hilbert space with scalar
product given by
\begin{align} \label{3.6}
(\phi,\psi)_{\ltm{{\mathbb{Z}}}} = \sum_{k=-\infty}^\infty\sum_{j=1}^m
\overline{(\phi(k))_j}(\psi(k))_j, \quad \phi,\psi\in\ltm{{\mathbb{Z}}}.
\end{align}
Finally, we note that a straightforward modification of the
above definitions also yields the Hilbert space $\ltm{J}$ as well
as the sets $\ltmn{J}$, $\sm{J}$, and $\smn{J}$ for any
$J\subset{\mathbb{Z}}$.
\end{remark}
We start by introducing our basic assumption:
\begin{hypothesis} \label{h3.2}
Let $m\in{\mathbb{N}}$ and assume $\alpha=\{\alpha_k\}_{k\in{\mathbb{Z}}}$ is a sequence of
$m \times m$ matrices with complex entries\footnote{We emphasize that $\alpha_k\in{\mathbb{C}^{m\times m}}$, $k\in{\mathbb{Z}}$, are general (not necessarily normal) matrices.}
and such that
\begin{equation} \label{3.7}
\norm{\alpha_k}_{{\mathbb{C}^{m\times m}}} < 1, \quad k\in{\mathbb{Z}}.
\end{equation}
\end{hypothesis}
Given a sequence $\alpha$ satisfying \eqref{3.7}, we define two sequences
of positive self-adjoint $m\times m$ matrices $\{\rho_k\}_{k\in{\mathbb{Z}}}$
and $\{\widetilde\rho_k\}_{k\in{\mathbb{Z}}}$ by
\begin{align}
\rho_k &= [I_m-\alpha_k^*\alpha_k]^{1/2}, \quad k\in{\mathbb{Z}}, \label{3.8}
\\
\widetilde\rho_k &= [I_m-\alpha_k\alpha_k^*]^{1/2}, \quad k\in{\mathbb{Z}}, \label{3.9}
\end{align}
and two sequences of $m\times m$ matrices with positive real parts,
$\{a_k\}_{k\in{\mathbb{Z}}}\subset {\mathbb{C}}^{m\times m}$ and
$\{b_k\}_{k\in{\mathbb{Z}}}\subset {\mathbb{C}}^{m\times m}$ by
\begin{align}
a_k &= I_m+\alpha_k, \quad k\in{\mathbb{Z}}, \label{3.10}
\\
b_k &= I_m-\alpha_k, \quad k \in {\mathbb{Z}}. \label{3.11}
\end{align}
Then \eqref{3.7} implies that $\rho_k$ and $\widetilde\rho_k$ are
invertible matrices for all $k\in{\mathbb{Z}}$, and using elementary power series
expansions one verifies the following identities for all $k\in{\mathbb{Z}}$,
\begin{align}
&\widetilde\rho_k^{\pm1}\alpha_k = \alpha_k\rho_k^{\pm1}, \quad
\alpha_k^*\widetilde\rho_k^{\pm1} = \rho_k^{\pm1}\alpha_k^*, \label{3.12}
\quad
a_k^*\widetilde\rho_k^{-2}a_k = a_k\rho_k^{-2}a_k^*, \\
&b_k^*\widetilde\rho_k^{-2}b_k = b_k\rho_k^{-2}b_k^*, \quad
a_k^*\widetilde\rho_k^{-2}b_k + a_k\rho_k^{-2}b_k^* =
b_k^*\widetilde\rho_k^{-2}a_k + b_k\rho_k^{-2}a_k^* = 2I_m. \label{3.13}
\end{align}
According to Simon \cite{Si04}, we call $\alpha_k$ the Verblunsky
coefficients in honor of Verblunsky's pioneering work in the theory
of orthogonal polynomials on the unit circle \cite{Ve35},
\cite{Ve36}.
Next, we introduce a sequence of $2\times 2$ block unitary matrices $\Theta_k$
with $m\times m$ matrix coefficients by
\begin{equation} \label{3.14}
\Theta_k = \begin{pmatrix} -\alpha_k & \widetilde\rho_k \\ \rho_k & \alpha_k^*
\end{pmatrix},
\quad k \in {\mathbb{Z}},
\end{equation}
and two unitary operators ${\mathbb{V}}$ and ${\mathbb{W}}$ on $\ltm{{\mathbb{Z}}}$ by their
matrix representations in the standard basis of $\ltm{{\mathbb{Z}}}$ by
\begin{align} \label{3.15}
{\mathbb{V}} &= \begin{pmatrix} \ddots & & &
\raisebox{-3mm}[0mm][0mm]{\hspace*{-5mm}\Huge $0$} \\ & \Theta_{2k-2} &
& \\ & & \Theta_{2k} & & \\ &
\raisebox{0mm}[0mm][0mm]{\hspace*{-10mm}\Huge $0$} & & \ddots
\end{pmatrix}, \quad
{\mathbb{W}} = \begin{pmatrix} \ddots & & &
\raisebox{-3mm}[0mm][0mm]{\hspace*{-5mm}\Huge $0$}
\\ & \Theta_{2k-1} & & \\ & & \Theta_{2k+1} & & \\ &
\raisebox{0mm}[0mm][0mm]{\hspace*{-10mm}\Huge $0$} & & \ddots
\end{pmatrix},
\end{align}
where
\begin{align}
\begin{pmatrix}
{\mathbb{V}}_{2k-1,2k-1} & {\mathbb{V}}_{2k-1,2k} \\ {\mathbb{V}}_{2k,2k-1} & {\mathbb{V}}_{2k,2k}
\end{pmatrix} = \Theta_{2k},
\quad
\begin{pmatrix}
{\mathbb{W}}_{2k,2k} & {\mathbb{W}}_{2k,2k+1} \\ {\mathbb{W}}_{2k+1,2k} & {\mathbb{W}}_{2k+1,2k+1}
\end{pmatrix} = \Theta_{2k+1},
\quad k\in{\mathbb{Z}}. \label{3.16}
\end{align}
Moreover, we introduce the unitary operator ${\mathbb{U}}$ on
$\ltm{{\mathbb{Z}}}$ as the product of the unitary operators ${\mathbb{V}}$ and ${\mathbb{W}}$ by
\begin{equation} \label{3.17}
{\mathbb{U}} = {\mathbb{V}}{\mathbb{W}},
\end{equation}
or in matrix form in the standard basis of $\ltm{{\mathbb{Z}}}$, by
\begin{align}
{\mathbb{U}} &= \begin{pmatrix} \ddots &&\hspace*{-8mm}\ddots
&\hspace*{-10mm}\ddots &\hspace*{-12mm}\ddots &\hspace*{-14mm}\ddots
&&& \raisebox{-3mm}[0mm][0mm]{\hspace*{-6mm}{\Huge $0$}}
\\
&0& -\alpha_{0}\rho_{-1} & -\alpha_{0}\alpha_{-1}^* & -\widetilde\rho_{0}\alpha_{1} &
\widetilde\rho_{0}\widetilde\rho_{1}
\\
&& \rho_{0}\rho_{-1} &\rho_{0}\alpha_{-1}^* & -\alpha_{0}^*\alpha_{1} &
\alpha_{0}^*\widetilde\rho_{1} & 0
\\
&&&0& -\alpha_{2}\rho_{1} & -\alpha_{2}\alpha_{1}^* & -\widetilde\rho_{2}\alpha_{3} &
\widetilde\rho_{2}\widetilde\rho_{3}
\\
&&\raisebox{-4mm}[0mm][0mm]{\hspace*{-6mm}{\Huge $0$}} &&
\rho_{2}\rho_{1} & \rho_{2}\alpha_{1}^* & -\alpha_{2}^*\alpha_{3} &
\alpha_{2}^*\widetilde\rho_{3}&0
\\
&&&&&\hspace*{-14mm}\ddots &\hspace*{-14mm}\ddots
&\hspace*{-14mm}\ddots &\hspace*{-8mm}\ddots &\ddots
\end{pmatrix}. \label{3.18}
\end{align}
Here terms of the form $-\alpha_{2k}\alpha_{2k-1}^*$ and
$-\alpha_{2k}^*\alpha_{2k+1}$, $k\in{\mathbb{Z}}$, represent the diagonal entries
${\mathbb{U}}_{2k-1,2k-1}$ and ${\mathbb{U}}_{2k,2k}$ of the infinite matrix ${\mathbb{U}}$ in
\eqref{3.18}, respectively. Then, with $\delta_{\rm even}$ and $\delta_{\rm odd}$ defined
in \eqref{2.10}, and by analogy with \eqref{2.9}, we see that as an operator acting
upon $\ltm{{\mathbb{Z}}}$, ${\mathbb{U}}$ can be represented
by
\begin{align}
{\mathbb{U}}&= \rho\rho^-\delta_{\rm even} S^{--} + [\rho(\alpha^-)^*\delta_{\rm even} -\alpha^*\rho\ \delta_{\rm odd} ]S^-
- \alpha^*\alpha^+\delta_{\rm even} - \alpha^+\alpha^*\delta_{\rm odd} \notag\\
& \hspace{4mm}+ [\alpha^*\widetilde\rho^{\hspace{2pt}+}\delta_{\rm even} - \widetilde\rho^{\hspace{2pt}+}\alpha^{++}\delta_{\rm odd} ]S^+
+\widetilde\rho^{\hspace{2pt}+}\widetilde\rho^{\hspace{2pt}++}\delta_{\rm odd} S^{++}.
\end{align}
We continue to call the operator ${\mathbb{U}}$ on
$\ltm{{\mathbb{Z}}}$ the CMV operator since \eqref{3.14}--\eqref{3.18} in the
context of the scalar-valued semi-infinite (i.e., half-lattice) case
were obtained by Cantero, Moral, and Vel\'azquez in \cite{CMV03} in
2003. Then, in analogy with Lemma~\ref{l2.2}, the following result is proven in
\cite{CGZ07}:
\begin{lemma} \label{l3.3}
Let $z\in{\mathbb{C}}\backslash\{0\}$ and $\{U(z,k)\}_{k\in{\mathbb{Z}}},
\{V(z,k)\}_{k\in{\mathbb{Z}}}$ be two ${\mathbb{C}^{m\times m}}$-valued sequences. Then the
following items $(i)$--$(iii)$ are equivalent:
\begin{align}
(i) & \quad ({\mathbb{U}} U(z,\cdot))(k) = z U(z,k), \quad ({\mathbb{W}} U(z,\cdot))(k)=z
V(z,k), \quad k\in{\mathbb{Z}}.
\\
(ii) & \quad ({\mathbb{W}} U(z,\cdot))(k) = z V(z,k), \quad ({\mathbb{V}} V(z,\cdot))(k) =
U(z,k), \quad k\in{\mathbb{Z}}. \label{3.21}
\\
(iii) & \quad \binom{U(z,k)}{V(z,k)} = {\mathbb{T}}(z,k)
\binom{U(z,k-1)}{V(z,k-1)}, \quad k\in{\mathbb{Z}}.\label{3.22}
\end{align}
Here ${\mathbb{U}}$, ${\mathbb{V}}$, and ${\mathbb{W}}$ are understood in the sense of difference
expressions on $\smm{{\mathbb{Z}}}$ rather than difference operators on
$\ltm{{\mathbb{Z}}}$ $($cf.\ Remark \ref{r3.1}$)$ and the transfer matrices
${\mathbb{T}}(z,k)$, $z\in{\C\backslash\{0\}}$, $k\in{\mathbb{Z}}$, are defined by
\begin{equation}\label{3.23}
{\mathbb{T}}(z,k) = \begin{cases}
\begin{pmatrix}
\widetilde\rho_{k}^{-1}\alpha_{k} & z\widetilde\rho_{k}^{-1} \\
z^{-1}\rho_{k}^{-1} & \rho_{k}^{-1}\alpha_{k}^*
\end{pmatrix}, & \text{$k$ odd,}
\\[20pt]
\begin{pmatrix}
\rho_{k}^{-1}\alpha_{k}^* & \rho_{k}^{-1} \\
\widetilde\rho_{k}^{-1} & \widetilde\rho_{k}^{-1}\alpha_{k}
\end{pmatrix}, & \text{$k$ even.}
\end{cases}
\end{equation}
\end{lemma}
\begin{definition}\label{d3.4}
If for some reference point, $k_0\in{\mathbb{Z}}$, we modify Hypothesis~\ref{h3.2} by allowing $\alpha_{k_0}=\gamma$, where $\gamma\in\mathbb{C}^{m\times m}$ is unitary, then the resulting operators defined in \eqref{3.15}--\eqref{3.17} will be denoted by
${\mathbb{V}}_{k_0}^{(\gamma)}, \ {\mathbb{W}}_{k_0}^{(\gamma)}$ and ${\mathbb{U}}_{k_0}^{(\gamma)}$.
\end{definition}
\begin{remark}
Strictly, speaking, allowing $\alpha_{k_0}$ to be unitary for some reference point $k_0\in{\mathbb{Z}}$ contradicts our basic Hypothesis \ref{h3.2}.
However, as long as the exception to Hypothesis \ref{h3.2} refers to
only one site, we will safely ignore this inconsistency in favor of
the notational simplicity it provides by avoiding the introduction of
a properly modified hypothesis on $\alpha=\{\alpha_k\}_{k\in{\mathbb{Z}}}$ and will refer to ${\mathbb{U}}_{k_0}^{(\gamma)}$ as a CMV operator.
\end{remark}
The operator ${\mathbb{U}}_{k_0}^{(\gamma)}$ splits into a direct sum of two half-lattice operators ${\mathbb{U}}_{-,k_0-1}^{(\gamma)}$ and ${\mathbb{U}}_{-,k_0}^{(\gamma)}$ acting on $\ltm{(-\infty,k_0-1]\cap{\mathbb{Z}}}$ and on $\ltm{[k_0,\infty)\cap{\mathbb{Z}}}$,
respectively. Explicitly, one obtains
\begin{align}
{\mathbb{U}}_{k_0}^{(\gamma)} = {\mathbb{U}}_{-,k_0-1}^{(\gamma)} \oplus {\mathbb{U}}_{+,k_0}^{(\gamma)} \, \text{ in } \,
\ltm{(-\infty,k_0-1]\cap{\mathbb{Z}}} \oplus \ltm{[k_0,\infty)\cap{\mathbb{Z}}}.
\end{align}
Similarly,
one obtains ${\mathbb{W}}_{-,k_0-1}^{(\gamma)}$, ${\mathbb{V}}_{-,k_0-1}^{(\gamma)}$ and ${\mathbb{W}}_{+,k_0}^{(\gamma)}$,
${\mathbb{V}}_{+,k_0}^{(\gamma)}$ such that
\begin{equation}\label{3.25}
{\mathbb{V}}_{k_0}^{(\gamma)}={\mathbb{V}}_{-,k_0-1}^{(\gamma)}\oplus {\mathbb{V}}_{+,k_0}^{(\gamma)},
\quad {\mathbb{W}}_{k_0}^{(\gamma)}={\mathbb{W}}_{-,k_0-1}^{(\gamma)}\oplus {\mathbb{W}}_{+,k_0}^{(\gamma)},
\end{equation}
and hence
\begin{equation}\label{3.26}
{\mathbb{U}}_{\pm,k_0}^{(\gamma)} = {\mathbb{V}}_{\pm,k_0}^{(\gamma)} {\mathbb{W}}_{\pm,k_0}^{(\gamma)}, \quad
{\mathbb{U}}_{k_0}^{(\gamma)} = {\mathbb{V}}_{k_0}^{(\gamma)} {\mathbb{W}}_{k_0}^{(\gamma)}.
\end{equation}
For the special case when $\gamma=I_m$, we simplify our notation by writing
\begin{align}
&{\mathbb{U}}_{k_0}={\mathbb{U}}_{k_0}^{(\gamma= I_m)}={\mathbb{V}}_{k_0}^{(\gamma= I_m)}{\mathbb{W}}_{k_0}^{(\gamma= I_m)}=
{\mathbb{V}}_{k_0}{\mathbb{W}}_{k_0}\label{3.27}\\
&{\mathbb{U}}_{\pm,k_0}={\mathbb{U}}_{\pm,k_0}^{(\gamma= I_m)}={\mathbb{V}}_{\pm,k_0}^{(\gamma= I_m)}{\mathbb{W}}_{\pm,k_0}^{(\gamma= I_m)}=
{\mathbb{V}}_{\pm,k_0}{\mathbb{W}}_{\pm,k_0}
\end{align}
In analogy with the scalar case, when emphasizing dependence of these operators on the sequence $\alpha=\{\alpha_k\}_{k\in{\mathbb{Z}}}$, we write, for example, ${\mathbb{U}}_{\alpha}$ and ${\mathbb{U}}_{\alpha;k_0}^{(\gamma)}$. Also in analogy to the scalar case, let $\sigma, \tau\in{\mathbb{C}^{m\times m}}$ be unitary, and let $\mathbb{A}$ and $\widetilde\mathbb{A}$ be the unitary operators defined on $\ltm{{\mathbb{Z}}}$ by
\begin{equation}\label{3.29}
\mathbb{A}=\sigma\delta_{\rm odd} + \tau\delta_{\rm even}, \quad \widetilde\mathbb{A}=\tau\delta_{\rm odd} + \sigma\delta_{\rm even},
\end{equation}
Then, for the full-lattice and direct sum of half-lattice CMV operators, we obtain the following analogs of \eqref{2.17b} and \eqref{2.17c}:
\begin{align}
& \mathbb{A}{\mathbb{U}}_{\beta}\mathbb{A}^* = \big[\mathbb{A}{\mathbb{V}}_{\beta}\widetilde\mathbb{A}^*\big] \big[\widetilde\mathbb{A}{\mathbb{W}}_{\beta}\mathbb{A}^*\big]
= {\mathbb{V}}_{\alpha}{\mathbb{W}}_{\alpha} = {\mathbb{U}}_{\alpha},\label{3.30}
\\
& \mathbb{A}{\mathbb{U}}_{\beta;k_0}^{(\gamma)}\mathbb{A}^* = \big[\mathbb{A}{\mathbb{V}}_{\beta;k_0}^{(\gamma)}\widetilde\mathbb{A}^*\big]
\big[\widetilde\mathbb{A}{\mathbb{W}}_{\beta;k_0}^{(\gamma)}\mathbb{A}^*\big] =
{\mathbb{V}}_{\alpha;k_0}^{(\sigma\gamma\tau^*)}{\mathbb{W}}_{\alpha;k_0}^{(\sigma\gamma\tau^*)} =
{\mathbb{U}}_{\alpha;k_0}^{(\sigma\gamma\tau^*)}.\label{3.31}
\end{align}
where $\beta=\{\beta_k\}_{k\in{\mathbb{Z}}}$, and $\alpha_k=\sigma\beta_k\tau^*$.
In particular, when $\sigma=\gamma^{-1/2}$, $\tau=\gamma^{1/2}$, we note, by \eqref{3.27} and \eqref{3.31}, that
\begin{equation}\label{3.32}
\mathbb{A}{\mathbb{U}}_{\beta;k_0}^{(\gamma)}\mathbb{A}^*={\mathbb{V}}_{\alpha;k_0}{\mathbb{W}}_{\alpha;k_0}={\mathbb{U}}_{\alpha;k_0}.
\end{equation}
The unitary transformations cited in \eqref{3.29}--\eqref{3.31} are relevant to our next result because of the following observation about $n\times n$ complex matrices:
Let $\alpha\in{\mathbb{C}^{m\times m}}$, where we are not assuming that $\alpha$ is unitary. Then, $\alpha$ has a (not necessarily unique) polar decomposition, $\alpha=U|\alpha|$, where $U, |\alpha|\in{\mathbb{C}^{m\times m}}$, $U$ is unitary, and $|\alpha|=(\alpha^*\alpha)^{1/2}\geq 0$ is nonnegative (cf., e.g.,
\cite[Theorem\ 3.1.9 (c)]{HJ94}). The nonnegative matrix $|\alpha|$ can then be diagonalized: $|\alpha|=U^*_0DU_0$, where $U_0, D\in{\mathbb{C}^{m\times m}}$, $U_0$ is unitary, and $D$ is diagonal. Thus, each $\alpha\in{\mathbb{C}^{m\times m}}$ has a (not necessarily unique) factorization of the form
\begin{equation}\label{3.33}
\alpha = \sigma\beta\tau,
\end{equation}
where, $\sigma,\beta,\tau\in{\mathbb{C}^{m\times m}}$, where $\sigma$ and $\tau$ are unitary, and where $\beta$ is diagonal.
We now present our principal result on
rank~$m$ perturbations:
\begin{theorem}\label{t3.6}
Given ${\mathbb{U}}_{\alpha}$, fix $k_0\in{\mathbb{Z}}$ and let
$\alpha_{k_0}=\sigma_{k_0}\beta_{k_0}\tau_{k_0}^*$ be a factorization
for $\alpha_{k_0}$, as described in \eqref{3.33}, where $\sigma_{k_0},\
\tau_{k_0}\in{\mathbb{C}^{m\times m}}$ are unitary, and where
$\beta_{k_0}\in{\mathbb{C}^{m\times m}}$ is the diagonal matrix
$\beta_{k_0}=\diag[\beta_{k_0,1},\dots,\beta_{k_0,m}].$ Let
$\gamma_j=\sigma_{k_0}\theta_j\tau_{k_0}^*$, $j=1,2$, where
$\theta_1=\diag[e^{it_1},\dots,e^{it_m}]$, and
$\theta_2=\diag[e^{is_1},\dots,e^{is_m}]$. Let
${\mathbb{U}}_{\alpha;k_0}^{(\gamma_1,\gamma_2)}$ denote the unitary operator acting
on $\ltm{{\mathbb{Z}}}$ and defined by
\begin{equation}
{\mathbb{U}}_{\alpha;k_0}^{(\gamma_1,\gamma_2)}={\mathbb{U}}_{\alpha;-,k_0-1}^{(\gamma_1)}\oplus{\mathbb{U}}_{\alpha;+,k_0}^{(\gamma_2)}.
\end{equation}
Then, for an arbitrary unitary matrix $\gamma\in{\mathbb{C}^{m\times m}}$, the difference ${\mathbb{U}}_{\alpha}-{\mathbb{U}}_{\alpha;k_0}^{(\gamma)}$ has rank greater than $m$, while the differences ${\mathbb{U}}_{\alpha}-{\mathbb{U}}_{\alpha;k_0}^{(\gamma_1,\gamma_2)}$ and $({\mathbb{U}}_{\alpha}-zI)^{-1} - \big({\mathbb{U}}_{\alpha;k_0}^{(\gamma_1,\gamma_2)} -zI\big)^{-1}$ are of rank~$m$ if and only if \
$t_j=2\arg [i(\beta_{k_0,j}e^{-is_j/2} - e^{is_j/2})]$, $j=1,\dots,m$, and otherwise possess rank greater than $m$.
\end{theorem}
\begin{proof}
${\mathbb{U}}^{(\gamma)}_{\alpha;k_0}={\mathbb{V}}^{(\gamma)}_{\alpha;k_0}{\mathbb{W}}^{(\gamma)}_{\alpha;k_0}$ by
\eqref{3.26}, and hence, it follows from \eqref{3.15} that
\begin{align}
{\mathbb{U}}_{\alpha}-{\mathbb{U}}^{(\gamma)}_{\alpha;k_0} =
\begin{cases}
{\mathbb{V}}_{\alpha}\big({\mathbb{W}}_{\alpha}-{\mathbb{W}}^{(\gamma)}_{\alpha;k_0}\big), & \text{$k_0$ odd,} \\
\big({\mathbb{V}}_{\alpha}-{\mathbb{V}}^{(\gamma)}_{\alpha;k_0}\big){\mathbb{W}}_{\alpha}, & \text{$k_0$ even.}
\end{cases}
\end{align}
For $k_0$ odd, $D={\mathbb{W}}_{\alpha}-{\mathbb{W}}^{(\gamma)}_{\alpha;k_0}$ is block-diagonal with all
of its $2m\times 2m$ blocks on the diagonal being zero except for one
which has the following form
\begin{equation}
A=
\begin{pmatrix}
D_{k_0-1,k_0-1} & D_{k_0-1,k_0} \\ D_{k_0,k_0-1} & D_{k_0,k_0}
\end{pmatrix}
=
\begin{pmatrix}
-\alpha_{k_0}+\gamma & \widetilde\rho_{k_0} \\ \rho_{k_0} & \alpha_{k_0}^*-\gamma^*
\end{pmatrix}
\end{equation}
Note that the following subspaces have only trivial intersection
with $\text{\rm{ker}}(A)$ since $\rho_{k_0}$ and $\widetilde\rho_{k_0}$ are
invertible matrices,
\begin{equation}
S_1=\left\{ \begin{pmatrix}\xi\\0 \end{pmatrix}\in\mathbb{C}^{2m}\mid \xi\in\mathbb{C}^{m}\right\},\quad
S_2=\left\{ \begin{pmatrix}0\\ \eta \end{pmatrix}\in\mathbb{C}^{2m}\mid \eta\in\mathbb{C}^{m}\right\}.
\end{equation}
As a consequence, $\text{\rm{rank}}(A)\ge m$. Further note, for any $c\in\mathbb{C}$, that
\begin{equation}\label{3.38}
\begin{pmatrix}\xi \\ c\xi \end{pmatrix} \in \text{\rm{ker}}(A)
\end{equation}
only when $\xi=0\in\mathbb{C}^m$. To see this, assume that \eqref{3.38}
holds for some $c\in\mathbb{C}$ and some $\xi\ne 0\in\mathbb{C}^m$. It follows that
\begin{equation}\label{3.39}
\begin{pmatrix} 0\\0\end{pmatrix}
= \begin{pmatrix} c\xi\\ \xi \end{pmatrix}^*A \begin{pmatrix} \xi\\ c\xi \end{pmatrix}=
\begin{pmatrix}-\bar c \xi^*(\alpha_{k_0}-\gamma)\xi+ |c|^2 \xi^*\widetilde\rho_{k_0}\xi\\
\xi^*\rho\xi + c\xi^*(\alpha_{k_0}-\gamma)^*\xi \end{pmatrix}
\end{equation}
and hence by conjugation in the first line of \eqref{3.39} that
\begin{align}
0&=-c \xi^*(\alpha_{k_0}-\gamma)^*\xi+ |c|^2 \xi^*\widetilde\rho_{k_0}\xi,\\
0&=\xi^*\rho\xi + c\xi^*(\alpha_{k_0}-\gamma)^*\xi.
\end{align}
Summing these equations, we see that
$\xi^*(\rho_{k_0}+|c|^2\widetilde\rho_{k_0})\xi=0$. However, strict
positivity of the self-adjoint matrix
$\rho_{k_0}+|c|^2\widetilde\rho_{k_0}$ implies that $\xi=0\in\mathbb{C}^m$; a
contradiction.
Noting again that $\text{\rm{rank}}(A)\ge m$, assume that $\text{\rm{rank}}(A)=m$. Given that
\begin{equation}
\mathbb{C}^{2m} = S_1\oplus S_2, \quad \text{\rm{ker}}(A)\cap S_j =\{0\}, j=1,2,
\end{equation}
then there exists a matrix $M\in{\mathbb{C}^{m\times m}}$ such that
\begin{equation}
\text{\rm{ker}}(A) = \left\{\begin{pmatrix}\xi\{\mathcal{M}}\xi \end{pmatrix} \mid \xi\in\mathbb{C}^m \right\}.
\end{equation}
However, this implies the existence of some $\xi\ne 0\in\mathbb{C}^m$ and some $c\in\mathbb{C}$ such that \eqref{3.38} holds; thus, again a contradiction. Hence, the rank of $A$, and as a consequence the rank of ${\mathbb{U}}_{\alpha}-{\mathbb{U}}_{\alpha;k_0}^{(\gamma)}$, are greater than $m$. The proof when $k_0\in{\mathbb{Z}}$ is even is similar to that just given.
Now consider the rank of the difference ${\mathbb{U}}_{\alpha}-{\mathbb{U}}_{\alpha;k_0}^{(\gamma_1,\gamma_2)}$, first noting by \eqref{3.25} and \eqref{3.26} that ${\mathbb{U}}^{(\gamma_1,\gamma_2)}_{\alpha;k_0}={\mathbb{V}}^{(\gamma_1,\gamma_2)}_{\alpha;k_0}{\mathbb{W}}^{(\gamma_1,\gamma_2)}_{\alpha;k_0}$ and hence, by \eqref{3.15} that
\begin{align}
{\mathbb{U}}_{\alpha}-{\mathbb{U}}^{(\gamma_1,\gamma_2)}_{\alpha;k_0} =
\begin{cases}
{\mathbb{V}}_{\alpha}\big({\mathbb{W}}_{\alpha}-{\mathbb{W}}^{(\gamma_1,\gamma_2)}_{\alpha;k_0}\big), & \text{$k_0$ odd,} \\
\big({\mathbb{V}}_{\alpha}-{\mathbb{V}}^{(\gamma_1,\gamma_2)}_{\alpha;k_0}\big){\mathbb{W}}_{\alpha}, & \text{$k_0$ even.}
\end{cases}
\end{align}
We again proceed by assuming that $k_0\in{\mathbb{Z}}$ is odd, and letting
$D={\mathbb{W}}_{\alpha}-{\mathbb{W}}^{(\gamma_1,\gamma_2)}_{\alpha;k_0}$ be the block-diagonal matrix all of whose
$2m\times 2m$ blocks on the diagonal are zero except for one
which has the following form:
\begin{align}
\begin{pmatrix}
D_{k_0-1,k_0-1} & D_{k_0-1,k_0} \\ D_{k_0,k_0-1} & D_{k_0,k_0}
\end{pmatrix}
&=\begin{pmatrix}-\alpha_{k_0} & \widetilde\rho_{k_0} \\ \rho_{k_0} & \alpha_{k_0}^*\end{pmatrix}
-
\begin{pmatrix}-\sigma_{k_0}\theta_1\tau_{k_0}^* & 0 \\ 0 & \tau_{k_0}\theta_2^*\sigma_{k_0}^*\end{pmatrix}\\
&=
\begin{pmatrix}\sigma_{k_0} & 0 \\ 0 & \tau_{k_0}\end{pmatrix}
\begin{pmatrix}-\beta_{k_0}+\theta_1 & \kappa_{k_0} \\ \kappa_{k_0} & \beta_{k_0}^*-\theta_2^*\end{pmatrix}
\begin{pmatrix}\tau_{k_0}^* & 0 \\ 0 & \sigma_{k_0}^*\end{pmatrix},
\end{align}
where $\kappa_{k_0} = (I_m-\beta_{k_0}\beta_{k_0}^*)^{1/2} = (I_m-\beta_{k_0}^*\beta_{k_0})^{1/2}=\diag[\kappa_{k_0,1},\dots,\kappa_{k_0,m}]$. Then, the rank of the difference
$D={\mathbb{W}}_{\alpha}-{\mathbb{W}}^{(\gamma_1,\gamma_2)}_{\alpha;k_0}$ equals the rank of the matrix
\begin{equation}\label{3.47}
B=\begin{pmatrix}-\beta_{k_0}+\theta_1 & \kappa_{k_0} \\ \kappa_{k_0} & \beta_{k_0}^*-\theta_2^*\end{pmatrix}.
\end{equation}
Since all $m\times m$ matrices on the right-hand side of
\eqref{3.47} are diagonal, performing an equivalent set of
permutations on the rows and on the columns of the matrix $B$ in
\eqref{3.47} results in a block diagonal matrix with $2\times 2$
blocks arrayed along the diagonal, each of the form,
\begin{equation}
B_j=
\begin{pmatrix}
-\beta_{k_0,j}+e^{it_j}& \kappa_{k_0,j}\\ \kappa_{k_0,j}&\overline\beta_{k_0,j}-e^{-is_j}
\end{pmatrix}, \quad j=1,\dots,m.
\end{equation}
It follows that $\text{\rm{rank}}(B)=m$ precisely when $\text{\rm{rank}}(B_j) =1$ for all $j=1,\dots,m$, and otherwise, that $\text{\rm{rank}}(B)>m$. We now observe that each $B_j$ has the form given in \eqref{2.21}. Hence, by the derivation following from \eqref{2.22a} we see that $\det(B_j)=0$, and hence that $\text{\rm{rank}}(B_j)=1$, precisely when
$
t_j=2\arg [i(\beta_{k_0,j}e^{-is_j/2} - e^{is_j/2})].$
As in the scalar case discussed in Theorem~\ref{t2.3}, the statement
in this theorem for the resolvents follows from the result just
proven for the difference ${\mathbb{U}}_{\alpha}-{\mathbb{U}}_{\alpha;k_0}^{(\gamma_1,\gamma_2)}$
and from the identity,
\begin{equation}
({\mathbb{U}}_{\alpha}-zI)^{-1} - \big({\mathbb{U}}_{\alpha;k_0}^{(\gamma_1,\gamma_2)} -zI\big)^{-1}
=
-({\mathbb{U}}_{\alpha}-zI)^{-1}
\big[{\mathbb{U}}_{\alpha}-{\mathbb{U}}^{(\gamma_1,\gamma_2)}_{\alpha;k_0}\big] \big({\mathbb{U}}^{(\gamma_1,\gamma_2)}_{\alpha;k_0}-zI\big)^{-1}.
\end{equation}
\end{proof}
\begin{remark}
In particular, we note that the difference $\big[{\mathbb{U}}_{\alpha}-{\mathbb{U}}_{\alpha;k_0}^{(\gamma)}\big]$ has rank greater than $m$ when $\gamma = I_m$.
\end{remark}
Next we present formulas for
different unitary $\gamma\in{\mathbb{C}^{m\times m}}$, linking various spectral theoretic objects
associated with half-lattice CMV operators ${\mathbb{U}}^{(\gamma)}_{\pm,k_0}$. For the special case when $\gamma=I_m$, these objects, and the relationships described below, have proven exceptionally useful (see, e.g., \cite{CGZ07}, \cite{CGZ08}, \cite{GZ06}, \cite{GZ06a}, and \cite{Zi08}).
We begin with an analog of Lemma
\ref{l3.3} for difference expressions ${\mathbb{U}}^{(\gamma)}_{\pm,k_0}$,
${\mathbb{V}}^{(\gamma)}_{\pm,k_0}$, and ${\mathbb{W}}^{(\gamma)}_{\pm,k_0}$. For the special case
$\gamma=I_m$, the result below is proven in \cite[Lemma 2.4]{CGZ07}. The general case
below follows immediately from the special case and the observation
of unitary equivalence in \eqref{3.32}.
\begin{lemma} \label{l3.8}
Let $z\in\mathbb{C}\backslash\{0\}$, $k_0\in{\mathbb{Z}}$, and let $\gamma\in{\mathbb{C}^{m\times m}}$ be unitary.
Let
$\big\{\widehat P^{(\gamma)}_+(z,k,k_0)\big\}_{k\geq k_0}$,
$\big\{\widehat R^{(\gamma)}_+(z,k,k_0)\big\}_{k\geq k_0}$ be two ${\mathbb{C}^{m\times m}}$-valued sequences. Then the
following items $(i)$--$(iii)$ are equivalent:
\begin{align}
(i) &\quad \big({\mathbb{U}}^{(\gamma)}_{+,k_0} \widehat P^{(\gamma)}_+(z,\cdot,k_0)\big)(k) = z \widehat
P^{(\gamma)}_+(z,k,k_0), \notag \\
&\quad \big({\mathbb{W}}^{(\gamma)}_{+,k_0} \widehat P^{(\gamma)}_+(z,\cdot,k_0)\big)(k) = z \widehat
R^{(\gamma)}_+(z,k,k_0), \quad k\geq k_0.
\\
(ii) &\quad \big({\mathbb{W}}^{(\gamma)}_{+,k_0} \widehat P^{(\gamma)}_+(z,\cdot,k_0)\big)(k) = z
\widehat R^{(\gamma)}_+(z,k,k_0), \notag \\
&\quad \big({\mathbb{V}}^{(\gamma)}_{+,k_0} \widehat R^{(\gamma)}_+(z,\cdot,k_0)\big)(k) = \widehat
P^{(\gamma)}_+(z,k,k_0), \quad k\geq k_0.
\\
(iii) &\quad \binom{\widehat P^{(\gamma)}_+(z,k,k_0)}{\widehat R^{(\gamma)}_+(z,k,k_0)}
= {\mathbb{T}}(z,k) \binom{\widehat P^{(\gamma)}_+(z,k-1,k_0)}{\widehat
R^{(\gamma)}_+(z,k-1,k_0)}, \quad k > k_0, \notag \\
&\quad \widehat P^{(\gamma)}_+(z,k_0,k_0) =
\begin{cases}
z\gamma \widehat R^{(\gamma)}_+(z,k_0,k_0), & \text{$k_0$ odd}, \\
\gamma^* \widehat R^{(\gamma)}_+(z,k_0,k_0), & \text{$k_0$ even}.
\end{cases}
\end{align}
Similarly, let $\big\{\widehat P^{(\gamma)}_-(z,k,k_0)\big\}_{k\leq k_0}$,
$\big\{\widehat R^{(\gamma)}_-(z,k,k_0)\big\}_{k\leq k_0}$ be two ${\mathbb{C}^{m\times m}}$-valued sequences. Then the
following items $(iv)$--$(vi)$are equivalent:
\begin{align}
(iv) &\quad \big({\mathbb{U}}^{(\gamma)}_{-,k_0} \widehat P^{(\gamma)}_-(z,\cdot,k_0)\big)(k) = z
\widehat P^{(\gamma)}_-(z,k,k_0), \notag \\
&\quad \big({\mathbb{W}}^{(\gamma)}_{-,k_0} \widehat P^{(\gamma)}_-(z,\cdot,k_0)\big)(k) = z \widehat
R^{(\gamma)}_-(z,k,k_0), \quad k\leq k_0.
\\
(v) &\quad \big({\mathbb{W}}^{(\gamma)}_{-,k_0} \widehat P^{(\gamma)}_-(z,\cdot,k_0)\big)(k) = z
\widehat R^{(\gamma)}_-(z,k,k_0), \notag \\
&\quad \big({\mathbb{V}}^{(\gamma)}_{-,k_0} \widehat R^{(\gamma)}_-(z,\cdot,k_0)\big)(k) = \widehat
P^{(\gamma)}_-(z,k,k_0), \quad k\leq k_0.
\\
(vi) &\quad \binom{\widehat P^{(\gamma)}_-(z,k-1,k_0)}{\widehat
R^{(\gamma)}_-(z,k-1,k_0)}={\mathbb{T}}(z,k)^{-1} \binom{\widehat
P^{(\gamma)}_-(z,k,k_0)}{\widehat R^{(\gamma)}_-(z,k,k_0)}, \quad k \leq k_0,\notag
\\
&\quad \widehat P^{(\gamma)}_-(z,k_0,k_0) =
\begin{cases}
-\gamma\widehat R^{(\gamma)}_-(z,k_0,k_0), & \text{$k_0$ odd,}
\\
-z\gamma^* \widehat R^{(\gamma)}_-(z,k_0,k_0), & \text{$k_0$ even.}
\end{cases}
\end{align}
\end{lemma}
Next, we denote by
$\Big(\begin{smallmatrix}P^{(\gamma)}_\pm(z,k,k_0) \\
R^{(\gamma)}_\pm(z,k,k_0)\end{smallmatrix}\Big)_{k\in{\mathbb{Z}}}$ and
$\Big(\begin{smallmatrix} Q^{(\gamma)}_\pm(z,k,k_0) \\ S^{(\gamma)}_\pm(z,k,k_0)
\end{smallmatrix}\Big)_{k\in{\mathbb{Z}}}$,
$z\in{\mathbb{C}}\backslash\{0\}$, four linearly independent solutions of
\eqref{3.21} satisfying the following initial conditions:
\begin{align}
&\binom{P^{(\gamma)}_+(z,k_0,k_0)}{R^{(\gamma)}_+(z,k_0,k_0)} =
\begin{cases}
\binom{z\gamma^{1/2}}{\gamma^{-1/2}}, & \text{$k_0$ odd,} \\[1mm]
\binom{\gamma^{-1/2}}{\gamma^{1/2}}, & \text{$k_0$ even,}
\end{cases} \; \; \,
\binom{Q^{(\gamma)}_+(z,k_0,k_0)}{S^{(\gamma)}_+(z,k_0,k_0)} =
\begin{cases}
\binom{z\gamma^{1/2}}{-\gamma^{-1/2}}, & \text{$k_0$ odd,} \\[1mm]
\binom{-\gamma^{-1/2}}{\gamma^{1/2}}, & \text{$k_0$ even.}
\end{cases} \label{3.56}
\\
&\binom{P^{(\gamma)}_-(z,k_0,k_0)}{R^{(\gamma)}_-(z,k_0,k_0)} =
\begin{cases}
\binom{\gamma^{1/2}}{-\gamma^{-1/2}}, & \text{$k_0$ odd,} \\[1mm]
\binom{-z\gamma^{-1/2}}{\gamma^{1/2}}, & \text{$k_0$ even,}
\end{cases}
\binom{Q^{(\gamma)}_-(z,k_0,k_0)}{S^{(\gamma)}_-(z,k_0,k_0)} =
\begin{cases}
\binom{\gamma^{1/2}}{\gamma^{-1/2}}, & \text{$k_0$ odd, }\\[1mm]
\binom{z\gamma^{-1/2}}{\gamma^{1/2}}, & \text{$k_0$ even.}
\end{cases} \label{3.57}
\end{align}
Then, $P^{(\gamma)}_\pm(z,k,k_0)$, $Q{(\gamma)}_\pm(z,k,k_0)$,
$R{(\gamma)}_\pm(z,k,k_0)$, and $S{(\gamma)}_\pm(z,k,k_0)$, $k,k_0\in{\mathbb{Z}}$, are
${\mathbb{C}^{m\times m}}$-valued Laurent polynomials in $z$.
\begin{lemma} \label{l3.9}
Let $z\in\mathbb{C}\backslash\{0\}$, $k_0\in{\mathbb{Z}}$, and let $\gamma_j\in{\mathbb{C}^{m\times m}}$, $j=1,2$, be unitary.
With
$C_1=\frac12(\gamma_1^{-1/2}\gamma_2^{1/2} + \gamma_1^{1/2}\gamma_2^{-1/2})$ and
$D_1=\frac12(\gamma_1^{-1/2}\gamma_2^{1/2} - \gamma_1^{1/2}\gamma_2^{-1/2})$,
then,
\begin{align}
&\binom{Q^{(\gamma_2)}_\pm(z,\cdot,k_0)}{S^{(\gamma_2)}_\pm(z,\cdot,k_0)} =
\binom{Q^{(\gamma_1)}_\pm(z,\cdot,k_0)}{S^{(\gamma_1)}_\pm(z,\cdot,k_0)}C_1 +
\binom{P^{(\gamma_1)}_\pm(z,\cdot,k_0)}{R^{(\gamma_1)}_\pm(z,\cdot,k_0)}D_1\label{3.58}
\\
&\binom{P^{(\gamma_2)}_\pm(z,\cdot,k_0)}{R^{(\gamma_2)}_\pm(z,\cdot,k_0)} =
\binom{Q^{(\gamma_1)}_\pm(z,\cdot,k_0)}{S^{(\gamma_1)}_\pm(z,\cdot,k_0)}D_1 +
\binom{P^{(\gamma_1)}_\pm(z,\cdot,k_0)}{R^{(\gamma_1)}_\pm(z,\cdot,k_0)}C_1.
\end{align}
With
$C_2=(2z^{k_0 \, ({\rm mod}\,2)})^{-1}(\gamma_1^{-1/2}\gamma_2^{1/2} - z\gamma_1^{1/2}\gamma_2^{-1/2})$,
$D_2=(2z^{k_0 \, ({\rm mod}\,2)})^{-1}(\gamma_1^{-1/2}\gamma_2^{1/2} + z\gamma_1^{1/2}\gamma_2^{-1/2}),$ then
\begin{align}
&\binom{Q^{(\gamma_2)}_-(z,\cdot,k_0)}{S^{(\gamma_2)}_-(z,\cdot,k_0)} =
\binom{Q^{(\gamma_1)}_+(z,\cdot,k_0)}{S^{(\gamma_1)}_+(z,\cdot,k_0)}C_2 +
\binom{P^{(\gamma_1)}_+(z,\cdot,k_0)}{R^{(\gamma_1)}_+(z,\cdot,k_0)}D_2,
\\
&\binom{P^{(\gamma_2)}_-(z,\cdot,k_0)}{R^{(\gamma_2)}_-(z,\cdot,k_0)} =
\binom{Q^{(\gamma_1)}_+(z,\cdot,k_0)}{S^{(\gamma_1)}_+(z,\cdot,k_0)}D_2 +
\binom{P^{(\gamma_1)}_+(z,\cdot,k_0)}{R^{(\gamma_1)}_+(z,\cdot,k_0)}C_2.\label{3.61}
\end{align}
With $C_3=\frac12(\gamma_1^{-1/2}\widetilde\rho_{k_0}^{-1}\alpha_{k_0}\gamma_2^{-1/2}-
\gamma_1^{1/2}\rho_{k_0}^{-1}\alpha_{k_0}^*\gamma_2^{1/2}) +
\frac12(\gamma_1^{-1/2}\widetilde\rho_{k_0}^{-1}\gamma_2^{1/2}-
\gamma_1^{1/2}\rho_{k_0}^{-1}\gamma_2^{-1/2})$ and
$D_3=\frac12(\gamma_1^{-1/2}\widetilde\rho_{k_0}^{-1}\alpha_{k_0}\gamma_2^{-1/2}+
\gamma_1^{1/2}\rho_{k_0}^{-1}\alpha_{k_0}^*\gamma_2^{1/2}) +
\frac12(\gamma_1^{-1/2}\widetilde\rho_{k_0}^{-1}\gamma_2^{1/2}+
\gamma_1^{1/2}\rho_{k_0}^{-1}\gamma_2^{-1/2}),$
then
\begin{equation}
\binom{Q^{(\gamma_2)}_-(z,\cdot,k_0-1)}{S^{(\gamma_2)}_-(z,\cdot,k_0-1)} =
\binom{Q^{(\gamma_1)}_+(z,\cdot,k_0)}{S^{(\gamma_1)}_+(z,\cdot,k_0)}C_3+
\binom{P^{(\gamma_1)}_+(z,\cdot,k_0)}{R^{(\gamma_1)}_+(z,\cdot,k_0)}D_3.\label{3.62}
\end{equation}
With $C_4 =-\frac12(\gamma_1^{-1/2}\widetilde\rho_{k_0}^{-1}\alpha_{k_0}\gamma_2^{-1/2}+
\gamma_1^{1/2}\rho_{k_0}^{-1}\alpha_{k_0}^*\gamma_2^{1/2}) +
\frac12(\gamma_1^{-1/2}\widetilde\rho_{k_0}^{-1}\gamma_2^{1/2}+
\gamma_1^{1/2}\rho_{k_0}^{-1}\gamma_2^{-1/2})$ and
$D_4=-\frac12(\gamma_1^{-1/2}\widetilde\rho_{k_0}^{-1}\alpha_{k_0}\gamma_2^{-1/2}-
\gamma_1^{1/2}\rho_{k_0}^{-1}\alpha_{k_0}^*\gamma_2^{1/2}) +
\frac12(\gamma_1^{-1/2}\widetilde\rho_{k_0}^{-1}\gamma_2^{1/2}-
\gamma_1^{1/2}\rho_{k_0}^{-1}\gamma_2^{-1/2}),$
\begin{equation}
\binom{P^{(\gamma_2)}_-(z,\cdot,k_0-1)}{R^{(\gamma_2)}_-(z,\cdot,k_0-1)} =
\binom{Q^{(\gamma_1)}_+(z,\cdot,k_0)}{S^{(\gamma_1)}_+(z,\cdot,k_0)}C_4 +
\binom{P^{(\gamma_1)}_+(z,\cdot,k_0)}{R^{(\gamma_1)}_+(z,\cdot,k_0)}D_4.\label{3.63}
\end{equation}
\end{lemma}
\begin{proof}
Since each sequence in equations \eqref{3.58}--\eqref{3.63} satisfies the recurrence relation \eqref{3.22}, it is necessary only to check equality at $k=k_0$. Substituting \eqref{3.56}, \eqref{3.57} into \eqref{3.58}--\eqref{3.61} suffices to verify the identities. In addition to use of \eqref{3.56}, \eqref{3.57}, application of the transfer matrix ${\mathbb{T}}(z,k_0)$ in \eqref{3.23} to the left sides of \eqref{3.62}, \eqref{3.63} suffices in the verification of the later two identities.
\end{proof}
Next, we introduce half-lattice Weyl-Titchmarsh m-functions associated with the half-lattice CMV operators, ${\mathbb{U}}_{\pm,k_0}^{(\gamma)}$, described in \eqref{3.26}.
Let $\Delta_k=\{\Delta_k(\ell)\}_{\ell\in{\mathbb{Z}}}\in\smm{{\mathbb{Z}}}$, $k\in{\mathbb{Z}}$, denote
the sequences of $m\times m$ matrices defined by
\begin{align}
(\Delta_k)(\ell) =
\begin{cases}
I_m, & \ell=k,\\
0, & \ell\neq k,
\end{cases}
\quad k,\ell\in{\mathbb{Z}}.
\end{align}
Then using right-multiplication by $m\times m$ matrices on
$\smm{{\mathbb{Z}}}$ defined in Remark \ref{r3.1}, we get the
identity
\begin{align}
(\Delta_k X)(\ell) =
\begin{cases}
X, & \ell=k,\\
0, & \ell\neq k,
\end{cases} \quad X\in{\mathbb{C}^{m\times m}},
\end{align}
and hence consider $\Delta_k$ as a map $\Delta_k\colon {\mathbb{C}^{m\times m}}\to\smm{{\mathbb{Z}}}$. In
addition, we introduce the map $\Delta_k^*\colon \smm{{\mathbb{Z}}}\to{\mathbb{C}^{m\times m}}$,
$k\in{\mathbb{Z}}$, defined by
\begin{align}
\Delta_k^* \Phi = \Phi(k), \,\text{ where }\,
\Phi=\{\Phi(k)\}_{k\in{\mathbb{Z}}}\in\smm{{\mathbb{Z}}}.
\end{align}
Similarly, one introduces the corresponding maps with ${\mathbb{Z}}$ replaced by
$[k_0,\pm\infty)\cap{\mathbb{Z}}$, $k_0\in{\mathbb{Z}}$, which, for notational brevity, we
will also denote by $\Delta_k$ and $\Delta_k^*$, respectively.
For $k_0\in{\mathbb{Z}}$, and for a unitary $\gamma\in{\mathbb{C}^{m\times m}}$, let $d\Omega_\pm^{(\gamma)}(\cdot,k_0)$ denote the ${\mathbb{C}^{m\times m}}$-valued measures on ${\partial\hspace*{.2mm}\mathbb{D}}$
defined by
\begin{equation}
d\Omega_\pm^{(\gamma)}(\zeta,k_0) = d(\Delta_{k_0}^* E_{{\mathbb{U}}_{\pm,k_0}^{(\gamma)}}(\zeta)\Delta_{k_0}),
\quad \zeta\in{\partial\hspace*{.2mm}\mathbb{D}},
\end{equation}
where $E_{{\mathbb{U}}_{\pm,k_0}^{(\gamma)}}(\cdot)$ denotes the family of spectral projections for
the half-lattice unitary operators ${\mathbb{U}}_{\pm,k_0}^{(\gamma)}$,
\begin{equation}
{\mathbb{U}}_{\pm,k_0}^{(\gamma)}=\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}dE_{{\mathbb{U}}_{\pm,k_0}^{(\gamma)}}(\zeta)\,\zeta.
\end{equation}
Then, the half-lattice Weyl-Titchmarsh m-functions, $m_\pm^{(\gamma)}(z,k_0)$, are defined by
\begin{align}
m_\pm^{(\gamma)}(z,k_0) &= \pm \Delta_{k_0}^*({\mathbb{U}}_{\pm,k_0}^{(\gamma)}+zI)({\mathbb{U}}_{\pm,k_0}^{(\gamma)}-zI)^{-1}
\Delta_{k_0} \label{3.69}
\\
& =\pm
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}d\Omega_{\pm}^{(\gamma)}(\zeta,k_0)\,\frac{\zeta+z}{\zeta-z},
\quad z\in{\mathbb{C}}\backslash{\partial\hspace*{.2mm}\mathbb{D}},\label{3.70}
\end{align}
with
\begin{equation}
\oint_{{\partial\hspace*{.2mm}\mathbb{D}}}d\Omega_{\pm}^{(\gamma)}(\zeta,k_0)= I_m.
\end{equation}
As defined, we note that $m_\pm^{(\gamma)}(z,k)$ are matrix-valued Caratheodory functions: see Appendix~\ref{sA} for definition and further properties.
Proof of the next result in the special case when $\gamma=I_m$ can be found in \cite[Lemma~2.13, Theorem~2.17]{CGZ07}.
The proof in the case for general unitary $\gamma\in{\mathbb{C}^{m\times m}}$ follows
the same lines and is omitted here for the sake of brevity.
\begin{theorem} \label{t3.10}
Let $k_0\in{\mathbb{Z}}$. Then, for each unitary $\gamma\in{\mathbb{C}^{m\times m}}$, for $P_{\pm,k_0}^{(\gamma)}$, $Q_{\pm,k_0}^{(\gamma)}$, $R_{\pm,k_0}^{(\gamma)}$, $S_{\pm,k_0}^{(\gamma)}$ defined in \eqref{3.56} and \eqref{3.57}, and for $m_\pm^{(\gamma)}(z,k_0)$ defined in \eqref{3.69}, the following relations hold for $z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\})$:
\begin{align}
\widehat U^{(\gamma)}_\pm(z,\cdot,k_0)=Q_\pm^{(\gamma)}(z,\cdot,k_0) + P_\pm^{(\gamma)}(z,\cdot,k_0)m_\pm^{(\gamma)}(z,k_0) \in
\ltmm{[k_0,\pm\infty)\cap{\mathbb{Z}}},\label{3.72}
\\
\widehat V^{(\gamma)}_\pm(z,\cdot,k_0)=S_\pm^{(\gamma)}(z,\cdot,k_0) + R_\pm^{(\gamma)}(z,\cdot,k_0)m_\pm^{(\gamma)}(z,k_0) \in
\ltmm{[k_0,\pm\infty)\cap{\mathbb{Z}}}.\label{3.73}
\end{align}
Moreover, there exist unique ${\mathbb{C}^{m\times m}}$-valued functions
$M_\pm^{(\gamma)}(\cdot,k_0)$ such that for all
$z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\})$,
\begin{align}
&U_\pm^{(\gamma)}(z,\cdot,k_0) = Q_+^{(\gamma)}(z,\cdot,k_0) + P_+^{(\gamma)}(z,\cdot,k_0)M_\pm^{(\gamma)}(z,k_0)
\in \ltmm{[k_0,\pm\infty)\cap{\mathbb{Z}}},\label{3.74}
\\
&V_\pm^{(\gamma)}(z,\cdot,k_0) = S_+^{(\gamma)}(z,\cdot,k_0) + R_+^{(\gamma)}(z,\cdot,k_0)M_\pm^{(\gamma)}(z,k_0)
\in \ltmm{[k_0,\pm\infty)\cap{\mathbb{Z}}}.\label{3.75}
\end{align}
\end{theorem}
\begin{remark}\label{r3.11}
Within the proof of Theorem~\ref{t3.10}, one observes that the sequences,
\begin{equation}\label{3.76}
\binom{U_\pm^{(\gamma)}(z,k,k_0)}{V_\pm^{(\gamma)}(z,k,k_0)}_{k\in{\mathbb{Z}}},
\end{equation}
defined by \eqref{3.74} and \eqref{3.75},
are unique up to right-multiplication by constant $m\times m$ matrices.
Hence, we shall call $U_\pm^{(\gamma)}(z,\cdot,k_0)$ the Weyl--Titchmarsh solutions of ${\mathbb{U}}$. Note also that $m_\pm^{(\gamma)}(z,k_0)$, as well as $M_\pm^{(\gamma)}(z,k_0)$, are said to be half-lattice Weyl--Titchmarsh $m$-functions associated with
${\mathbb{U}}_{\pm,k_0}^{(\gamma)}$. (See also \cite{Si04a} for a comparison of various
alternative notions of Weyl--Titchmarsh $m$-functions for
${\mathbb{U}}_{+,k_0}^{(\gamma)}$ with scalar-valued Verblunsky coefficients.)
\end{remark}
For fixed $k_0\in{\mathbb{Z}}$, $z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\})$, and unitary $\gamma\in{\mathbb{C}^{m\times m}}$, using \eqref{3.70} and Theorem~\ref{t3.10}, we obtain,
\begin{equation}\label{3.77}
M_+^{(\gamma)}(z,k_0)=m_+^{(\gamma)}(z,k_0),\quad
M_+^{(\gamma)}(0,k_0)=I_m.
\end{equation}
In particular, by \eqref{3.77} and the uniqueness up to right multiplication by constant $m\times m$ matrices noted in Remark~\ref{r3.11}, note for \eqref{3.72}--\eqref{3.75} that
\begin{equation}
\widehat U^{(\gamma)}_+( z,k,k_0)= U^{(\gamma)}_+( z,k,k_0), \quad
\widehat V^{(\gamma)}_+( z,k,k_0)= V^{(\gamma)}_+( z,k,k_0).
\end{equation}
Following the line of reasoning presented for the special case when $\gamma=I_m$ found in \cite{CGZ07}, one uses the equations in Lemma~\ref{l3.9}, together with Theorem~\ref{t3.10}, to obtain the following equations where $C_3,\ D_3$, and $ C_4,\ D_4$ are defined in \eqref{3.62} and \eqref{3.63} respectively, with $\gamma=\gamma_1=\gamma_2$.
\begin{align}
M_-^{(\gamma)}(z,k_0)&=
\big[D_3+D_4m_-^{(\gamma)}(z,k_0-1)\big]\big[C_3+C_4m_-^{(\gamma)}(z,k_0-1)\big]^{-1}
\label{3.79}\\
&=\big[m_-^{(\gamma)}(z,k_0)+I_m + z(m_-^{(\gamma)}(z,k_0)-I_m)\big] \\
&\hspace{15pt}\times
\big[m_-^{(\gamma)}(z,k_0)+I_m - z(m_-^{(\gamma)}(z,k_0)-I_m)\big]^{-1},\,
z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\}), \notag \\
M_-^{(\gamma)}(0,k_0)&=
[\gamma^{-1/2}\widetilde\rho^{-1}_{k_0}\gamma^{-1/2}+\gamma^{1/2}\rho^{-1}_{k_0}\gamma^{1/2}]
[\gamma^{-1/2}\widetilde\rho^{-1}_{k_0}\gamma^{-1/2}-\gamma^{1/2}\rho^{-1}_{k_0}\gamma^{1/2}]^{-1},\label{3.81}\\
m_-^{(\gamma)}(z,k_0)&= \big[z(M_-^{(\gamma)}(z,k_0) +I_m) - (M_-^{(\gamma)}(z,k_0)-I_m)\big]
\label{3.82} \\
&\hspace{15pt}\times \big[z(M_-^{(\gamma)}(z,k_0) +I_m) + (M_-^{(\gamma)}(z,k_0)-I_m)\big]^{-1},\, z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\}). \notag
\end{align}
By their relations to the Caratheodory functions $m_\pm^{(\gamma)}(z,k)$ given in \eqref{3.77} and \eqref{3.79}, we see that $M_\pm^{(\gamma)}(z,k)$ are also matrix-valued Caratheodory functions.
Next, we introduce the ${\mathbb{C}^{m\times m}}$-valued Schur functions $\Phi_\pm^{(\gamma)}(\cdot,k)$,
$k\in{\mathbb{Z}}$, by
\begin{align}
\Phi_\pm^{(\gamma)}(z,k) = \big[M_\pm^{(\gamma)}(z,k)-I_m\big]
\big[M_\pm^{(\gamma)}(z,k)+I_m\big]^{-1}, \quad
z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}}. \label{3.83}
\end{align}
Then, by \eqref{3.82} and \eqref{3.83}, one verifies that
\begin{align}
M_\pm^{(\gamma)}(z,k) &= \big[I_m-\Phi_\pm^{(\gamma)}(z,k)\big]^{-1}
\big[I_m+\Phi_\pm^{(\gamma)}(z,k)\big], \quad
z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}},
\\
m_-^{(\gamma)}(z,k) &= \big[zI_m+\Phi_-^{(\gamma)}(z,k)\big]^{-1}
\big[zI_m-\Phi_-^{(\gamma)}(z,k)\big], \quad
z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}}. \label{3.85}
\end{align}
Moreover, by \eqref{3.56}, \eqref{3.83}, and Theorem~\ref{t3.10}, it follows, as in \cite[Lemma 2.18]{CGZ07}, that for $k\in{\mathbb{Z}},\ z\in\mathbb{C}\backslash{\partial\hspace*{.2mm}\mathbb{D}}$,
\begin{equation}\label{3.86}
\Phi_\pm^{(\gamma)}(z,k) =
\begin{cases}
z\gamma^{1/2}V_\pm^{(\gamma)}(z,k,k_0)U_\pm^{(\gamma)}(z,k,k_0)^{-1}\gamma^{1/2},& k \text{ odd,}\\
\gamma^{1/2}U_\pm^{(\gamma)}(z,k,k_0)V_\pm^{(\gamma)}(z,k,k_0)^{-1}\gamma^{1/2},& k \text{ even,}
\end{cases}
\end{equation}
where $U_\pm^{(\gamma)}(z,k,k_0)$ and $V_\pm^{(\gamma)}(z,k,k_0)$ are sequences defined in \eqref{3.74} and \eqref{3.75}. Since the Weyl-Titchmarsh solutions defined in \eqref{3.76} are unique up to right multiplication by a constant complex $m\times m$ matirx, \eqref{3.86} implies that $\gamma^{-1/2}\Phi_\pm^{(\gamma)}(\cdot,k)\gamma^{-1/2}$ is $\gamma$-independent, and hence for unitary $\gamma_1,\gamma_2\in{\mathbb{C}^{m\times m}}$, and $k\in{\mathbb{Z}}$, that
\begin{align}
\Phi_\pm^{(\gamma_2)}(\cdot,k)&=\gamma_2^{1/2}\gamma_1^{-1/2}\Phi_\pm^{(\gamma_1)}(\cdot,k)\gamma_1^{-1/2}\gamma_2^{1/2},\\
M_\pm^{(\gamma_2)}(\cdot,k) &=\big[(\gamma_2^{-1/2}\gamma_1^{1/2}+\gamma_2^{1/2}\gamma_1^{-1/2})M_\pm^{(\gamma_1)}(\cdot,k)+(\gamma_2^{-1/2}\gamma_1^{1/2}-\gamma_2^{1/2}\gamma_1^{-1/2})\big]\notag\\
&\hspace{15pt}\times \big[(\gamma_2^{-1/2}\gamma_1^{1/2}-\gamma_2^{1/2}\gamma_1^{-1/2})
M_\pm^{(\gamma_1)}(\cdot,k)+(\gamma_2^{-1/2}\gamma_1^{1/2}+\gamma_2^{1/2}\gamma_1^{-1/2})\big]^{-1}.
\end{align}
Full and half-lattice resolvent operators lie at the heart of our analysis of the Weyl-Titchmarsh theory for full and half-lattice CMV operators; in particular, as a tool in obtaining our Borg-Marchenko-type uniqueness results in \cite{CGZ07} for CMV operators with matrix-valued coefficients. Hence, we conclude with a discussion of resolvent operators for a general unitary $\gamma\in{\mathbb{C}^{m\times m}}$.
First, we note the utility of the identities contained in the following lemma. This lemma was proven in \cite[Lemma~3.2]{CGZ07} for matrix-Laurent polynomial solutions of \eqref{3.22} defined by \eqref{3.56} in the special case when $\gamma=I_m$. The identities listed below were central to the derivation of the full-lattice resolvent operator in \cite[Lemma 3.3]{CGZ07}.
\begin{lemma}\label{l3.12}
Let $k,k_0\in{\mathbb{Z}}$ and $z\in{\C\backslash\{0\}}$. Then, for a unitary $\gamma\in{\mathbb{C}^{m\times m}}$, the following identities hold for the matrix-Laurent polynomial solutions of \eqref{3.22} defined in \eqref{3.56} and \eqref{3.57}:
\begin{align}
& P^{(\gamma)}_\pm(z,k,k_0)Q^{(\gamma)}_\pm(1/\overline{z},k,k_0)^* +
Q^{(\gamma)}_\pm(z,k,k_0)P^{(\gamma)}_\pm(1/\overline{z},k,k_0)^* = 2(-1)^{k+1}I_m,\label{3.89}
\\
& R^{(\gamma)}_\pm(z,k,k_0)S^{(\gamma)}_\pm(1/\overline{z},k,k_0)^* +
S^{(\gamma)}_\pm(z,k,k_0)R^{(\gamma)}_\pm(1/\overline{z},k,k_0)^* = 2(-1)^{k}I_m,
\\
& P^{(\gamma)}_\pm(z,k,k_0)S^{(\gamma)}_\pm(1/\overline{z},k,k_0)^* +
Q^{(\gamma)}_\pm(z,k,k_0)R^{(\gamma)}_\pm(1/\overline{z},k,k_0)^* = 0,
\\
& R^{(\gamma)}_\pm(z,k,k_0)Q^{(\gamma)}_\pm(1/\overline{z},k,k_0)^* +
S^{(\gamma)}_\pm(z,k,k_0)P^{(\gamma)}_\pm(1/\overline{z},k,k_0)^* = 0.\label{3.92}
\end{align}
\end{lemma}
\begin{proof}
For the case $k=k_0$, each of the equations \eqref{3.89}--\eqref{3.92} follows from \eqref{3.56} and \eqref{3.57}. The induction argument described in \cite[Lemma~3.2]{CGZ07} then suffices to establish \eqref{3.89}--\eqref{3.92} when $k\ne k_0$. As already noted, the proof in \cite[Lemma~3.2]{CGZ07} treats the special case when $\gamma=I_m$ for solutions defined by \eqref{3.56}. In all cases under consideration, the proof involves a number of cases all following a similar pattern. We outline one of these cases for a solution of \eqref{3.22} defined in \eqref{3.57}.
Suppose equations \eqref{3.89}--\eqref{3.92} hold when $k\in{\mathbb{Z}}$ is even.
Then utilizing \eqref{3.22} together with \eqref{3.8} and \eqref{3.9},
one computes
\begin{align}
&P^{(\gamma)}_-(z,k+1,k_0)Q^{(\gamma)}_-(1/\overline{z},k+1,k_0)^* +
Q^{(\gamma)}_-(z,k+1,k_0)P^{(\gamma)}_-(1/\overline{z},k+1,k_0)^* \notag
\\
&\hspace{5pt} = \widetilde\rho_{k+1}^{-1}\alpha_{k+1}
\big[P^{(\gamma)}_-(z,k,k_0)Q^{(\gamma)}_-(1/\overline{z},k,k_0)^* +
Q^{(\gamma)}_-(z,k,k_0)P^{(\gamma)}_-(1/\overline{z},k,k_0)^*\big] \alpha_{k+1}^*\widetilde\rho_{k+1}^{-1}
\notag
\\
&\hspace{15pt} + \widetilde\rho_{k+1}^{-1} \big[R^{(\gamma)}_-(z,k,k_0)S^{(\gamma)}_-(1/\overline{z},k,k_0)^*
+ S^{(\gamma)}_-(z,k,k_0)R^{(\gamma)}_-(1/\overline{z},k,k_0)^*\big] \widetilde\rho_{k+1}^{-1}\notag
\\
&\hspace{15pt} + z\widetilde\rho_{k+1}^{-1} \big[R^{(\gamma)}_-(z,k,k_0)Q^{(\gamma)}_-(1/\overline{z},k,k_0)^*
+ S^{(\gamma)}_-(z,k,k_0)P^{(\gamma)}_-(1/\overline{z},k,k_0)^*\big]
\alpha_{k_0}^*\widetilde\rho_{k+1}^{-1} \notag
\\
&\hspace{15pt} + \widetilde\rho_{k+1}^{-1}\alpha_{k_0}
\big[P^{(\gamma)}_-(z,k,k_0)S^{(\gamma)}_-(1/\overline{z},k,k_0)^* +
Q^{(\gamma)}_-(z,k,k_0)R^{(\gamma)}_-(1/\overline{z},k,k_0)^*\big] \widetilde\rho_{k+1}^{-1}z^{-1} \notag
\\
&\hspace{5pt} = 2(-1)^{k+1}
\big[\widetilde\rho_{k+1}^{-1}\alpha_{k+1}\alpha_{k+1}^*\widetilde\rho_{k+1}^{-1} -
\widetilde\rho_{k+1}^{-2}\big] = 2(-1)^{(k+1)+1}I_m.
\end{align}
Similarly, one checks all remaining cases at the point $k+1$. Then by inverting the matrix ${\mathbb{T}}(z,k)$ and utilizing \eqref{3.22} in the form
\begin{align}
\binom{P_-(z,k-1,k_0)}{R_-(z,k-1,k_0)} = {\mathbb{T}}(z,k)^{-1}
\binom{P_-(z,k,k_0)}{R_-(z,k,k_0)},
\end{align}
where
\begin{equation}
{\mathbb{T}}(z,k)^{-1}=
\begin{cases}
\begin{pmatrix}
-\rho^{-1}_k\alpha^*_k& z\rho^{-1}_k\\z^{-1}\widetilde\rho^{-1}_k&-\widetilde\rho^{-1}_k\alpha_k
\end{pmatrix}
&k\ odd,\\[15pt]
\begin{pmatrix}
-\widetilde\rho^{-1}_k\alpha_k&\widetilde\rho^{-1}_k\\ \rho^{-1}_k&-\rho^{-1}_k\alpha^*_k
\end{pmatrix}
&k\ even,
\end{cases}
\end{equation}
one verifies the equations \eqref{3.89}--\eqref{3.92} at the point $k-1$. Similarly,
one verifies \eqref{3.89}--\eqref{3.92} at the points $k+1$ and $k-1$
under the assumption that $k\in{\mathbb{Z}}$ is odd.
\end{proof}
The next lemma introduces the half-lattice resolvent operators for ${\mathbb{U}}_{\pm,k_0}^{(\gamma)}$; this appears to be a new result:
\begin{lemma}\label{l3.13}
Let $z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\})$ and fix $k_0\in{\mathbb{Z}}$. Then, for a unitary $\gamma\in{\mathbb{C}^{m\times m}}$, the
resolvent $({\mathbb{U}}_{\pm,k_0}^{(\gamma)}-zI)^{-1}$ for the unitary CMV operator ${\mathbb{U}}_{\pm,k_0}^{(\gamma)}$ on
$\ltm{[k_0,\pm\infty)\cap{\mathbb{Z}}}$ is given in terms of its matrix representation in the
standard basis of $\ltm{[k_0,\pm\infty)\cap{\mathbb{Z}}}$ by
\begin{align}
&({\mathbb{U}}_{+,k_0}^{(\gamma)}-zI)^{-1}=\frac1{2z}
\begin{cases}
-P^{(\gamma)}_+(z,k,k_0)\widehat U^{(\gamma)}_+(1/\bar z,k',k_0)^*,\\
\hspace{48pt} k<k'\ and\ k=k'\ odd,\\
\widehat U^{(\gamma)}_+( z,k,k_0)P^{(\gamma)}_+(1/\bar z,k',k_0)^*,\\
\hspace{33pt} k>k'\ and\ k=k'\ even,
\end{cases}
k, k'\in[k_0,\infty)\cap{\mathbb{Z}},\label{3.95}
\intertext{where $P^{(\gamma)}_+(z,k,k_0)$ is defined in \eqref{3.56}, and
$\widehat U^{(\gamma)}_+( z,k,k_0)$ is defined in \eqref{3.72};}
&({\mathbb{U}}_{-,k_0}^{(\gamma)}-zI)^{-1}=\frac1{2z}
\begin{cases}
-\widehat U^{(\gamma)}_-(z,k,k_0)P^{(\gamma)}_-(1/\bar z,k',k_0)^*,\\
\hspace{48pt} k<k'\ and\ k=k'\ odd,\\
P^{(\gamma)}_-(z,k,k_0)\widehat U^{(\gamma)}_-(1/\bar z,k',k_0)^*,\\
\hspace{33pt} k>k'\ and\ k=k'\ even,
\end{cases}
\hspace{-5pt}k, k'\in(-\infty,k_0]\cap{\mathbb{Z}},\label{3.96}
\end{align}
where $P^{(\gamma)}_-(z,k,k_0)$ is defined in \eqref{3.57}, and $\widehat U^{(\gamma)}_-( z,k,k_0)$ is defined in \eqref{3.72}.
\end{lemma}
\begin{proof}
We begin by noting that the following equations hold for $k\in{\mathbb{Z}}$:
\begin{align}
&R^{(\gamma)}_\pm(z,k,k_0)\widehat U^{(\gamma)}_\pm(1/\bar z,k,k_0)^* +
\widehat V^{(\gamma)}_\pm( z,k,k_0)P^{(\gamma)}_\pm(1/\bar z,k,k_0)^*=0,\label{3.98}\\
&P^{(\gamma)}_\pm(z,k,k_0)\widehat U^{(\gamma)}_\pm( 1/\bar z,k,k_0)^* +
\widehat U^{(\gamma)}_\pm( z,k,k_0)P^{(\gamma)}_\pm(1/\bar z,k,k_0)^* = 2(-1)^{k+1}I_m.\label{3.99}
\end{align}
These equations are a consequence of \eqref{3.72}, \eqref{3.73}, \eqref{3.89}, \eqref{3.92}, and the fact that
$m^{(\gamma)}_\pm(z,k_0)$ are matrix-valued Caratheodory functions and hence satisfy the property given in \eqref{A.7}.
For $k,k'\in[k_0,\infty)\cap{\mathbb{Z}}$, let $G^{(\gamma)}_+(z,k,k',k_0)$ be defined by
\begin{equation}
G^{(\gamma)}_+(z,k,k',k_0)=
\begin{cases}
-P^{(\gamma)}_+(z,k,k_0)\widehat U^{(\gamma)}_+(1/\bar z,k',k_0)^*,& k<k'\ and\ k=k'\ odd,\\
\widehat U^{(\gamma)}_+( z,k,k_0)P^{(\gamma)}_+(1/\bar z,k',k_0)^*,& k>k'\ and\ k=k'\ even.
\end{cases}
\end{equation}
Then, \eqref{3.95} is equivalent to showing that
\begin{equation}\label{3.101}
\big({\mathbb{U}}_{+,k_0}^{(\gamma)}-zI\big)G^{(\gamma)}_+(z,\cdot,k',k_0) = 2z\Delta_{k'},\quad k'\in[k_0,\infty)\cap{\mathbb{Z}}.
\end{equation}
Assume that $k'\in[k_0,\infty)\cap{\mathbb{Z}}$ is odd. Then, for $\ell \in([k_0,\infty)\cap{\mathbb{Z}})\backslash
\{k',k'+1\}$, note that
\begin{align}\label{3.102}
\big(\big({\mathbb{U}}_{+,k_0}^{(\gamma)}-zI\big)G^{(\gamma)}_+(z,\cdot,k',k_0)\big)(\ell)
= \big(\big({\mathbb{V}}_{+,k_0}^{(\gamma)}{\mathbb{W}}_{+,k_0}^{(\gamma)}-zI\big)G^{(\gamma)}_+(z,\cdot,k',k_0)\big)(\ell) =0.
\end{align}
Next, by \eqref{3.98}, \eqref{3.99}, note that
\begin{align}
&\begin{pmatrix}
\big(\big({\mathbb{U}}_{+,k_0}^{(\gamma)}-zI\big)G^{(\gamma)}_+(z,\cdot,k',k_0)\big)(k')\\
\big(\big({\mathbb{U}}_{+,k_0}^{(\gamma)}-zI\big)G^{(\gamma)}_+(z,\cdot,k',k_0)\big)(k'+1)
\end{pmatrix}\notag
\\
&\hspace{10pt}=
\begin{pmatrix}
\big(\big({\mathbb{V}}_{+,k_0}^{(\gamma)}{\mathbb{W}}_{+,k_0}^{(\gamma)}-zI\big)G^{(\gamma)}_+(z,\cdot,k',k_0)\big)(k')\\
\big(\big({\mathbb{V}}_{+,k_0}^{(\gamma)}{\mathbb{W}}_{+,k_0}^{(\gamma)}-zI\big)G^{(\gamma)}_+(z,\cdot,k',k_0)\big)(k'+1)
\end{pmatrix}\notag
\\
&\hspace{10pt}=\Theta(k'+1)
\begin{pmatrix}
-zR^{(\gamma)}_+(z,k',k_0)\widehat U^{(\gamma)}_+(1/\bar z,k',k_0)^*\\
z
\widehat V^{(\gamma)}_+( z,k'+1,k_0)P^{(\gamma)}_+(1/\bar z,k',k_0)^*
\end{pmatrix}
-z\begin{pmatrix}
G^{(\gamma)}_+(z,k',k',k_0))\\G^{(\gamma)}_+(z,k'+1,k',k_0))
\end{pmatrix}\notag
\\
&\hspace{10pt}=\Theta(k'+1)
\begin{pmatrix}
z\widehat V^{(\gamma)}_+( z,k',k_0)P^{(\gamma)}_+(1/\bar z,k',k_0)^*\\
z
\widehat V^{(\gamma)}_+( z,k'+1,k_0)P^{(\gamma)}_+(1/\bar z,k',k_0)^*
\end{pmatrix}
-z\begin{pmatrix}
G^{(\gamma)}_+(z,k',k',k_0))\\G^{(\gamma)}_+(z,k'+1,k',k_0))
\end{pmatrix}\notag
\\
&\hspace{10pt}=
z\begin{pmatrix}
\widehat U^{(\gamma)}_+( z,k,k_0)P^{(\gamma)}_+(1/\bar z,k,k_0)^* +
P^{(\gamma)}_+(z,k,k_0)\widehat U^{(\gamma)}_+( 1/\bar z,k,k_0)^*
\\
0
\end{pmatrix}\notag
\\
&\hspace{10pt}=
\begin{pmatrix}
2z(-1)^{k'+1}I_m\\0
\end{pmatrix}
=\begin{pmatrix}2zI_m\\0 \end{pmatrix}.\label{3.103}
\end{align}
Hence, when $k'\in[k_0,\infty)\cap{\mathbb{Z}}$ is odd, \eqref{3.101} is a consequence of \eqref{3.102} and \eqref{3.103}.
Assume that $k'\in[k_0,\infty)\cap{\mathbb{Z}}$ is even. Then, for $\ell\in ([k_0,\infty)\cap{\mathbb{Z}})\backslash
\{k'-1,k'\}$, note that \eqref{3.102} holds. Again, by \eqref{3.98} and \eqref{3.99}, note that
\begin{align}
&\begin{pmatrix}
\big(\big({\mathbb{U}}_{+,k_0}^{(\gamma)}-zI\big)G^{(\gamma)}_+(z,\cdot,k',k_0)\big)(k'-1)\\
\big(\big({\mathbb{U}}_{+,k_0}^{(\gamma)}-zI\big)G^{(\gamma)}_+(z,\cdot,k',k_0)\big)(k')
\end{pmatrix}\notag
\\
&\hspace{7pt}=
\begin{pmatrix}
\big(\big({\mathbb{V}}_{+,k_0}^{(\gamma)}{\mathbb{W}}_{+,k_0}^{(\gamma)}-zI\big)G^{(\gamma)}_+(z,\cdot,k',k_0)\big)(k'-1)\\
\big(\big({\mathbb{V}}_{+,k_0}^{(\gamma)}{\mathbb{W}}_{+,k_0}^{(\gamma)}-zI\big)G^{(\gamma)}_+(z,\cdot,k',k_0)\big)(k')
\end{pmatrix}\notag
\\
&\hspace{7pt}=\Theta(k')
\begin{pmatrix}
-zR^{(\gamma)}_+(z,k'-1,k_0)\widehat U^{(\gamma)}_+(1/\bar z,k'-1,k_0)^*\\
z
\widehat V^{(\gamma)}_+( z,k',k_0)P^{(\gamma)}_+(1/\bar z,k',k_0)^*
\end{pmatrix}
-z\begin{pmatrix}
G^{(\gamma)}_+(z,k'-1,k',k_0))\\G^{(\gamma)}_+(z,k',k',k_0))
\end{pmatrix}\notag
\\
&\hspace{7pt}=\Theta(k')
\begin{pmatrix}
-zR^{(\gamma)}_+(z,k'-1,k_0)\widehat U^{(\gamma)}_+(1/\bar z,k'-1,k_0)^*\\
-zR^{(\gamma)}_+(z,k',k_0)\widehat U^{(\gamma)}_+(1/\bar z,k',k_0)^*
\end{pmatrix}
-z\begin{pmatrix}
G^{(\gamma)}_+(z,k'-1,k',k_0))\\G^{(\gamma)}_+(z,k',k',k_0))
\end{pmatrix}\notag
\\
&\hspace{7pt}=
-z\begin{pmatrix}
0\\
P^{(\gamma)}_+(z,k,k_0)\widehat U^{(\gamma)}_+( 1/\bar z,k,k_0)^*+\widehat U^{(\gamma)}_+( z,k,k_0)P^{(\gamma)}_+(1/\bar z,k,k_0)^*
\end{pmatrix}\notag
\\
&\hspace{7pt}=
\begin{pmatrix}
0\\
2z(-1)^{k'+2}I_m
\end{pmatrix}
=\begin{pmatrix}0\\2zI_m \end{pmatrix}.\label{3.104}
\end{align}
Hence, when $k'\in[k_0,\infty)\cap{\mathbb{Z}}$ is even, \eqref{3.101} is a consequence of \eqref{3.102} and \eqref{3.104}.
The proof of \eqref{3.96} is omitted here for brevity, but follows a line of reasoning similar that just completed for the proof of \eqref{3.95} by using \eqref{3.72}, \eqref{3.73}, \eqref{3.98}, and \eqref{3.99}.
\end{proof}
Before stating our final result for the full-lattice resolvent of ${\mathbb{U}}$, let us recall the definition of, and some facts about, the matrix-valued Wronskian, defined in \cite{CGZ07}, for two ${\mathbb{C}^{m\times m}}$-valued sequences $U_j(z,\cdot)$, $j=1,2$. First, the Wronskian is defined for $k\in{\mathbb{Z}}, \; z\in{\C\backslash\{0\}},$ by
\begin{align}\label{3.105}
&W(U_1(1/\overline{z},k),U_2(z,k))\notag\\
&\hspace{10pt}= \frac{(-1)^{k+1}}{2} \big[
U_1(1/\overline{z},k)^*U_2(z,k)-({\mathbb{V}}^*U_1(1/\overline{z},\,\cdot\,))(k)^*
({\mathbb{V}}^*U_2(z,\,\cdot\,))(k)\big],
\end{align}
where ${\mathbb{V}}$ is defined in \eqref{3.15}. It is shown in \cite[Lemma 3.1]{CGZ07} when ${\mathbb{U}} U_j(z,\cdot)=zU_j(z,\cdot)$, and hence ${\mathbb{V}}^* U_j(z,\cdot)= V_j(z,\cdot)$, $j=1,2$, where ${\mathbb{U}}$ is viewed as a difference expression rather than as an operator acting on $\ell^2({\mathbb{Z}})^{m\times m}$, that the Wronskian of $U_j(z,\cdot)$, $j=1,2$, is independent of $k\in{\mathbb{Z}}$. Moreover, for $P_+^{(\gamma)}(z,\cdot,k_0)$ and $Q_+^{(\gamma)}(z,\cdot,k_0)$ defined in \eqref{3.56}, and for $U_\pm^{(\gamma)}(z,\cdot,k_0)$ defined in \eqref{3.74}, with $k,k_0\in{\mathbb{Z}}, \; z\in{\C\backslash\{0\}}$, as a consequence of \eqref{3.56}, \eqref{3.57}, \eqref{3.74}, \eqref{3.75}, and property \eqref{A.7}, we see that
\begin{align}
W\big(P_+^{(\gamma)}(1/\overline{z},k,k_0),Q_+^{(\gamma)}(z,k,k_0)\big) &= I_m,
\\
W\big(U_+^{(\gamma)}(1/\overline{z},k,k_0),U_-^{(\gamma)}(z,k,k_0)\big) &= M_-^{(\gamma)}(z,k_0)-M_+^{(\gamma)}(z,k_0).\label{3.107}
\end{align}
For notational simplicity, we abbreviate the Wronskian of $U_+^{(\gamma)}$ and
$U_-^{(\gamma)}$ by
\begin{equation}
W^{(\gamma)}(z,k_0)= - W\big(U_+^{(\gamma)}(1/\overline{z},k,k_0),U_-^{(\gamma)}(z,k,k_0)\big).
\end{equation}
Then, using \eqref{3.77}, \eqref{3.81}, and \eqref{3.107}, one analytically continues $W^{(\gamma)}(z,k_0)$ to $z=0$ and obtains
\begin{equation}\label{3.109}
W^{(\gamma)}(z,k_0) = M^{(\gamma)}_+(z,k_0)-M^{(\gamma)}_-(z,k_0), \quad k\in{\mathbb{Z}}, \; z\in{\mathbb{C}}.
\end{equation}
Moreover, one verifies the following symmetry property of the Wronskian $W^{(\gamma)}(z,k_0)$, for $k\in{\mathbb{Z}}, \; z\in\mathbb{C}$
\begin{equation}
M^{(\gamma)}_+(z,k_0)W^{(\gamma)}(z,k_0)^{-1}M^{(\gamma)}_-(z,k_0) =
M^{(\gamma)}_-(z,k_0)W^{(\gamma)}(z,k_0)^{-1}M^{(\gamma)}_+(z,k_0).
\end{equation}
Then, using \eqref{3.74}, \eqref{3.75}, \eqref{3.89}, \eqref{3.92}, \eqref{3.107}, \eqref{3.109}, and following the steps in the proof for \cite[Lemma 3.2]{CGZ07} for the special case when $\gamma=I_m$, we find that
\begin{align}
&U^{(\gamma)}_+(z,k,k_0){W^{(\gamma)}(z,k_0)}^{-1}U^{(\gamma)}_-(1/\overline{z},k,k_0)^*\notag\\
&\quad - U^{(\gamma)}_-(z,k,k_0)W^{(\gamma)}(z,k_0)^{-1}U^{(\gamma)}_+(1/\overline{z},k,k_0)^* = 2(-1)^{k+1}I_m,
\label{3.111}
\\
&V^{(\gamma)}_+(z,k,k_0)W^{(\gamma)}(z,k_0)^{-1}U^{(\gamma)}_-(1/\overline{z},k,k_0)^*\notag\\
&\quad -
V^{(\gamma)}_-(z,k,k_0)W^{(\gamma)}(z,k_0)^{-1}U^{(\gamma)}_+(1/\overline{z},k,k_0)^* = 0. \label{3.112}
\end{align}
Then, using \eqref{3.111} and \eqref{3.112}, and following the steps in the proof for \cite[Lemma 3.3]{CGZ07} for the special case when $\gamma=I_m$, we obtain the next result for the resolvent of the full-lattice operator ${\mathbb{U}}$.
\begin{lemma} \label{l3.14}
Let $z\in{\mathbb{C}}\backslash({\partial\hspace*{.2mm}\mathbb{D}}\cup\{0\})$, fix $k_0\in{\mathbb{Z}}$, and let $\gamma\in{\mathbb{C}^{m\times m}}$ be unitary. Then the
resolvent $({\mathbb{U}}-zI)^{-1}$ of the unitary CMV operator ${\mathbb{U}}$ on
$\ltm{{\mathbb{Z}}}$ is given, for $k,k' \in{\mathbb{Z}}$, in terms of its matrix representation in the
standard basis of $\ltm{{\mathbb{Z}}}$ by
\begin{align}
({\mathbb{U}}-zI)^{-1}(k,k') = \frac{1}{2z}
\begin{cases}
U^{(\gamma)}_-(z,k,k_0)W^{(\gamma)}(z,k_0)^{-1}U^{(\gamma)}_+(1/\overline{z},k',k_0)^*,\\
\hspace*{4.1cm} k < k' \text{ or } k = k' \text{ odd},
\\
U^{(\gamma)}_+(z,k,k_0)W^{(\gamma)}(z,k_0)^{-1}U^{(\gamma)}_-(1/\overline{z},k',k_0)^*,\\
\hspace*{3.92cm} k > k' \text{ or } k = k' \text{ even},
\end{cases} \label{3.113}
\end{align}
Moreover, since $0\in{\mathbb{C}}\backslash\sigma({\mathbb{U}})$, \eqref{3.113}
analytically extends to $z=0$.
\end{lemma}
|
\section{Introduction}\label{intro-scn}
The study of descriptors of geometrical objects encompassed by terms of \emph{form} or \emph{shape} be it for artisanry, biological and morphological interest or medical applications dates back to the very origins of mankind. It took, and
it still takes, however, until modern days to fully realize, that underlying most seemingly intuitive concepts of shape are rather counter-intuitive non-Euclidean geometries. Adding to counter-intuition, going from 2D to 3D shapes, these geometries cease to be well behaved. For this reason,
the statistical analysis of 3D shapes is highly challenging and has gained much less attention in theory and practice.
This work has been motivated by a joint research with the Institute for Forest Biometry and Informatics at the University of G\"ottingen studying the temporal evolution of the 3D shape of tree boles over their growing periods. The task tackled here
is to assess their growth towards and away from cylinders over time. This research is not only of interest in understanding fundamentals of biological growth and subsequent model building, cf. \cite{SH98,KoiHir06}. Also, the deviation from cylindricity
has a direct economical impact by reducing log volume and increasing the number of turns to reposition logs for commercial sawing processes, cf.
\cite{RBWE07}.
While Procrustes analysis is a well established tool for the statistical analysis of shape (e.g. \cite{DM98} and \cite{Dshapes} for numerical routines), to the knowledge of the author, for 3D there are no asymptotic results available. Rather there is the belief that Procrustes sample means are unqualified for inference even in 2D because they may be ``inconsistent estimators'' of the ``shape of the mean'' -- the latter being the mean of a \emph{perturbation model} -- unless the error is isotropically distributed, see \cite{Lele93}, \cite{KM97} as well as \cite{Le98}. This perturbation model introduced by \cite{G91} and
discussed in detail in Section \ref{pert_mods:scn} (cf. (\ref{perturbation_model:eq}) on page \pageref{perturbation_model:eq}), assumes error on the original configurations. The results cited above allowed inference based on Procrustes means only for a limited set of 2D scenarios, e.g. when randomness is caused by a procedure of acquiring landmarks independent of rotation and location. For more general settings as often occur in applications of biology, e.g. when randomness is also the result of non-uniform biological growth, however, Procrustes means could only contribute to data description while they were not considered for use in inferential statistics.
Such more general applications in mind, \cite{BP03,BP05} established an asymptotic theory for \emph{intrinsic} and \emph{extrinsic} means on manifolds allowing for
\emph{non-parametric} asymptotic inference. Here ``non-parametric'' stands for the general approach not silently relying on a mean given by a perturbation model in the configuration space. Curiously in 2D, Procrustes means coincide with extrinsic means, the latter being ``consistent'' in the statistical sense i.e. satisfying a Strong Law of Large Numbers,
see \cite{Z77}, \cite{Le98} as well as \cite{BP03}.
Due to the dominating paradigm of the perturbation model, it seems
that the notion of a Procrustes \emph{population} mean has not quite found its way into the community.
Recently, a perturbation model has also motivated a non-standard approach in the context of neuro-imaging by \cite{DKZL09}, employing Procrustes analysis for estimating a mean diffusion tensor.
Although ``perturbation consistency'' has not been established,
practical imaging results in particular for nearly degenerate tensors have been of very good quality.
It is the intention of this work to
\begin{enumerate}
\item[(a)] make 3D Procrustes means with their well behaved asymptotics available for the practioners in the field,
\item[(b)] to provide for an assessement of ``inconsistency'' with perturbation models on Kendall's shape spaces as well as for the approach of \cite{DKZ09} for diffusion tensor imaging, and
\item[(c)] to derive a framework to obtain the temporal shape motion relative to the shape of cylinders for a sample of Douglas fir tree stems.
\end{enumerate}
To this end, the following Section \ref{GPA:scn} reviews Procrustes analysis and related results. Section \ref{SLLN_CLT_PZ_mean:scn} places Procrustes means and a mean introduced by \cite{Z94} in the context of Fr\'echet means and gives the Central-Limit-Theorem (CLT) as the main theoretical result. Here, two aspects require special attention: first, a CLT can only hold on manifolds while in 3D, the underlying shape spaces cease to be that; secondly, since 3D Procrustes means are neither intrinsic nor extrinsic means, the CLT of \cite{BP05} is accordingly modified as detailed in the appendix.
In Section \ref{pert_mods:scn} it is clarified
that the reported ``(in)consistency'' of Procrustes means expresses ``(in)compatibility'' of a perturbation model with the canonical shape space geometry. Moreover for 3D, simulations based on the CLT show
that for isotropic errors the perturbation model
can be considered nearly compatible in most practical situations unless the shape of the perturbation mean is almost degenerate.
In contrast by simulation in Section \ref{midly:scn}, a similar perturbation model for mean diffusion tensors
is not compatible, not even for small isotropic errors. Curiously, however, with increased degeneracy of the perturbation mean considered, incompatibility seems not increasing. This is also the case for an extension of the model to \emph{mildly rank deficient} diffusion tensors, presented here.
Section \ref{Conical_boles:scn} introduces a framework for the assessment of shape of frusta cut from tree boles. It turns out that the shapes of cylinders form a geodesic in the shape space allowing for the computation of the distance of arbitrary frusta to the space of cylinders.
For statistical inference, Boostrap confidence intervals for this distance are simulated from a sample of reconstructed tree ring structures of small size. Comparison with two typical growth scenarios gives that
the bole shape of young trees with little competition grows uniformly or stronger towards cylinders, while the motion of shapes of old tree boles with heavy competition away from cylinders is similar to a growth maintaining cross-sectional ellipticality and tapering.
\section{Procrustes Analysis and Kendall's Shape Spaces}\label{GPA:scn}
In the statistical analysis of similarity shapes based on landmark configurations, geometrical $m$-dimensional objects (usually $m=2,3$) are studied by placing $k>m$ \emph{landmarks} at specific locations of each object. Each object is then described by a matrix in the space $M(m,k)$ of $m\times k$ matrices, each of the $k$ columns denoting an $m$-dimensional landmark vector. $\langle x,y\rangle := \tr(xy^T)$ denotes the usual inner product with norm $\|x\| = \sqrt{\langle x,x\rangle}$. For convenience and without loss of generality for the considerations below, only \emph{centered} configurations are considered. Centering in way treating all landmarks equally can be achieved by multiplying with a sub-Helmert matrix ${\cal H}$
\begin{eqnarray}\label{Helmert_matrix:def}{\cal H} &=& \left(\begin{array}{cccc}
\frac{1}{\sqrt{2}} &\frac{1}{\sqrt{6}} & \dots& \frac{1}{\sqrt{k(k-1)}} \\
-\frac{1}{\sqrt{2}} &\frac{1}{\sqrt{6}} & \dots& \frac{1}{\sqrt{k(k-1)}} \\
0&-\frac{2}{\sqrt{6}} &\dots& \frac{1}{\sqrt{k(k-1)}}\\
\vdots&\vdots&\ddots&\vdots \\
0&0 &\dots& -\frac{k-1}{\sqrt{k(k-1)}}
\end{array}\right) \in M(k,k-1)
\end{eqnarray}
from the right, yielding $x{\cal H}$ in $M(m,k-1)$. This corresponds to the relocating of objects in space such that their mean landmark is zero and isometrically projecting the linear sub space of matrices with vanishing mean landmark to the space of matrices with $k-1$ of landmarks. For this method and other centering methods cf. \citet[Chapter 2]{DM98}. Excluding also all matrices with all landmarks coinciding gives the space of \emph{configurations}
\begin{eqnarray*
F_m^k&:=& M(m,k-1) \setminus \{0\} \,.
\end{eqnarray*}
Since only the similarity shape is of concern, all configurations are considered modulo the group of similarity transformations $H = SO(m) \times \mathbb R_+$ (recall that we excluded w.l.o.g. translations) where the action is given by the
operation $(g,\lambda)x= g\lambda x.$
Here, $SO(m)$ denotes the special orthogonal group (the orientation preserving orthogonal transformations). The \emph{shape} of a configuration $x\in F_m^k$ is the orbit $[x] = \{hx:h\in H\}$, the \emph{shape space} is the quotient $F_m^k/H = \{[x]:x\in F_m^k\}$ with a suitable metric.
\paragraph{A na\"ive shape space.}
A straightforward metric structure for $F_m^k/H$ is given by the canonical quotient metric:
$$\inf_{h_1,h_2\in H} \|h_1x_1 -h_2x_2\|\mbox{ for }x_1,x_2\in F_m^k\,.$$
Unfortunately, due to the scaling action of $\mathbb R_+$, this metric is identically zero, a dead end for statistical ambition.
\paragraph{General Procrustes analysis (GPA).}
As a workaround, \cite{Gow} introduced a constraining condition which allowed for the definition of a mean. For a sample of configurations $x_1,\ldots,x_n\in F_m^k$ a \emph{full Procrustes sample mean} is given by the shape of
$$\frac{1}{n}\sum_{j=1}^n h^*_jx_j$$
where $h^*_1,\ldots,h^*_n$ are minimizers over $h_1,\ldots, h_n \in H$ of the \emph{Procrustes sum of squares}
\begin{eqnarray*
\sum_{i,j=1}^n\|h_jx_j -h_ix_i\|^2 \mbox{ under the constraining condition } \left\|\frac{1}{n}\sum_{j=1}^n h_jx_j\right\| ~=~ 1\,.
\end{eqnarray*}
Letting $\mu = \frac{1}{n}\sum_{j=1}^n h_jx_j$, $h_j = (g_j,\lambda_j)$, the Procrustes sum of squares is
$$ \sum_{i,j=1}^n\|h_jx_j -h_ix_i\|^2 = 2n\sum_{j=1}^n\|h_jx_j -\mu\|^2 = 2n\sum_{j=1}^n\|\lambda_jg_jx_j -\mu\|^2\,.$$
W.l.o.g. we may assume that all configurations are contained in the unit sphere
$$ S_m^k :=\{x\in M(m,k-1): \|x\|=1\}\,.$$
Then, any minimizing $\mu$ is the orthogonal projection of the mean of minimizing $h_1x_1,\ldots, h_nx_n$ to $S_m^k$. In consequence, minimization can be performed sequentially, first for the $\lambda_j$, then for the $g_j$ and finally for $\mu$, as noted. Partial differentiation in particular gives the minimizing $\lambda^*_j = \langle g_jx_j,\mu\rangle >0$ (unless $\langle gx_j,\mu\rangle = 0 $ for all $g\in SO(m)$) such that every $\mu^*$ having the shape of a full Procrustes sample mean is a minimizer of
$$ \min_{\mu\in S_m^k} \min_{\footnotesize\begin{array}{l}g_1,\ldots,g_n \in SO(m)\\\langle g_ix_i,y\rangle \geq 0,i=1,\ldots,n\end{array}} \sum_{j=1}^n \| \langle g_jx_j,\mu\rangle \,g_jx_j -\mu\|^2 = \min_{\mu\in S_m^k} \sum_{j=1}^n \delta(x_j,\mu)^2$$
with the \emph{residual distance}
\begin{eqnarray}\label{res_dist:def}
\delta(x,y) &:=& \min_{\footnotesize\begin{array}{l}g\in SO(m)\\\langle gx,y\rangle \geq 0\end{array}}\|\langle gx,y\rangle gx - y\|\,.
\end{eqnarray}
\paragraph{Kendall's shape spaces.}
\cite{K77} proposed a slightly different work\-around. Instead of considering the quotient w.r.t. the action of $\mathbb R_+$, he projected all configurations to $S_m^k$ called the \emph{pre-shape sphere} and considered the metric quotient w.r.t. $SO(m)$ only, which is called \emph{Kendall's shape space}
$$\Sigma_m^k := S_m^k/SO(m) = \{[x]:x\in S_m^k\}\mbox{ with the \emph{fiber} } [x] = \{gx:g\in SO(m)\}\,.$$
Kendall's shape space is thus a quotient of a sphere allowing for two non-trivial canonical quotient metrics
$$d_{\Sigma_m^k}([x],[y]) := \inf_{g,h\in SO(m)} d(gx,hy) = \inf_{g\in SO(m)} d(gx,y)\,,$$
$d$ denoting either the Euclidean distance on $F_m^k$ or the \emph{spherical} distance on $S_m^k$:
$$d^{(e)}_{F_m^k}(x,y) :=\|x-y\|,\quad d^{(s)}_{S_m^k}(x,y) :=2\arcsin\frac{\|x -y\|}{2}\,.$$
In some applications only the \emph{form}, i.e. shape without size filtered out is of interest. The corresponding \emph{size-and-shape space} is $S\Sigma_m^k := F_m^k/SO(m)$ with \emph{size-and-shape}
$[x] := \{gx: g\in SO(m)\}\in S\Sigma_m^k $ of $x\in F_m^k$.
For a sample $x_1,\ldots, x_n \in F_m^k$, \emph{partial Procrustes sample means}, the size-and-shapes of $\mu^*\in F_m^k$ are then considered which minimize
$$\min_{g_1,\ldots,g_n\in SO(m)} \sum_{j=1}^n\|\mu - g_j x_j\|$$
over $\mu \in F_m^k$. Obviously, $\mu^* =\frac{1}{n}\sum_{j=1}^n g_j^*x_j$ with minimizers $g_1^*,\ldots,g_n^* \in SO(m)$. These means are ``partial'' because no equivalence w.r.t. scaling is considered. In contrast, in the definition of full Procrustes means above equivalence under the full group of similarities is considered.
\paragraph{2D full Procrustes means are extrinsic means.}
For $m=2$, \cite{K84} observed that one may
identify $F_2^k$ with $\mathbb C^{k-1}\setminus \{0\}$ such that every landmark column corresponds to a complex number. This means in particular that $z\in \mathbb C^{k-1}$ is a complex row-vector. With the Hermitian conjugate $a^* = (\overline{a_{kj}})$ of a complex matrix $a=(a_{jk})$ the pre-shape sphere $S_2^k$ is identified with $\{z\in \mathbb C^{k-1}: zz^*=1\}$ on which $SO(2)$ identified with $S^1=\{\lambda \in\mathbb C: |\lambda|=1\}$ acts by complex scalar multiplication. Then the well known Hopf-Fibration gives $\Sigma_2^k=S_2^k/S^1=\mathbb CP^{k-2}$, the complex projective $(k-2)$-dimensional space. Moreover, denoting by $M(k-1,k-1,\mathbb C)$ all complex $(k-1)\times (k-1)$ matrices, the \emph{Veronese-Whitney embedding} is given by
\begin{eqnarray*}
\frak{v}:\Sigma_2^k &\to& \{a \in M(k-1,k-1,\mathbb C): a^*=a\}
,~~[z] ~\mapsto~ z^*z\,.
\end{eqnarray*}
If $Z\in \mathbb C^{k-1}$ is a random pre-shape, identifying $\frak{v}(\Sigma_2^k)$ with $\Sigma_2^k$, the set of \emph{extrinsic means} (cf. \cite{BP03}) of $[Z]$ is the set of shapes of the orthogonal projection of the usual expected value $\mathbb E(Z^*Z)$ to $\frak{v}(\Sigma^k_2)$. Employing complex linear algebra, the set of extrinsic means is easily identified as the
shapes of the eigenvectors to the largest eigenvalue of the \emph{complex integral of squares matrix} $\mathbb E(Z^*Z)$.
Since on the other hand
$$\mathbb E\big(\delta(Z,w)^2\big) = 1 - w \mathbb E(Z^*Z) w^*\,.$$
with the residual shape distance from (\ref{res_dist:def}), we have that that any eigenvector of $\mathbb E(Z^*Z)$ to its largest eigenvalue is a pre-shape of a full Procrustes mean.
\begin{Th} The set of 2D full Procrustes means is the inverse image under the Veronese-Whitney embedding of the set of extrinsic means on $\frak{v}(\Sigma_2^k)$.
\end{Th}
In most applications, the largest eigenvalue of $\mathbb E(Z^*Z)$ is simple, then the 2D full Procrustes mean is uniquely determined.
A similar procedure gives extrinsic means using the \emph{Schoenberg embedding} for the related space of Kendall's \emph{reflection shapes} of arbitrary dimension, not discussed here, cf. \citet{B08}.
Procrustes means for three- and higher-dimensional configurations, however, cannot be modeled in this vein.
\section{Asymptotics for Procrustes and Ziezold Means}\label{SLLN_CLT_PZ_mean:scn}
In this section we will see that 3D and higher-dimensional Procrustes means as well as a mean introduced by \cite{Z94} are special cases of Fr\'echet $\rho$-means. In the following, \emph{smooth} means at least twice continuously differentiable.
\subsection{Kendall's Higher-dimensional Shape Spaces}
As we have seen above, $\Sigma_2^k$ is a manifold, namely a complex projective space. Similarly, $S\Sigma_2^m$ can be given a manifold structure. For $m\geq 3$, however, $\Sigma_m^k$ and $S\Sigma_m^k$ cease to be manifolds, they only contain dense and open submanifolds
{\footnotesize $$\begin{array}{rclcrcl}
(\Sigma_m^k)^*&:=&(S_m^k)^*/SO(m)&\mbox{with}&(S_m^k)^*&:=&\{x\in S_m^k: \rank(x)\geq m-1\}\\
(S\Sigma_m^k)^*&:=&(F_m^k)^*/SO(m)&\mbox{with}&(F_m^k)^*&:=&\{x\in F_m^k: \rank(x)\geq m-1\}
\end{array}\,$$}called the \emph{manifold parts} of \emph{regular} or equivalently \emph{non-degenerate} shapes, size-and-shapes, pre-shapes and configurations, respectively. A pre-shape or configuration $x$ or its shape or its size-and-shape $[x]$ is called \emph{strictly regular} if $\rank(x)=m$. In case of $\Sigma_m^k$, $m\geq 3$, approaching degenerate shapes, some sectional curvatures are unbound. A detailed discussion can be found in \cite{KBCL99}
as well as \cite{HHM07}.
Note that the square of a distance may not be smooth at so called cut points. E.g the square of the spherical distance between $e^{it}$ and $e^{is}$ on the unit circle $S^1\subset \mathbb C$ is smooth except when $e^{i(t-s)} =-1$. In general on a Riemannian manifold, the \emph{cut locus} $C(p)$ of $p$ comprises all points $q$ such that the extension of a length minimizing geodesic joining $p$ with $q$ is no longer minimizing beyond $q$. If $q\in C(p)$ or $p\in C(q)$ then $p$ and $q$ are \emph{cut points}.
\begin{LemDef}\label{all_are_metrics:rm}
All of
{\footnotesize $$\begin{array}{rclcl}
d^{(i)}_{\Sigma_m^k}([x],[y]) &:=& \min_{g\in SO(m)}d^{(s)}_{S_m^k}(gx,y)\,, \\
d^{(p)}_{\Sigma_m^k}([x],[y]) &:=& \sin d^{(i)}_{\Sigma_m^k}([x],[y])&=&\min_{{\footnotesize\begin{array}{l}g\in SO(m)\\\langle gx,y\rangle \geq 0\end{array}}}\sqrt{1-\langle gx,y\rangle^2} \,,\\
d^{(z)}_{\Sigma_m^k}([x],[y]) &:=& 2\sin \frac{d^{(i)}_{\Sigma_m^k}([x],[y])}{2}&=&\min_{g\in SO(m)}\sqrt{2(1-\langle gx,y\rangle)} \,,\\
d^{(i)}_{S\Sigma_m^k}([x],[y]) &:=& \min_{g\in SO(m)}d^{(e)}_{F_m^k}(gx,y)&=&\min_{g\in SO(m)}\sqrt{\|x\|^2+\|y\|^2-2\langle gx,y\rangle}
\end{array}$$}\noindent
are metrics on the shape space and the size-and-shape space, respectively, and their squares are smooth on the respective manifold parts except at cut points.
\end{LemDef}
\begin{proof} If the canonical quotient $Q=M/G$ of a smooth Riemannian manifold $M$ due to the action of a Lie group $G=SO(m)$ on $M$ is a Riemannian manifold, then the intrinsic distance on $Q$ is a smooth metric (cf. \citet[Chapter 4.1]{AM78}). This gives the smoothness of $(d^{(i)}_{\cdot})^2$ on the respective manifold parts which are dense in the shape space and the size-and-shape space, respectively, except at cut points.
Hence, the $d^{(i)}_{\cdot}$ extend to metrics on these. Moreover, $d^{(z)}_{\Sigma_m^k}$ is a metric due to convexity and monotonicity of $t \to 2\sin (t/2)$ for $t\in [0,\pi]$; for $d^{(p)}_{\Sigma_m^k}$ we rely on \citet[p. 206]{KBCL99}.
\end{proof}
We say that $x$ \emph{is in optimal position to $y$} if $\max_{g\in SO(m)} \langle gx,y\rangle = \langle x,y\rangle$. In particular then $d^{(z)}_{\Sigma_m^k}([x],[y]) =\|x-y\|$ and $d^{(i)}_{S\Sigma_m^k}([x],[y]) =\|x-y\|$, respectively.
\subsection{Fr\'echet $\rho$-Means}\label{Frech_means:scn}
The Central-Limit-Theorem (CLT) derived below relies on the ``$\delta$-method'' for smooth transformations. In consequence, we can only expect a CLT theorem to hold on the manifold part of the shape space and the size-and-shape space, respectively. For the following we assume that $(Q,d)$ is a metric space and $P$ is a $D$-dimensional smooth Riemannian manifold. For example, $Q=\Sigma_m^k$ and $P = (\Sigma_m^k)^*$. Suppose that $X, X_1,X_2,\ldots$ are i.i.d. random variables mapping from an abstract probability space $(\Omega,\cal A,\mathbb P)$ to $Q$ equipped with its Borel $\sigma$-field. For a random vector $Y\in \mathbb R^D$, $\mathbb E(Y)$ denotes the usual expectation, if defined.
\begin{Def}\label{Frechet_means:def} For a continuous function $\rho:Q\times P \to [0,\infty)$ define the \emph{set of population Fr\'echet $\rho$-means of $X$ in $P$} by
$$ E^{(\rho)}(X) = \argmin_{\mu\in P} \mathbb E\big(\rho(X,\mu)^2\big)
\,.$$
For $\omega\in \Omega$ denote the \emph{set of sample Fr\'echet $\rho$-means} by
$$ E^{(\rho)}_n(\omega) = \argmin_{\mu\in P} \sum_{j=1}^n \rho\big(X_j(\omega),\mu\big)^2\,.$$
\end{Def}
Since their original definition by \cite{F48} for $P=Q$
and $\rho=d$, such means have found much interest. \cite{Z77} extended the concept to $\rho$ being a quasi-metric only.
On Riemannian manifolds ($Q=P$) w.r.t. the Riemannian metric $\rho$, \cite{BP03,BP05} introduced the corresponding means as \emph{intrinsic means}, and, taking $\rho$ to be the metric of an ambient Euclidean space, as \emph{extrinsic means}.
In consequence of Lemma and Def. \ref{all_are_metrics:rm} and the connection with the residual distance (\ref{res_dist:def}),
$d^{(p)}_{\Sigma_m^k}([x],[y])=\delta(x,y)$, we can at once identify Procrustes means.
\begin{Cor} The set of $\rho$-Fr\'echet sample means on the shape space and size-and-shape space, respectively, are the sets of
\begin{enumerate}
\item[(i)] full Procrustes sample means for $\rho = d^{(p)}_{\Sigma_m^k}$,
\item[(ii)] partial Procrustes sample means for $\rho = d^{(i)}_{S\Sigma_m^k}$.
\end{enumerate}
\end{Cor}
For $m=2$, \cite{Z94} introduced mean shapes w.r.t. $\rho = d^{(z)}_{\Sigma_m^k}$ which have been studied also by \cite{Le98}. The following first two definitions honor his memory. The other extend sample Procrustes means to the population case.
\begin{Def}
Call $d^{(z)}_{\Sigma_m^k}$ the \emph{Ziezold metric} on the shape space, and the Fr\'echet $d^{(z)}_{\Sigma_m^k}$-means the set of \emph{Ziezold means}. The Fr\'echet $d^{(p)}_{\Sigma_m^k}$-means are the set of \emph{full Procrustes means}, the Fr\'echet $d^{(i)}_{S\Sigma_m^k}$-means the set of \emph{partial Procrustes means}.
\end{Def}
Ziezold means and full Procrustes means, even though defined earlier, can be thought of as a generalization of extrinsic means and means for crude residuals (cf. \cite{MJ00}), respectively. Partial Procrustes means are extensions of intrinsic means to non-manifolds.
Let us now touch on the issues of existence and uniqueness for the above introduced means. Both are are rather simple issues for extrinsic means, since they are orthogonal projections of classical Euclidean means in an ambient space to the embedded manifold in question, this projection being well defined except for a set of Lebesgue measure zero (cf. \cite{BP03}). For non-extrinsic means as above, however, existence, namely that the means are assumed on the manifold part is a deeper issue tackled in \cite{H_meansmeans_10}. Intrinsic means are unique if the underlying random elements are sufficiently concentrated (cf. \cite{L01} for the elaborate proof). For higher-dimensional Ziezold and full Procrustes means, to the knowledge of the author, there are no uniqueness results available.
Since we are concerned with metrics,
the following is a consequence of the Strong Law by \cite{Z77}, cf. also \cite{BP03}.
\begin{Th}\label{SLLN:th} Suppose that $X$ is a random pre-shape in $S_m^k$ or a random configuration in $F_m^k$ with unique Ziezold, full Procrustes or partial Procrustes mean $[\mu]$.
Then every measurable selection $[\mu_n(\omega)]$ from the sets of sample Ziezold, full Procrustes or partial Procrustes means, respectively, is a strongly consistent estimator of $[\mu]$.
\end{Th}
For the following we need the condition that $[X]$ is bounded away from the cut locus of
the mean a.s. in the Ziezold or full/partial Procrustes
sense
\begin{eqnarray}\label{bdd_away:cond}
\exists \epsilon >0 &\st& d([X],[\mu])
d(C([\mu]),[\mu])-\epsilon \mbox{ a.s.}
\end{eqnarray}
where $[\mu]$ and $d$ are corresponding Ziezold, full Procrustes or partial Procrustes means and distances, respectively.
\subsection{The Central-Limit-Theorem}\label{CLT:scn}
The Central-Limit-Theorem for Fr\'echet $\rho$-means requires additional setup.
\begin{Def}\label{CLT:def} Let $P$ be a $D$-dimensional manifold. We say that a $P$-valued estimator $\mu_n(\omega)$ of $\mu \in P$ satisfies a \emph{Central-Limit-Theorem} (CLT), if in any local chart $(\phi,U)$ near $\mu = \phi^{-1}(0)$ there are a suitable $D\times D$ matrix $A_{\phi}$ and a Gaussian $D\times D$ matrix ${\cal G}_{\phi}$ with zero mean and semi-definite symmetric covariance matrix $\Sigma_{\phi}$ such that
$$ \sqrt{n}A_{\phi}\big(\phi(\mu_n) - \phi(\mu)\big)~\to~{\cal G}_{\phi}$$
in distribution as $n\to \infty$.
\end{Def}
In most applications $A_{\phi}$ is non-singular, then in consequence of the ``$\delta$-method'', for any other chart $(\phi',U)$ near $\mu = \phi'^{-1}(0)$ we have simply
$$A^{-1}_{\phi'}\Sigma_{\phi'} (A^{-1}_{\phi'})^T = J(\phi'\circ\phi^{-1})_0 A^{-1}_{\phi} \Sigma_{\phi}(A^{-1}_{\phi})^TJ(\phi'\circ\phi^{-1})_0^T\,$$
where $J(\cdot)_0$ denotes the Jacobian matrix of first derivatives at the origin.
In consequence of Theorem \ref{SLLN:th} and Theorem \ref{CLT:th} from the appendix (assertion (ii) follows from \cite{BP05}) we have the following.
\begin{Th}\label{CLT_PZ:th
Suppose that $X$ is a random pre-shape in $S_m^k$ or a random configuration in $F_m^k$, $[X]$ having a unique Ziezold, full Procrustes or partial Procrustes mean $[\mu]$ on the respective manifold part.
Then every measurable selection $[\mu_n(\omega)]$ from the sets of sample Ziezold, full Procrustes or partial Procrustes means, respectively, satisfies a Central-Limit-Theorem if $[X]$ is bounded away from the cut locus
of the mean a.s. in the Ziezold or full/partial Procrustes sense (\ref{bdd_away:cond}) under the following additional condition:
\begin{enumerate}
\item[(i)] none in case of Ziezold or full Procrustes means,
\item[(ii)] in case of partial Procrustes means, if the Euclidean second moment $\mathbb E(\|X\|^2)$ is finite.
\end{enumerate}\noindent
In a suitable chart $(\phi,U)$ the corresponding matrices from Definition \ref{CLT:def} are given by
$$ A_{\phi} = \mathbb E(H\rho([X],[\mu])),~~\Sigma_{\phi} = \cov(\grad\rho([X],[\mu]))$$
where $\rho$ denotes the distances $d_{\Sigma_m^k}^{(p)},d_{\Sigma_m^k}^{(z)}$ and $d_{S\Sigma_m^k}^{(i)}$, respectively. Moreover, $\grad$ and $H$ denote the gradient and Hessian of $x\mapsto \rho([X],[\phi^{-1}(x)])^2$, respectively.
\end{Th}
Numerical simulations show a rather good finite sample accuracy of the CLT, cf. Figure \ref{geod_pert_1_sample_test:fig}.
\begin{figure}[h!]
\begin{minipage}{0.4\textwidth}
\includegraphics[angle=-90,width=1\textwidth]{hor_pert_mod_density_1_sample_test.eps}
\end{minipage}
\begin{minipage}{0.05\textwidth}\hfill \end{minipage}
\begin{minipage}{0.55\textwidth}
\caption{\it Depicting the proximity to the standard normal density (dashed line) of the normalized density (solid line) of the non-zero coordinate of full Procrustes means of samples of size $n=10$ from the geodesic perturbation model (\ref{geod_perturbation_model:eq}) in $\Sigma_3^4$ with $\epsilon_i=0$ and $s_i \sim N(0,0.1^2)$ with the two-dimensional mean $\mu=\diag(1,-1,0)/\sqrt{2}{\cal H}^T$ and $\nu=\diag(1,1,1)/\sqrt{3}{\cal H}^T$.\label{geod_pert_1_sample_test:fig}}
\end{minipage}
\end{figure}
\begin{Rm}
For $m\geq 3$, computing Ziezold means (for an algorithm cf. \cite{Z94}) is computationally less costly than computing full Procrustes means (corresponding
algorithms
are discussed in \citet[Chapter 5.3]{DM98}): for full Procrustes means in every iteration step every single optimal positioned datum needs additionally to be projected to the tangent space. The simulations reported in Section \ref{isotropic_error:scn} and the Bootstrap simulations in Section \ref{Conical_boles:scn}
give similar results but are approximately $15 \,\%$ faster when using Ziezold means instead of full Procrustes means.
\end{Rm}
\begin{Rm}
\cite{Gr05} proves that the above algorithms converge under rather broad conditions and supplies error estimates.
\end{Rm}
\section{Perturbation Inconsistency for Kendall Shapes}\label{pert_mods:scn}
For brevity in this section unless otherwise specified, Procrustes means refer to full Procrustes means.
\cite{G91} proposed to model a sample of landmark configuration matrices $y_i$ ($i=1,\ldots,n$) with a \emph{perturbation model} that since then, has been highly popular in the community:
\begin{eqnarray}\label{perturbation_model:eq}
y_i &=& \lambda_i g_i(\mu + \epsilon_i) + t_i\,.
\end{eqnarray}
The $\lambda_i >0$ convey scaling, the $g_i$ are rotation matrices and the $t_i$ stand for translations. Obviously, these three sets of nuisance parameters keep the \emph{shape} invariant, of interest is only the deterministic \emph{perturbation mean} $\mu$ and the random perturbations $\epsilon_i$ with zero expectation. If the size is also of interest then set $\lambda_i=1$ keeping only the form invariant under rotation and translation. Under the assumption of isotropic Gaussian errors $\epsilon_i$,
for estimation and inference on the population perturbation means $\mu_i$ and error covariances $\epsilon_i$, \cite{G91} proposed to use \emph{sample Procrustes means} obtained from GPA, cf. Section \ref{GPA:scn}. In the sequel, the property that (for arbitrary errors) sample Procrustes means converge to the shape of the mean of an underlying perturbation model has been coined as the ``consistency of Procrustes means''.
\subsection{Isotropic 3D Error}\label{isotropic_error:scn}
In order to validate Goodall's proposal, \cite{KM97} studied the analog of the perturbation model (\ref{perturbation_model:eq}) on the pre-shape sphere
and showed that for 2D configurations with isotropic $\epsilon_i$, the Procrustes population mean from (\ref{perturbation_model:eq}) is identical with the shape of the perturbation mean.
\cite{Le98} extended these results and showed that under slightly relaxed conditions for 2D configurations, intrinsic, Ziezold and Procrustes means all agree with the shape of $\mu$.
For 3D shapes the above arguments are no longer valid because the shape-fibres in $S_m^k$ are not spanned by geodesics in general (see \citet[Example 5.1]{HHM07}), hence equality of Procrustes means with the shape of a perturbation model cannot be expected, even for very small isotropic error.
In order to assess the practical impact of this effect, we measure the distance between the shape of the perturbation mean and its corresponding Procrustes mean. In view of Theorem \ref{CLT_PZ:th}, this can be done by determining the distance between the shape of the perturbation mean and a corresponding sampled Procrustes mean while confidence or equivalently its accuracy can be estimated by distances of correspondingly sampled sample Procrustes means to their sample Procrustes mean. The following considerations detail this setup.
Consider a random $X = Y{\cal H}/\|Y{\cal H}\| \in S_m^k$ with $Y$ from a perturbation model (\ref{perturbation_model:eq}) such that the following means are unique. Assume $\mu \in S_m^k$ and denote by
\begin{description}
\item[$\nu$] the pre-shape of the Procrustes population mean
in optimal position to $\mu$;
\item[$\hat{\nu}^{(n)}$] the random pre-shape in optimal position to $\nu$ of the Procrustes sample mean
of $X_1,\ldots,X_n$ i.i.d. as $X$;
\item[$\hat{\nu}^{(n)}_j$] for $j=1,\ldots,N$ i.i.d. realizations of $\hat{\nu}^{(n)}$,
\item[$\hat{\nu}^{(n,N)}$] a realization of the Procrustes sample mean of $\hat{\nu}^{(n)}_1,\ldots,\hat{\nu}^{(n)}_N$ in optimal position to $\nu$.
\end{description}
Then in consequence of Theorem \ref{CLT_PZ:th},
\begin{eqnarray*}
\hat{\sigma}^{(n,N)}&:=&
\sqrt{\frac{1}{N}\sum_{j=1}^Nd^{(p)}_{\Sigma_m^k}([\hat{\nu}^{(n)}_j], [\hat{\nu}^{(n,N)}])^2}
~\approx~ \sqrt{\mathbb E\big(d^{(p)}_{\Sigma_m^k}([\hat{\nu}^{(n)}],[\nu])^2\big)}
\end{eqnarray*}
gives an approximation of the confidence into or accuracy of the measurement of $\nu$ by $\hat{\nu}^{(n)}$.
On the other hand, we have the approximation
$$\hat{d}^{(n,N)}:=\sqrt{\frac{1}{N}\sum_{j=1}^Nd^{(p)}_{\Sigma_m^k}([\nu_j^{(n)}],[\mu])^2}\approx \sqrt{\mathbb E\big(d^{(p)}_{\Sigma_m^k}(\nu^{(n)},\mu)^2\big)}\,.$$
The goodness of both approximations depends on $N$.
Moreover for fixed $N$, $\hat{\sigma}^{(n,N)}$ can be made arbitrarily small by choosing $n$ sufficiently large. If alongside, $\hat{d}^{(n,N)}$ also becomes equally small, this can be taken as strong evidence that the shape of the perturbation model agrees with the corresponding Procrustes mean. Otherwise, we have strong evidence, that the two disagree. Table \ref{pert_mod_3d:tab} reports values of $\hat{d}^{(n,N)}$ and $\hat{\sigma}^{(n,N)}$ for typical simulations of (\ref{perturbation_model:eq}) for $m=3,k=4$ with isotropic i.i.d Gaussian error $\epsilon$ of variance $\sigma^2$. The following remark summarizes the results of these and further simulations.
\begin{table}\centering
\begin{tabular}{c|c|cc}$\mu {\cal H}$&$\sigma$&$\hat{d}^{(n,N)}$&$\hat{\sigma}^{(n,N)}$\\\hline
\multirow{3}{*}{$\frac{1}{\sqrt{3}}\left(\begin{array}{ccc}1&0&0\\0&1&0\\0&0&1\end{array}\right)$}
&$0.5$&$ 0.011$&$ 0.0097$\\
&$0.1$&$0.0023 $&$ 0.0021$\\
&$0.01 $&$0.00022 $&$0.00020$\\ \hline
\multirow{3}{*}{$\frac{1}{\sqrt{1.1}}\left(\begin{array}{ccc}1&0&0\\0&0.3&0\\0&0&0.1\end{array}\right)$}
&$0.5$&$0.30$&$ 0.030$\\
&$0.1$&$0.016$&$ 0.0022$\\
&$0.01$&$ 0.00026$&$ 0.00020$\\ \hline
\multirow{3}{*}{$\frac{1}{\sqrt{1.001}}\left(\begin{array}{ccc}1&0&0\\0&0.01&0\\0&0&0\end{array}\right)$}&$0.1 $&$ 0.12$&$ 0.082$\\
&$0.01$&$0.0055$&$ 0.00044$\\
&$0.001$&$5.0e-05$&$ 2.0-05$
\end{tabular}
\caption{\it Estimated distance $\hat{d}^{(n,N)}$ between shape of the mean $\mu$ of the perturbation model (\ref{perturbation_model:eq}) and corresponding Procrustes mean with standard error $\hat{\sigma}^{(n,N)}$ for $n=10,000$ and $N=10$. The error follows independently ${\cal N}(0,\sigma^2)$ in every component. For convenience the Helmertized pre-shape $\mu {\cal H}$ is reported displaying proximity to degeneracy.\label{pert_mod_3d:tab}}
\end{table}
\begin{Rm}
In approximation the perturbation model (\ref{perturbation_model:eq}) with isotropic errors is compatible with the geometry of $\Sigma_3^k$ if the perturbation mean is far from degenerate. For nearly degenerate perturbation means, however, the perturbation model may be incompatible even for small isotropic errors. Using Ziezold means instead of Procrustes means gives similar results.
\end{Rm}
\subsection{Non-Isotropic 2D Error}\label{Kents_reserv:scn}
\paragraph{Global effects.}
Recall from Section \ref{GPA:scn}, modeling
$\Sigma_2^k$ by a complex projective space. Since the eigenvectors of the complex integral of squares matrix
may vary discontinuously under continuous matrix variation this led \cite{KM97} to the conclusion that Procrustes means can be inconsistent where in fact they are statistically consistent but possibly discontinuous.
\begin{figure}[h!]
\begin{minipage}{0.6\textwidth}
\includegraphics[angle=-90,width=1\textwidth]{Kents_Reservation.eps}
\end{minipage}
\begin{minipage}{0.35\textwidth}
\caption{\it Great circle in $\Sigma_2^3$ spanned by the shapes of the perturbation (\ref{Kent_pert:mod}). The little circles denote $100$ sampled shapes, the star their Procrustes sample mean and the triangle the respective Procrustes population mean. The respective headers give the standard deviation of the Gaussian $\epsilon$.
\label{Kents_Reservation:fig}}
\end{minipage}
\end{figure}
In view of the perturbation model (\ref{perturbation_model:eq}) consider in the spirit of \citet[p. 285]{KM97} the complex configuration $\mu = (\sqrt{2},0,1/\sqrt{2})\in \mathbb C^3$ additively perturbed by the non-isotropic random $\epsilon = (0,0,\sqrt{6}\eta t /2)\in \mathbb C^3$, $t\sim {\cal N}(0,1)$, $\eta>0$, modeling linear configurations with two fixed endpoints and a random landmark varying in the middle. The Kendall pre-shape of
\begin{eqnarray}\label{Kent_pert:mod}
\mu + \epsilon
\end{eqnarray}
is given by $z = \frac{1}{\sqrt{1+\eta^2t^2}}(1, \eta t)$ with the complex integral of squares matrix
$$ \left(\begin{array}{cc} \mathbb E\left(\frac{1}{1+\eta^2t^2}\right)&0\\
0& \eta^2 \mathbb E\left(\frac{t^2}{1+\eta^2t^2}\right)\end{array}\right)\,.$$
For small error intensity $\eta$, the shape of $\mu$ gives the Procrustes population mean. For higher error intensity, however, the Procrustes population mean indeed changes abruptly to the shape of the configuration $\nu=(\sqrt{2},\sqrt{2},-2\sqrt{2})$.
In order to visualize the situation in the shape space recall the Hopf fibration $$S_2^3 \to \Sigma_2^3 = \mathbb CP^1: (\alpha_1,\alpha_2)\mapsto \left(\re(\alpha_1\overline{\alpha_2}),\im(\alpha_1\overline{\alpha_2}), \frac{|\alpha_1|^2-|\alpha_2|^2}{2} \right)$$
mapping from a three-sphere to a two-sphere. The counter-intuitive behavior of discontinuity of the Procrustes mean is visualized in Figure \ref{Kents_Reservation:fig}. As is clearly visible, the discontinuity of the Procrustes mean in the error's standard deviation $\sqrt{6}\eta/2$ is due to the fact, that the corresponding shapes move along a common great circle in $\Sigma_2^3$, from clustering at the top (the shape of $\mu$) to clustering at the bottom (the shape of $\nu$). The discontinuity observed by \cite{KM97} thus reflects a global effect of an ``incompatibility'' of the perturbation model with the geometry of the shape space.
\paragraph{Local effects.}
As above,
\citet[Figure 3 on p. 598]{Lele93} considers an example of a distribution of planar but now quadrangular configurations along a straight line in the configuration space $F_2^4$. For simplicity of the argument consider $\mu = (i,1,-i,-1),~~\epsilon = (0,i,0,-i)$ and the sample $z_1=\mu +\epsilon,~~z_2=\mu-\epsilon\,$, $\epsilon >0$,
with mean $\mu$ in Euclidean $F_2^4$. One verifies immediately that $z_1$ and $z_2$ are not in optimal position, rather $\lambda = \frac{1}{\sqrt{5}}(1+2i)$ puts $z_1$ into optimal position $\lambda z_1$ to $z_2$ w.r.t. to the action of $SO(2)$. Then the form of $(\lambda z_1 + z_2)/2 $ is the partial Procrustes mean, which is different from the form of $\mu$. The alleged ``inconsistency'' of the partial Procrustes mean is now a local effect of an ``incompatibility'' of the perturbation model with the canonical geometry of the size-and-shape space $S\Sigma_2^4$.
\section{Asymptotics for Diffusion Tensors}\label{midly:scn}
In diffusion tensor neuro-imaging (for a short introduction into this young field, e.g. \cite{VilZhKL06}), the dominating eigenvector of a symmetric semi-positive definite (SPD) matrix from the space
$P(m) = \{0\neq a\in M(m,m): a^T=a\geq 0\}$
exhibits the dominating direction of molecular displacement due to a flow in tissue fibres of interest. The statistical and non-statistical literature to the task of reconstructing SPD matrices which are called \emph{diffusion tensors} (DTs) in this context is vast. Recently, matching techniques involving shape analysis (e.g. \cite{CMMWY06}) have gained momentum.
A non-standard statistical approach to reconstruct a mean DT from an observed sample of neighboring DTs based on Procrustes methods has been proposed by \cite{DKZL09}. One motivation comes from perturbation models, which, as it turns out below are ``inconsistent'' with Procrustes means. Coming as a surprise, however, practical image reconstructions based on this method in particular for nearly degenerate DTs, have produced results of convincingly good quality, cf. \cite{DKZ09}.
In the following, this approach is first extended to \emph{mildly rank-deficient DTs} and secondly, perturbation inconsistency is assessed.
Even though in most applications to imaging, the flow is observed in 2D or 3D, i.e. $m=2,3$, the following
applies to arbitrary $m\in \mathbb N$.
Diffusion tensors observed in the ``real world'' fall into the sub-space
$P^{+}(m)=\{a\in P(m): a>0\}$
of symmetric strictly positive definite matrices. This space -- even though much more complicated than complex projective space -- has been well studied (as the ``universal symmetric space of non-compact type with non-positive non-constant sectional curvatures'' e.g. \citet[Chapter XII]{L99}). Via the well known \emph{Cholesky factorization}
$$ {\rm chol} : P^{+}(m) \to UT^{+}(m)\,,$$
this space can equivalently be modelled by the group of $m$-dimensional upper triangular matrices $UT^{+}(m)$ with positive diagonal.
Traditionally diffusion tensors have been modelled within these spaces, statistical analysis has either been carried out by using the extrinsic distance inherited from $M(m,m)$ (e.g. \cite{PajBass03}) or the intrinsic distance due to the aforementioned structure (cf. \cite{fletch4}). Modeling flow in ideal micro-fibres, obviously leads to rank-deficient diffusion tensors, which, however, are not contained in these manifolds. Unless one utilizes a Euclidean embedding e.g. in the space of symmetric matrices, a new structure has to be found.
\subsection{A CLT for Mildly Rank-deficient Diffusion Tensors}
Here, an embedding $P(m)\hookrightarrow \Sigma_{m}^{m+1}$ is proposed, that maps the space $$P^*(m) = \{a\in P(m): \rank(a)\geq m-1\}$$
of \emph{mildly rank-deficient diffusion tensors} into the manifold part $(\Sigma_{m}^{m+1})^*$. This embedding is inspired by recent work of \cite{DKZ09} who, in effect, model $P^{+}(m)$ as a subspace of $\Sigma_{m}^{m+1}$. More subtly one could model as well using Kendall's \emph{reflection size-and-shape space}. In fact, modeling within size-and-shape space is equivalent to embedding reflection size-and-shape space in size-and-shape space.
To this end consider the following canonical domain for upper triangular matrices
{\footnotesize\begin{eqnarray*}
\lefteqn{UT(m) }\\&=& \Big\{a\in M(m,m): \mbox{ there is }1\leq i_0 \leq m\mbox{ such that for } \\
&&\left.\begin{array}{rl}
i=1:& \mbox{there is }m\geq j_{1} \geq 1 \mbox{ with }a_{ij_1} >0\mbox{ and } a_{il} = 0 \mbox{ for all }1\leq l < j_1,\\
2\leq i \leq i_0: &
\mbox{there is }m\geq j_{i} \geq j_{i-1}+1\mbox{ with } a_{il} = 0 \mbox{ for all }1\leq l<j_i,\mbox{ and }\\& \hspace{3.4cm}a_{ij_i} > 0\mbox{ in case } i < m,\\
i_0<i:
&a_{ij}=0\mbox{ for all } j=1,\ldots,m\end{array}
\right\}\,,
\end{eqnarray*}}the
sphere $SUT(m) = \{a\in UT(m): \|a\|=1\}$, $UT^\geq(m) = \{a\in UT(m): a_{mm}\geq 0\}$ and $SUT^\geq(m) = SUT(m)\cap UT^\geq(m)$. Moreover, with a bijective extension of the Cholesky factorization, the corresponding canonical projections from Section \ref{GPA:scn}:
$\pi$ and $s\pi$, $s(x) = \|x\|$ and bijections $\phi, \phi_s$, consider the following diagram of mappings:
\vspace{-0.2cm}
{\footnotesize\begin{eqnarray}\label{wild_P(m)_structure:def}
\left.\begin{array}{rclclcl}
P(m) \\
\updownarrow {\rm chol}\\
UT^\geq(m)&\hookrightarrow& SO(m)\times UT(m)&\stackrel{(g,a) \mapsto ga}{\rightarrow}&F_m^{m+1}&\stackrel{(\pi,s)}{\to}&\Sigma_m^{m+1}\times \mathbb R_+\\
&&&&\downarrow s\pi&& \updownarrow \phi \\
&&&&S\Sigma_{m}^{m+1}&& SUT^\geq(m)\\
&&&&\updownarrow \phi_s\\
&&&&UT^\geq(m)
\end{array}\right\}
\end{eqnarray}}In
particular, (\ref{wild_P(m)_structure:def}) gives rise to the following mappings
$$\tau_s:P(m) \to S\Sigma_m^{m+1}\mbox{ and } \tau:P(m) \to \Sigma_m^{m+1}\,.$$
\begin{Th}\label{diff_tens_kend_ss:thm} The mappings $\tau_s$ and $\tau$ are well defined, open and continuous. Moreover
\begin{enumerate}\item[(i)] $\tau_s$ restricted to the space of mildly rank-deficient diffusion tensors $P^*(m)$ is an injective mapping into the manifold part $(S\Sigma_m^{m+1})^*$ of Kendall's size-and-shape space. Restricted to the space of full rank diffusion tensors it is a diffeomorphism onto an open subset.
\item[(ii)] $\tau$ restricted to the space of mildly rank-deficient diffusion tensors $P^*(m)$ maps into the manifold part $(\Sigma_m^{m+1})^*$. Restricted to the space of full rank diffusion tensors it is a submersion of codimension $1$ onto an open subset.
\end{enumerate}
\end{Th}
For the proof of Theorem \ref{diff_tens_kend_ss:thm} and
diagram (\ref{wild_P(m)_structure:def}) we refer to the appendix.
\begin{Rm}
In consequence, inference for mildly deficient-rank diffusion tensors can be carried out via $\tau_s$ or $\tau$ utilizing the CLT: Theorem \ref{CLT_PZ:th}. Moreover, mean shapes can be pulled back under $(\tau_s)^{-1}$ to obtain \emph{mean diffusion tensors} if
they
stay away from the region corresponding to $a_{mm} <0$ in $UT(m)$.
\end{Rm}
\subsection{Perturbation Inconsistency for Diffusion Tensors}
A typical perturbation model considered by \cite{DKZ09}
is
\begin{eqnarray}\label{cov_matrix_pert:mod}x_i &=& (\mu + \epsilon_i)^T(\mu +\epsilon_i)\end{eqnarray}
with perturbation mean $\mu \in UT(m)$ and error $\epsilon_i$ following a Gaussian distribution ${\cal N}(0,\sigma^2)$
independently in every component or in every upper diagonal component and zero in every strictly lower diagonal component. In order to relate the quality of the embedding $\tau_s$ from (\ref{wild_P(m)_structure:def}) to other structures for the space of diffusion tensors (e.g. that of the universal symmetric space), \cite{DKZ09} compare the partial Procrustes distance between $\tau_s(\mu^T\mu)$ and partial Procrustes means from samples of (\ref{cov_matrix_pert:mod}).
For illustration we report in Table \ref{pert_mod_tens_3d:tab} a simulation in analogy to Section \ref{isotropic_error:scn} (using partial Procrustes means instead of full Procrustes means,
cf. Table \ref{pert_mod_3d:tab}) for three typical diffusion tensors for
upper triangular isotropic errors. The situation is similar for isotropic errors. The values of $\hat{d}^{(n,N)}$ for $\sigma=0.1$ in our Table \ref{pert_mod_tens_3d:tab} correspond to the ``RMSE($d_S$) of $\hat{\Sigma}_S$'' reported in the blocks labelled ``II'' in ``Table 2'' (corresponding to our second block) and ``Table 3'' (corresponding to our third block) of \cite{DKZ09}. Here, however, we used $n=10,000$ and identified these numbers as a measure only for perturbation inconsistency by additionally computing the standard error.
We remark the consequence.
\begin{table}[h!]\centering
\begin{tabular}{c|c|cc}$\mu$&$\sigma$&$\hat{d}^{(n,N)}$&$\hat{\sigma}^{(n,N)}$\\\hline
\multirow{3}{*}{$\left(\begin{array}{ccc}1&0&0\\0&1&0\\0&0&1\end{array}\right)$}
&$0.1$&$0.12$&$ 0.0017$\\
&$0.01$&$0.012$&$ 0.00017$\\
&$0.001 $&$0.0012$&$1.6e-05$\\ \hline
\multirow{3}{*}{$\left(\begin{array}{ccc}1&0&0\\0&0.3&0\\0&0&0.1\end{array}\right)$}
&$0.1$&$0.023$&$0.0020$\\
&$0.01$&$0.00026$&$0.00023$\\
&$0.001$&$2.4e-05$&$ 2.5e-05$\\ \hline
\multirow{3}{*}{$\left(\begin{array}{ccc}1&0&0\\0&0.01&0\\0&0&0\end{array}\right)$}
&$0.1 $&$ 0.14$&$ 0.0019$\\
&$0.01$&$0.0085$&$ 0.00021$\\
&$0.001$&$0.00081$&$ 2.2-05$
\end{tabular}
\caption{\it Estimated distance $\hat{d}^{(n,N)}$ between size-and-shape of $\mu$ from the perturbation model (\ref{cov_matrix_pert:mod}) and corresponding partial Procrustes mean with standard error $\hat{\sigma}^{(n,N)}$ for $n=10,000$ and $N=10$. The error follows independently ${\cal N}(0,\sigma^2)$ in every upper triangular component.\label{pert_mod_tens_3d:tab}}
\end{table}
\begin{Rm}\label{DTI-inconsistency:rm}
Suppose that $X$ follows a perturbation model (\ref{cov_matrix_pert:mod}). For upper triangular perturbation mean $\mu$ and error $\epsilon$ this is an anisotropic perturbation model for $S\Sigma_m^{m+1}$. For $\epsilon$ independent Gaussian in every component, $\mathbb E\big({\rm chol}(X)\big)\neq {\rm chol}(\mu)$ in general (cf. Lemma \ref{chol:lem} in the appendix).
Simulations corroborate inconsistency
with
the geometry of $S\Sigma_3^k$.
Surprisingly it seems that incompatibility is not increasing, near degeneracy. This observation may be taken as an explanation for successful modeling of nearly rank deficient diffusion tensors via Kendall's size-and-shape spaces and certainly deserves further research.
\end{Rm}
\section{Assessing Tree Bole Cylindricity
}\label{Conical_boles:scn}
We conclude with an application from forest biometry.
As basic descriptors for the shape of tree boles,
\emph{taper curves}
relate height above ground level with the area of a cross section at that height.
The shape of a tree stem is thus described by a
two-dimensional curve. Empirical curves can be directly analyzed with methods of shape analysis, e.g. \cite{Kr02} or, sophisticated models based on biological and elastomechanical context can be sought for, e.g. \cite{CS94} or \cite{GaSlobMats98}. For a discussion of current taper curve models cf. \cite{LiWeis10}.
While taper curves, following Cavalieri's principle, model tree stems by \emph{circular frusta}, recently bole ellipticality has been studied more closely, e.g. \cite{SH98} or \cite{KoiHir06}. In particular, \cite{RBWE07} model 3D logs by \emph{elliptical frusta} and discuss the impact of their findings on commercial logging: proximity to circular frusta increases the volume produced and reduces the number of turns to reposition logs in sawing machines and thus the total sawing time.
\begin{figure}[htb!]
\includegraphics[width=1\textwidth]{hex-cones8.36.eps}
\vspace{-0.5cm}
\includegraphics[width=1\textwidth]{hex-cones30.36.eps}
\vspace{-0.5cm}
\includegraphics[width=1\textwidth]{hex-cones62.36.eps}
\vspace{-0.5cm}
\caption{\it Distorted view (the center frustum in the bottom row has a diameter of approx. $50$ cm) along 36 angles of 1 m butt logs of five Douglas firs at ages 8 years (top row), 30 years (middle row) and 62 years (bottom row).\label{hex-cones:fig}}
\end{figure}
\subsection{Data Description}
Usually, tree stems are divided into three different parts: a short neiloid bottom part with strong tapering connecting with the root system, the main bole with little tapering -- which is of prime commercial interest -- and a conical top (e.g. \cite{LiWeis10}). Typically a \emph{butt log} that is a frustum taken above \emph{breast height} (1.3 m) is used to assess bole quality. For our application, we use 1 m butt logs from five Douglas fir trees typical for the inside of a small experimental stand in the Netherlands as detailed by \cite{gaslob01}. At about the age of 10 to 15 years, tree crowns met; subsequently with almost no thinning, competition for light increased strongly. The entire ring structure of bottom and top disk has been elaborately reconstructed along 36 equally spaced angles allowing to reconstruct the butt logs for every age beginning from 8 years to 62 years as displayed in Figure \ref{hex-cones:fig} for early, intermediate and ultimate age.
\subsection{Elliptical-Like Frusta: A Totally Geodesic Subspace}
In order assess the deviation of a 1 m frustum from a cylinder (of unknown dimensions), we compute the shape distance to the space of all cylinders. To this end, and in order to compare with a model building on elliptical frusta, we introduce a suitable subspace of configurations and shapes.
For given ${\bf a}^T,{\bf b}^T,{\bf 1}^T=(1,\ldots,1)^T \in \mathbb R^{\kappa}$ with
\begin{eqnarray}\label{parallel_cone_section:cond}
\langle {\bf a},{\bf b}\rangle =0 =\langle {\bf a},{\bf 1}\rangle = \langle {\bf b},{\bf 1}\rangle\,
\end{eqnarray}
and $r,\alpha,t>0$ let
\begin{eqnarray*
w_{\alpha,\beta,r,t} &:= &\left(\begin{array}{cc} r\,{\bf a}&r t \,{\bf a}\\
r\alpha\,{\bf b}&r \beta t\, {\bf b}\\
0\, {\bf 1}&\,{\bf 1}
\end{array}\right)\,
\end{eqnarray*}
be the configuration of an \emph{elliptical-like} unit height frustum. In case of ${\bf a} = \big(\cos(2\pi/\kappa), \ldots, \cos(2\pi\kappa/\kappa)\big)$ and ${\bf b} = \big(\sin(2\pi/\kappa), \ldots, \sin(2\pi\kappa/\kappa)\big)$ we have an \emph{elliptical} frustum of mean radius $r$, tapering $t$ with bottom half axes of length $r,r\alpha$ and top half axes of length $tr,tr\beta$. A \emph{straight frustum} has $t=1$, a \emph{circular frustum} has $\alpha = 1=\beta$ and a \emph{cylinder} is a straight circular frustum. On the grounds of the findings of \cite{CS94}, that for a large middle part of the bole, there is little shape variation when moving upward, a possible torsion between top and bottom ellipse is neglected in this model.
Denote by
$$ PS_{{\bf a},{\bf b}} := \left\{ \frac{w}{\|w\|} : w \in P_{{\bf a},{\bf b}} \right\}\subset S_3^{2\kappa}$$
the pre-shapes of all \emph{elliptical-like frusta} determined by ${\bf a},{\bf b}\in \mathbb R^\kappa$ satisfying (\ref{parallel_cone_section:cond}). Here
$$ P_{{\bf a},{\bf b}} := \{ w_{\alpha,\beta,r,t}{\cal H}: \alpha,\beta,r,t>0\}\subset F_3^{2\kappa}$$
and ${\cal H}$ denotes the Helmert sub-matrix from (\ref{Helmert_matrix:def}) in Section \ref{GPA:scn}.
Recall that a submanifold $P\subset M$ is \emph{totally geodesic} if for any two points in $P$ any minimal geodesic segment joining the two in $M$ is contained in $P$.
\begin{Th}\label{parallel_cone_scn_geodesic:thm} Consider ${\bf a}^T,{\bf b}^T\in \mathbb R^{\kappa}$ satisfying (\ref{parallel_cone_section:cond}). Then the shapes of all elliptical-like frusta form
a totally geodesic submanifold of $(\Sigma_3^{2\kappa})^*$ with horizontal lift $PS_{{\bf a},{\bf b}}$ to $S_3^{2\kappa}$.
\end{Th}
The proof of Theorem \ref{parallel_cone_scn_geodesic:thm}
is found in the appendix. We have at once:
\begin{Cor}\label{ideal_circular_cylinders:col}
The shapes of cylinders form a segment on a geodesic in $(\Sigma_3^{2\kappa})^*$.
\end{Cor}
\subsection{Data Analysis}
The distance of the shape of a frustum to the geodesic spanned by the shapes of cylinders is computed via a method proposed in \cite{HHM07}. Since for every age considered there are only five frusta, a $95 \,\% $ confidence band for the distance of the full Procrustes mean to the geodesic of cylinders has been computed by $200$ Bootstrap resamples. As clearly visible in Figure \ref{conf_dist_ideal_circ_geod:fig}, young frusta until the age of approx. 15 years tend to toward cylinders. At later ages they tend away again. The change point at approx. 15 years can be explained by the fact that at this time
competition for light began.
Note that
uniform growth naturally decreases tapering and ellipticality. In order to compare the initial growth to uniform growth, three curves of elliptical frusta
starting with $t$ and $\alpha=\beta$ in the range of corresponding estimated butt log values as well as
with size growth identical to the mean size growth are depicted in the left display of Figure \ref{conf_dist_ideal_circ_geod:fig} ($95 \,\%$ of estimated tapering coefficients $t$ observed are between $0.864 $ and $0.986$; a crude estimate for $\alpha$ and $\beta$ gives $95 \,\%$ of observed values between $1.02$ and $1.13$ (after rotating such that $\alpha,\beta\geq 1$)).
This comparison yields that observed initial growth towards cylinders appears even stronger than uniform.
In order to assess the growth of the older logs, compare it to the growth of elliptical frusta keeping tapering and ellipticality constant while letting size grow with the mean size of the original data. Three of such curves are depicted in the right display of Figure \ref{conf_dist_ideal_circ_geod:fig}. The observed movement away from the geodesics of cylinders appears rather similar.
\begin{figure}[h!]
\includegraphics[angle=-90,width=0.5\textwidth]{boundaries_confidence_band_uniform-bw.eps}
\includegraphics[angle=-90,width=0.5\textwidth]{boundaries_confidence_band_circular_mixed-bw.eps}
\caption{\it Solid: confidence band for the distance of the Procrustes mean shape of 1 m butt logs above breast height to the geodesic of cylinders for each year based on 200 boostrapped Procrustes means. For comparison dashed: uniform equalsize growth of elliptical cylinders (left display) starting with $(\alpha=\beta,t) = (1.07,0.93), (1.1,0.9), (1.15,0.85) $ from left to right; equalsize growth of elliptical cylinders (right display) with constant $(\alpha=\beta,t) = (1.025, 0.975), (1.05,0.95), (1.1,0.9)$ from left to right. \label{conf_dist_ideal_circ_geod:fig}}
\end{figure}
Summarizing, this study indicates that tree boles of young Douglas fir trees with little competition grow uniformly or even stronger towards cylinders. Older tree boles their crowns competing for light, grow as if they keep ellipticality and tapering constant. These findings
certainly call for more elaborate research, e.g. tapering and ellipticality can be investigated by developing a method to study distance and projection to the submanifold of elliptical-like frusta.
\section{Conclusion and Outlook}
In this paper for the statistical analysis of 3D shape, an asymptotic result for mean shape has been derived, a classical perturbation model and a newly proposed perturbation model for the statistical analysis of diffusion tensors has been revisited, and inference on the temporal deviation of the 3D shape of tree boles from cylinders has been performed.
Although Procrustes means are very popular, it seems that due to a misunderstanding, the notion of 3D \emph{Procrustes population means} had not been quite available in the community. Instead, population means of a perturbation model
had been estimated by \emph{Procrustes sample means} which were, as was well known, ``consistent'' estimators under isotropic 2D errors. Introducing Fr\'echet $\rho$-means in this work and extending the available Central-Limit Theorem
to underlying non-metrical distances, in particular, the issue of ``consistency'' has been identified as an ``incompatibility'' of the perturbation model with the shape space's geometry. Moreover, we recalled a Fr\'echet $\rho$-mean introduced by \cite{Z94}
the computation of which
in 3D is computationally slightly less costly than the computation of Procrustes means. This result allows for one-sample tests for a population Procrustes or Ziezold mean on the manifold part of the shape space, since due to strong consistency, sample Procrustes or Ziezold means will eventually come to lie on the manifold part a.s. For a two-sample test, one would need to ensure that Procrustes and Ziezold means are contained on the manifold part, if the underlying random shapes are a.s. contained in the manifold part. Settling this issue is the subject of a separate research, cf. \cite{H_meansmeans_10}.
While many means, different from classical intrinsic, extrinsic or Procrustes means are Fr\'echet $\rho$-means (e.g. geometric medians of \cite{fletch9} and (penalized) weighted Procrustes means of \cite{DKZL09}) it would be interesting to verify whether a larger class of means, e.g. the semi-metrical mean introduced in \cite{STCB07b} and successfully employed in computer vision also fall into this setup.
Within the discussion of ``consistency'', the modeling of \cite{DKZ09} for diffusion tensor imaging has been extended
to diffusion tensors mildly deficient in rank. To the knowledge of the author, this is the first framework allowing for statistical inference on non-regular diffusion tensors without utilizing a Euclidean embedding, say, in the space of symmetric matrices. Of course in many applications, the interest lies specifically in degenerate tensors because these indicate a strong directional flow. The finding that mildly rank-deficient diffusion tensors are only ``mildly perturbation inconsistent'' may provide for another motivation for the approach of \cite{DKZ09,DKZL09}. Following \cite{DKZ09}, we have used Procrustes means. For the underlying reflection shape space, extrinsic Schoenberg means qualify as well, they compute much faster than Procrustes or Ziezold means. As the most important advantage, the two-sample tests of \cite{B08} can then be performed even including rank 1 diffusion tensors. A drawback, however, may result from a possible insensitivity of Schoenberg means to degeneracy thus yielding a smaller discrimination power (cf. \cite{H_meansmeans_10} for a detailed discussion). These issues certainly warrant further research.
Briefly compiling the findings on tree boles we can add to the well known fact that
longest boles can be found in dense stands,
it seems that the most cylindrical boles can be found in young stands with low competition.
~\\~\\ \noindent{\bf
A geodesic perturbation model.}
As demonstrated in this work, for most practical applications in 3D avoiding degenerate configurations, perturbation models may be considered compatible with the shape space's geometry for isotropic error. Note that the shapes of entire tree boles are nearly singular (cf. \cite{HHM07}). For short bole frusta, however, such models may still serve intuition and for sufficiently small error provide an adequate approximation. As one of such consider the \emph{geodesic perturbation model}
\begin{eqnarray}\label{geod_perturbation_model:eq}
x_i &=& g_i\lambda_i(\gamma(s_i) + \epsilon_i
\end{eqnarray}
with a straight line $s \to \gamma(s) = \mu + s\nu$ in configuration space $F_m^k$ such that locally all points on $\gamma$ are in optimal position to each other.
For 3D frusta,
the geodesic of cylinders has been used here, any other geodesic of specific frusta can be used similarly. Moreover, based on larger samples, a geodesic perturbation model within the space of elliptical-like frusta can be used to assess specific model parameters. In particular, a close investigation of such model parameters may lead to inference on the impact of environmental effects on the shape of tree boles building on artificially induced modifications of biological tree parameters as in \cite{gaslob04}.
\vspace{0.4cm}
\noindent\textbf{Acknowledgments:}
The author would like to express his sincerest gratitude to the late Herbert Ziezold (1942 -- 2008) whose untimely death was most unfortunate. He would also like to thank Dieter Gaffrey and the colleagues from the Institute for Forest Biometry and Informatics at the University of G\"ottingen for discussion of and supplying with the tree-bole data. Also he is grateful for helpful discussions with Thomas Hotz, John Kent and Axel Munk.
|
\section{Introduction}
The doping evolution of the electronic properties of lightly-doped cuprates have attracted considerable interest because of their fundamental importance in the context of both physics near Mott transition and mechanism of superconductivity..
In the cuprates, in particular, the phase competition and the nature of the competing phases in the underdoped region have provided major experimental and theoretical challenges.
Angle-resolved photoemission spectroscopy (ARPES) studies of lightly-doped La$_{2-x}$Sr$_{x}$CuO$_{4}$ (LSCO)~\cite{Yoshida_LSCO} have shown that a sharp dispersive quasi-particle (QP) peak crosses the Fermi level ($E_{\rm F}$) in the nodal $\vec{k}$ = (0,0)-($\pi$,$\pi$) direction, which explains their metallic behavior~\cite{LSCO_metallic} and that the other part of the Fermi surface around ($\pi$,0) is gapped or pseudo-gapped, resulting in an ``arc'' of the Fermi surface around the nodal direction.
A remnant of the lower Hubbard band (LHB) (or the polaronic side band~\cite{Roesch}) persists at $\sim -$(0.4-0.6) eV as the QP peak appears near $E_{\rm F}$.
As holes are further doped, spectral weight is transferred from the ``LHB'' to the QP peak~\cite{LSCO_mu}.
ARPES studies on lightly-doped Ca$_{2-x}$Na$_x$CuO$_2$Cl$_2$ (Na-CCOC)~\cite{kfCCOC} and Bi$_2$Sr$_{2-x}$La$_x$CuO$_{6+\delta}$ (Bi2201)~\cite{Bi2201_mu}, on the other hand, have shown quite different behaviors:
With hole doping, the ``LHB'' feature is shifted toward $E_{\rm F}$ and merges into the QP band crossing $E_{\rm F}$, making the LHB-to-QP spectral weight transfer less obvious.
These observations are consistent with the pinning of the chemical potential in LSCO and the rapid chemical potential shift in Na-CCOC, Bi2201 and Bi2212 as implied by core-level x-ray photoemission spectroscopy (XPS)~\cite{LSCO_mu,Bi2201_mu,Bi2212_mu,CCOC_mu}.
Recent in-plane transport studies of YBa$_2$Cu$_3$O$_y$ (YBCO) have revealed their metallic behavior ($d\rho/dT>0$) at high temperatures even in the lightly-doped ``insulating'' region as in the case of LSCO~\cite{Ando_transport}.
The $E_{\rm F}$ crossing of QP peak in the nodal direction has been observed in lightly-doped~\cite{YBCO_nodalQP} as well as in underdoped to overdoped YBCO~\cite{YBCO_bilayer}.
It is therefore important to see whether the electronic structure of lightly-doped YBCO is similar to that of LSCO or to that of Na-CCOC, Bi2201, and Bi2212.
ARPES studies of YBCO have not been advanced compared to those of other high-$T_c$ cuprates because intense surface-state signals near $E_{\rm F}$ mask the bulk electronic states around the X [=($\pi$,0)] and Y [=(0,$\pi$)] points in the Brillouin zone.
Lu \textit{et al.}~\cite{ARPES_Lu} have made high-resolution ARPES measurments on YBa$_2$Cu$_3$O$_{6.993}$ and have succeeded in distinguishing bulk electronic states from the surface states around X and Y by aging the sample surfaces and deduced the superconducting gap.
They have reported that the superconducting peak is accompanied by a peak-dip hump structure, similar to Bi2212.
Yoshida \textit{et al.}~\cite{YBCO_nodalQP} and Borisenko \textit{et al.}~\cite{YBCO_bilayer} have reported a bilayer splitting in lightly-doped and optimally-doped YBCO, respectively.
In a recent ARPES study of optimally-doped YBCO by Nakayama \textit{et al.}~\cite{YBCO_Takahashi}, bulk and surface Fermi surfaces have been separated. More recently, Hossain \textit{et al.} reported the doping evolution of the electronic structure of the topmost CuO$_2$ plane from overdoped to the underdoped region~\cite{YBCO_Hossain}.
The recent Hall resistance measurment of underdoped YBCO revealed the existence of an small electron pocket~\cite{QO1_YBCO,QO2_YBCO}, although in the ARPES studies of other high-$T_{\rm c}$ cuprates, only a Fermi arc around the nodal direction has been observed.
Therfore, it is important to clarify whether an small electron pocket is observed in ARPES studies of underdoped YBCO or not.
In this paper, ARPES and core-level XPS studies of lightly-doped to underdoped YBCO are presented to clarify the above issues and to gain further insight into the electronic structure of YBCO.
\begin{table}
\caption{Oxygen content $y$, hole concentration $\delta$ in the CuO$_2$ plane estimated from the electrical resistivity
and thermopower~\cite{Ong&Uchida}, and $T_c$ of the YBCO samples studied in the present work.}
\label{tab:table1}
\begin{tabular}{p{2.5cm}|p{2.5cm}|p{2.5cm}}
\hline
\hline
$y$ & $\delta$ & $T_c$ (K) \\
\hline
6.20 & 0.01 & \\
6.25 & 0.02 & \\
6.28 & 0.03 & \\
6.30 & 0.04 & \\
6.35 & 0.05 & \\
6.40 & 0.07 & \\
6.45 & 0.08 & 20 \\
6.50 & 0.09 & 35 \\
6.60 & 0.12 & 57 \\
\hline
\hline
\end{tabular}
\end{table}
\section{Experiment}
Untwinned single crystals of YBa$_2$Cu$_3$O$_y$ ($y$=6.25, 6.28, 6.30, 6.35, 6.40, 6.45, 6.60) were grown by the flux method as described in Ref.~\cite{sample_growth}.
In the XPS study, we measured polycrystals of YBa$_{2}$Cu$_{3}$O$_y$
($y=$6.20, 6.40, 6.50, 6.60), too.
In YBCO, how the in-plane hole concentration $\delta$ changes with oxygen content $y$ has been controversial.
We have adopted the estimates from electrical resistivity and thermopower~\cite{Ong&Uchida}.
$y$, $\delta$ and $T_c$ of the measured samples are listed in Table~\ref{tab:table1}.
The ARPES measurements were carried out at BL10.0.1 of Advanced Light Source (ALS) using incident photons of 55 eV and a SCIENTA R4000 analyzer, at BL28 of Photon Factory (PF) using incident photons of 65 eV, a SCIENTA SES2002 analyzer, and at BL5-4 of Stanford Synchrotron Radiation Laboratory (SSRL) using photons of 28 eV and a SCIENTA SES-200 analyzer.
An R-Dec Co. Ltd. i GONIO LT goniometer~\cite{igonio} was used at PF.
The total energy and momentum resolution was about 20 meV and 0.02$\pi$ in units of 1/\textit{a} ($a=3.86~$\AA~is the lattice constant), respectively.
The samples were cleaved \textit{in situ} in an ultra high vaccum of $10^{-11}$~Torr and cooled to $\sim$ 10 K during the measurement.
The XPS measurements were carried out using the Mg $K{\alpha}$ line (${\it h}{\nu}=1253.6$~eV) and a VSW EAC 125 hemispherical analyzer for sintered polycrystalline samples.
Single cyrstals were measured using the Mg $K{\alpha}$ line (${\it h}{\nu}=1253.6$~eV) and a SCIENTA SES-100 analyzer.
The energy resolution was about $0.8$~eV, which was largely due to the width of the photon source.
Nevertheless we could determine the core-level shifts with an accuracy of about $\pm$50 meV because most of the spectral line shapes were identical between different compositions.
Details of the core-level XPS studies are described elsewhere~\cite{mu_review,CCOC_mu}.
\section{Results and Discussion}
Figure \ref{bilayer_EDC} shows energy distribution curves (EDC's) along high symmetry lines in the second Brillouin zone for the $y$=6.35 and 6.60 samples.
Two dispersive features are observed and both of them become more pronounced in going from y=6.35 to 6.60, reflecting the increased hole doping.
One (open circles) is around $-0.2$ eV at $\Gamma$, dispersing upward and crossing $E_{\rm F}$ around 60 \% of the $\Gamma$X and $\Gamma$Y lines.
In optimally-doped YBCO, this feature was assigned to surface states in Ref.~\cite{ARPES_Lu} and assigned to surface antibonding band in Ref.~\cite{YBCO_Takahashi}
The other feature (open triangles) was considered to be the surface bonding band according to Ref.~\cite{YBCO_Takahashi}.
The dispersion of the surface bonding band is unclear around $\Gamma$ because the broad and strong antibonding band masks this feature.
It should be noted that the energy position of the surface bonding band showed clear anisotropy, $-0.18$ eV at X and $-0.14$ eV at Y.
The 1D like chain-derived states which was observed in the spectra of optimally doped YBCO~\cite{ARPES_Lu} is not observed here, perhaps due to the oxygen deficiency in the CuO chain and also due to the difference in the photon energy and the polarization of the incident light, i.e., difference in the matrix elements.
Figure \ref{bilayer_mapping}(a) shows the plot of spectral weight in the second Brillouin zone integrated within 40 meV of $E_{\rm F}$, which reflects the Fermi surface, for $y$=6.45.
Not only a large hole-like Fermi surface centered at the S point, but also an apparently electron-like Fermi surface centered at $\Gamma$ are observed, indicating the bilayer splitting~\cite{YBCO_nodalQP,YBCO_bilayer}.
Open circles in Fig.~\ref{bilayer_mapping}(a) are the $k_{\rm F}$ positions
of the Fermi surfaces determined by the peak positions of momentum distribution curves (MDC's) at $E_{\rm F}$.
In Fig.~\ref{bilayer_mapping}(b), the $k_{\rm F}$ positions thus determined are plotted for various compositions in the same panel.
The shape of the Fermi surface does not vary appreciably in this doping range, suggesting that these Fermi surfaces are not derived from bulk.
Because the cleaved surface of YBCO tend to be overdoped~\cite{YBCO_bilayer,YBCO_Takahashi}, we compare the experiment with the Fermi surfaces predicted by local-density-approximation (LDA) calculation for $y$=7~\cite{LDA_Andersen} in Fig.~\ref{bilayer_mapping}(b).
In the LDA calculation, the Fermi surface has four sheets derived from the CuO$_2$ planes and the CuO chain.
However, because we did not observe the CuO chain-derived bands, we have plotted only the CuO$_2$ plane-derived sheets, that is, those derived from the bonding and antibonding bands of the CuO$_2$ bilayer.
The Fermi surfaces for two values of $k_{\rm z}$=0 and $\pi/c$ are indicated because for the optimally-doped YBCO the CuO chain mediates orbital overlap along the $c$-axis and therefore strong $k_{\rm z}$ dispersions are predicted.
The experimental results are more similar to the $k_{\rm z}$=0 cross-section than to the $k_{\rm z}$=$\pi/c$ one.
The shape of the Fermi surface of the bonding band is more similar to that of Bi2212 than to that of LSCO.
According to the tight binding model including hopping parameters $t$, $t'$, and $t''$, this suggests a large next-nearest-neighbour hopping $|t'|$ in YBCO~\cite{Tohyama_SST}.
In Fig.~\ref{bilayer_band}(a), we summarize the surface band dispersions of YBCO for various $y$'s obtained from EDC's.
Here, the surface bonding band for the $y$=6.28 sample was taken in a different experimental geometry ($\bm{E}\perp \bm{b}$) and photon energy (${\it h}{\nu}=65$~eV) than the rest of the data ($\bm{E}\parallel \rm{XY}$), and was clearly resolved even around the $\Gamma$ point.
Although the spectral line shape and intensities dramatically changed with hole doping (see Fig.~\ref{bilayer_EDC}), the band dispersions did not change appreciably in this doping range.
The lower bound for the bilayer splitting at X and Y is estimated to be 180 meV and 140 meV, respectively, because the surface antibonding band should be above $E_{\rm F}$ at these momenta.
Figure~\ref{bilayer_band}(b) shows the LDA band dispersions of YBCO ($y$=7) for $k_{\rm z}$=0~\cite{LDA_Andersen}.
Red curves are the bonding and antibonding bands of the CuO$_2$ bilayer.
The experimental surface band dispersions are qualitatively reproduced by the LDA band structure except for the $E_{\rm F}$ crossing points along the symmetry lines and the overall band narrowing (by a factor of $\sim$ 0.5) in experiment.
This similarity indicates that the origin of the surface state is the overdoped cleaved surface.~\cite{YBCO_Takahashi}
In the LDA calculation, the bonding-antibonding splitting of the CuO$_2$ bilayers at X and Y was 610 meV and 560 meV, respectively, which is much larger than the experimental value of $\gtrsim$ 180 meV and $\gtrsim$ 140 meV.
The reason for this dicrepancy is that LDA calculation neglects electron correlation and generally overestimates band dispersions.
Indeed, overdoped Bi2212 has shown a bilayer splitting of 88 meV~\cite{Feng}, much smaller than 300 meV predicted by LDA calculation~\cite{Chakravarty,LDA_Andersen}.
Bilayer Hubbard model calculations for two coupled CuO$_2$ layers, which explicitly include the on-site Coulomb repulsion, predicted a maximum splitting of 40 meV~\cite{bilayer_Hubbard}.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=9.5cm]{Bilayer_EDC.eps}
\caption{(Color online) EDC's of lightly-doped to underdoped YBCO. Two features in the spectra are surface bonding state (open triangles) and surface antibonding state (open circles).}
\label{bilayer_EDC}
\end{center}
\end{figure}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=8cm]{Bilayer_mapping.eps}
\caption{(Color online) (a) Spectral weight mapping at $E_{\rm F}$ for a 40 meV integration window. Open circles are the $k_{\rm F}$ positions determined from MDC's and the white arrow denotes the polarization of the incident light. (b) Experimentally determined $k_{\rm F}$ positions in each doping level and LDA Fermi surface for $k_{\rm z}$=0 and $k_{\rm z}$=$\pi$/$c$ related to the CuO$_2$ plane.}
\label{bilayer_mapping}
\end{center}
\end{figure}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=8cm]{Bilayer_band.eps}
\caption{(Color online) Band dispersions in YBCO obtained from EDC's (a) compared with the LDA band dispersions for YBCO (y=7) at $k_{\rm z}$=0~\cite{LDA_Andersen} (b). In (b), red curves are the antibonding and bonding bands of the CuO$_2$ planes.}
\label{bilayer_band}
\end{center}
\end{figure}
Figure 4(a)-(c) shows spectral weight mapping at $E_{\rm F}$ with a 40 meV integration window in the first Brillouin zone for the $y$=6.35, 6.45 and 6.60 samples.
A Fermi "arc" around the nodal direction is observed as in the LSCO case, which is completely different from Fig.~2(a).
Figure 2(a) does not show a Fermi arc but a hole-like and an electron-like Fermi surface arising from the surface state.
Figure 4(e) shows the intensity plot in the \textit{E}-$\bm{k}$ plane in the nodal direction for the $y$=6.35 sample together with the energy band dispersion determined from MDC peaks.
The dispersion of this band shows a kink structure around $\sim$ 50 meV like in many other high-$T_c$ cuprates, while the surface bonding band is reported to show a straight dispersion~\cite{YBCO_Takahashi}.
In addition, the $k_{\rm F}$ positions in the nodal direction in the first Brillouin zone are different from those in the second Brillouin zone.
These findings suggest that the Fermi arc observed in the first Brillouin zone is distinct from the Fermi surfaces observed in the second BZ and we attribute it to a bulk-derived feature.
To investigate the doping evolution of the Fermi arc, the $k_{\rm F}$ positions determiend by the peak positions of the MDC's are plotted by filled circles in Fig.~4(a)-(c).
The $k_{\rm F}$ positions for $y$=6.35, 6.45, and 6.60 are overlaid in Fig.~4(d).
As holes are doped, the hole-like Fermi surface centered at S slightly expands in the nodal region, which is also consistent with the assumption that these Fermi surfaces are bulk derived.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=8cm]{Fig4.eps}
\caption{(Color online) (a)-(c) Spectral weight mapping at $E_{\rm F}$ for a 40 meV integration window in the first Brillouin zone. Filled circles are the $k_{\rm F}$ positions determined from MDC's and the white arrow denotes the polarization of the incident light. (d) $k_{\rm F}$ positions for $y$=6.35, 6.45, and 6.60 determined from MDC's. (e) Intensity plot in the \textit{E}-$\bm{k}$ plane in the nodal direction for the $y$=6.35 sample. Filled circles denote the energy band dispersion determined by fitting the MDC's to Lorentzians.}
\label{Fig4}
\end{center}
\end{figure}
Next, we present the result of the measurements of XPS core levels in YBCO, and deduce the chemical potential shift as a function of doping.
As shown in the inset of Fig.~\ref{chemicalpotential}, the O 1\textit{s} and Y 3\textit{d} core levels show the same shifts.
As for the Ba 4\textit{d} core level, the peak becomes broader with hole doping, perhaps due to some surface effects, but apart from this the Ba 4\textit{d} peak position follow the shifts of the O 1\textit{s} and Y 3\textit{d} core levels, too.
The line shape of the Cu 2\textit{p} peak was not identical between the different compositions but its peak position moved in the opposite direction to those of the O 1\textit{s}, Y 3\textit{d} and Ba 4\textit{d} core levels.
The different behavior of the Cu 2\textit{p} core level can be attributed to the creation and the growth of ``Cu$^{3+}$" signals with hole doping on the higher binding energy side of the Cu$^{2+}$ main component~\cite{LSCO_mu}.
The inset of Fig.~\ref{chemicalpotential} shows the binding energy shifts referenced to $y$=6.2 of each core level thus estimated.
As in the previous studies~\cite{mu_review}, we consider that the parallel shifts of the O 1\textit{s}, Y 3\textit{d} and Ba 4\textit{d} core levels reflect the chemical potential shift $\Delta \mu$.
Recently, Maiti {\it et al.} reported the shifts of the core levels in YBCO using hard x-ray photoemission spectroscopy~\cite{Maiti}.
Our Y 3\textit{d} and O 1\textit{s} shifts are consistent with their results except for their lowest doped sample ($y$ = 6.15).
In the main panel of Fig.~\ref{chemicalpotential}, therefore, $\Delta \mu$ has been obtained by taking the average of the O 1\textit{s} and Y 3\textit{d} core-level shifts as shown in the figure.
In the same panel, the chemical potential shift in other high $T_c$ cuprates are plotted for comparison~\cite{LSCO_mu,CCOC_mu,Bi2212_mu,NCCO_mu}.
Theoretically, if the long-range antiferromagnetic order of the parent Mott insulator disappears for a very small hole concentration, the charge suceptibility $\partial n$/$\partial \mu$ diverges in the limit $\delta \to 0$~\cite{suppression} and hence the chemical potential is suppressed.
In LSCO and Bi2212, the antiferromagnetic order quickly disappears for a small amount of hole doping and indeed the shift is suppressed for small $\delta$.
In YBCO and NCCO, on the other hand, the antiferromagnetic order perisits up to $\delta \sim 0.07$ and $\delta \sim 0.14$, respectivly, and thefore it is reasonable that little or no suppression of the shift was observed for small $\delta$.
According to $t$-$t'$-$t''$-$J$ model calculations~\cite{cps_tj}, the larger the $|t'|$ becomes, the stronger the chemical potential shift becomes.
The observed differences between Bi2212 and LSCO, such as the different dispersion width of the LHB from ${\bm k}\sim (\pi/2,\pi/2)$ to $\sim (\pi,0)$, the different shape of the Fermi surface as well as the different behavior of the chemical potential shift, can be explained by a larger $|t'|$ value in Bi2212 than that in LSCO~\cite{cps_tj,tj_tanaka}.
$\Delta \mu$ in YBCO is the largest among those compounds, which means that the $|t'|$ of YBCO may be the largest, consistent with the conclusion from the ARPES results described above.
The suppression of the shift in the underdoped LSCO is too strong compared to
the $t$-$t'$-$t''$-$J$ model calculation and has been attributed to a fluctuation of charge stripes~\cite{LSCO_mu}.
The dynamical stripes seen in LSCO change their periodicity linearly with doping~\cite{stripe_LSCO2}.
As a result, the local charge density is nearly unchanged so that the hole rich region expands and the hole poor region shrinks as the average hole concentration is increased.
Under such a circumstance, the chemical potential would not move with doping.
If the periodicity of the stripes remains unchanged with doping, on the other hand, the chemical potential is expected to move in a rigid-band-like manner.
In this regard, the observations of in-plane anisotropic magnetoresistance~\cite{Ando_magnetoresistance} and electrical resistivity~\cite{Ando_transport} in untwinned single crystals of YBCO are most consistent with the idea that charge stripes also exist in lightly-doped YBCO.
The presence of the stripes and the strong chemical potential shift can be reconciled if the stripes in the lightly-doped YBCO are of the type of fixed periodicity.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=6cm]{chemicalpotential.eps}
\caption{(Color online) Chemical potential shift in YBCO compared with those in other high $T_c$ superconductors~\cite{LSCO_mu,CCOC_mu,Bi2212_mu,NCCO_mu}.
The antiferromagnetic region is displayed by shaded region for YBCO and NCCO. Inset shows the shift of each core level for YBCO as a function of hole concentration $\delta$.}
\label{chemicalpotential}
\end{center}
\end{figure}
To investigate the doping dependence of the shape of the Fermi surface and the QP band quantitatively, we have fitted the $k_{\rm F}$ positions shown in Fig.~4(d) to the two-dimensional $t$-$t'$-$t''$ tight binding model
\begin{eqnarray*}
\epsilon _k - \mu = &-&2t[{\rm cos}(k_x a) + {\rm cos}(k_y a)]-4t'{\rm cos}(k_x a){\rm cos}(k_y a)\\ &-&2t''[{\rm cos}(2k_x a)+{\rm cos}(2k_y a)]+\epsilon _0
\end{eqnarray*}
Here, $\epsilon _0$ is the position of the band center relative to the chemical potential.
We have assumed that $t''/t' = -1/2$ as before~\cite{Yoshida_cond,cps_tj,Bi2201_mu}, and regarded $-t'/t$ and $- \epsilon _0 /t$ as fitting parameters.
Figure 6(a) shows the shape of the Fermi surface from the tight-binding fit.
One can see the squarelike Fermi surface which suggests large $-t'/t$~\cite{Tohyama_SST}.
Figure 6(b) shows fitted tight-binding parameters as function of hole concentration.
As in the case of LSCO and Bi2201, $-t'/t$ shows weak doping dependence~\cite{Yoshida_cond,Bi2201_mu}.
However, the $-t'/t$ value is much larger in YBCO ($\sim$ 0.4) than that in LSCO ($\sim$ 0.2) and Bi2201 ($\sim$ 0.2), consistent with the XPS results.
If the chemical potential moves in a rigid-band-like manner, $\Delta \mu/t = \epsilon _0 /t$, as is not the case in YBCO and LSCO.
The doping dependence of $\epsilon _0 /t$ in YBCO is similar to that in LSCO while $-\Delta \mu /t$ increases faster than that in LSCO.
This difference may be attributed to the different type of charge stripes as described above.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=8cm]{Fig6.eps}
\caption{(Color online) Doping dependence of the electronic structure in YBCO. (a) Doping dependence of the Fermi surface shape. (b) Doping dependence of the fitted tight-binding parameters.}
\label{tbparameters}
\end{center}
\end{figure}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=8cm]{kf.eps}
\caption{(Color online) ARPES spectra taken at 10 K for lightly-doped
YBa$_2$Cu$_3$O$_y$ along the nodal direction in the first Brillouin zone.
$\delta$ indicates hole concentration per CuO$_2$ plane. (a) Top panel shows EDC's and bottom panel shows intensity plot in the \textit{E}-$\bm{k}$ plane. (b) EDC's at $k_{\rm F}$ for $y$=6.25, 6.30, 6.35, and 6.40 samples. (c) $k_{\rm F}$ positions for various doping levels in YBCO, LSCO~\cite{Yoshida_LSCO}, Na-CCOC~\cite{kfCCOC}, Bi2201~\cite{Bi2201_mu}, and Bi2212~\cite{Bi2212_kf}.}
\label{kf}
\end{center}
\end{figure}
Figure~\ref{kf}(a) shows EDC's (top) and corresponding intensity plots in \textit{E}-$\bm{k}$ space (middle) in the nodal (0,0)-($\pi$,$\pi$) direction of the $y$=6.30 sample.
As in the case of LSCO, a dispersive QP feature crossing $E_{\rm F}$ is seen for the lightly-doped $y\ge$6.30 ($\delta$=0.04) sample and its intensity increased with doping as shown in Fig.~\ref{kf}(b), giving rise to the metallic transport at high temperatures~\cite{Ando_transport}.
Figure \ref{kf}(c) shows the $k_{\rm F}$ positions along the nodal direction for various doping levels in YBCO, LSCO, Na-CCOC, Bi2201, and Bi2212.
In the antiferromagnetic (AF) insulating state of the undoped compound, the valence-band maximum along the nodal direction should be located at the AF zone boundary, that is, exactly at ($\pi/2,\pi/2$) due to the band folding in the AF state.
In Na-CCOC and the Bi-based compounds, indeed, with decreasing hole concentration the $k_{\rm F}$ position moves toward ($\pi/2,\pi/2$).
In LSCO, on the other hand, $k_{\rm F}$ extrapolates to $\sim$($0.44\pi ,0.44\pi$) and not to ($\pi/2,\pi/2$)~\cite{Yoshida_cond}.
In LSCO, in addition to the QP feature crossing $E_{\rm F}$, the broad ``LHB'' or the polaronic feature at $\sim$ -(0.4-0.6) eV is recognized for low hole concentrations $x\leq 0.05$.
Such an electronic structure has been regarded as the presence of the coherent and incoherent parts of the spectral function influenced by electron-electron and electron-phonon interacitons~\cite{Roesch} and/or due to a nano-scale electronic inhomogeneity such as charge stripes~\cite{Mayr}.
Since such an inhomogeneity is also expected to exist in lightly-doped YBCO~\cite{Ando_magnetoresistance,Ando_transport}, the fact that $k_{\rm F}$ in YBCO did not extrapolate to ($\pi/2,\pi/2$) suggests that QP band is also separated from the ``LHB'' in YBCO.
Unfortunately, such a remnant of the ``LHB'' is not seen in YBCO at least down to $\sim -$0.80 eV, partly because it is masked by the strong overcapping tail of the O 2\textit{p} band and partly because spectral weight transfer from the LHB to the QP band occurs very quickly with hole doping.
\section{Conclusion}
We have observed the band dispersions and Fermi surfaces of lightly-doped to underdoped YBCO by ARPES using untwinned single crystals.
The shape of the Fermi surface and the large chemical potential shift indicated large $|t'|$ like Bi2212, Bi2202, and Na-CCOC.
On the other hand, we observed a rather clear dispersive QP feature crossing $E_{\rm F}$ along the nodal direction, revealing a similarity to LSCO.
Although the coexistence of the remnant LHB could not be confirmed unlike the other cuprates, we conclude that very rapid spectral weight transfer ocurrs from the LHB to the QP band upon hole doping in the lightly-doped to underdoped YBCO.
We have also studied the chemical potential shift with hole doping by the XPS measurements of single and polycrystals.
The shift was as rapid as those in Na-CCOC, Bi2201, and Bi2212 in spite of the fact that charge stripes are observed in lightly-doped YBCO like LSCO.
We attribute the different chemical potential shifts between YBCO and LSCO to different types of charge stripes.
\section{Acknowledgement}
This work was supported by a Grant-in Aid for Scientific Research in Priority Area ``Invention of Anomalous Quantum Materials" (16076208) from MEXT, Japan.
ALS and SSRL is operated by the Department of Energy's Office of Basic Energy Science, Division of Chemical Sciences and Materials Sciences.
Experiment at Photon Factory was done under the approval of the Photon Factory Program Advisory Committee (Proposal Nos.~2006S2-001 and 2009S2-005).
The work at Osaka University was supported by KAKENHI Grant Nos.~19674002, 20030004, and 20740196.
|
\section{Introduction}\label{i}
In a previous paper \cite{cl3} we investigate the influence of the diffusion coefficient $a$ on the life span of solutions to the one dimensional Smoluchowski-Poisson system
\begin{eqnarray}
\label{he1}
\partial_t{u} &=& \partial_x \left( a(u) \partial_x u -u \partial_x v \right) \;\;\mbox{in}\;\;(0,\infty)\times (0,1),\\
\label{he12}
0&=&\partial_x^2 v + u - M \;\;\mbox{in}\;\;(0,\infty)\times (0,1),\\
\label{he2}
a(u) \partial_x u & = & \partial_x v = 0\;\;\mbox{on} \;\; (0,\infty)\times \{0,1\},\\
\label{he3}
u(0)&=& u_0\ge 0 \;\;\mbox{in} \;\;(0,1),\;\;\int_0^1v(t,x)dx=0\;\;\mbox{for any}\;\; t\in (0,\infty),
\end{eqnarray}
where
$$
M := \langle u_0 \rangle = \int_0^1 u_0(x) dx
$$
denotes the mean value of $u_0$, and uncover a fundamental difference with the quasilinear Smoluchowski-Poisson system in higher space dimensions. More precisely, when the space dimension $n$ is greater or equal to two, there is a critical diffusion $a_*(r):= (1+r)^{(n-2)/2}$ which separates different behaviours for the quasilinear Smoluchowski-Poisson system. Roughly speaking,
\begin{description}
\item[(a)] if the diffusion coefficient $a$ is stronger than $a_*$ (in the sense that $a(r)\ge C (1+r)^\alpha$ for some $\alpha>(n-2)/n$ and $C>0$), then all solutions exist globally whatever the value of the mass of the initial condition $u_0$ \cite{cw_1},
\item[(b)] if the diffusion coefficient $a$ is weaker than $a_*$ (in the sense that $a(r)\le C (1+r)^\alpha$ for some $\alpha<(n-2)/n$ and $C>0$), then there exists for all $M>0$ an initial condition $u_0$ with $\langle u_0 \rangle=M$ for which the corresponding solution to the quasilinear Smoluchowski-Poisson system blows up in finite time (in the sense that $\|u(t)\|_\infty\to\infty$ as $t\to T$ for some $T\in (0,\infty)$) \cite{cl1,cw_1,Nagai},
\item[(c)] if the diffusion coefficient $a$ behaves as $a_*$ for large values of $r$, solutions starting from initial data $u_0$ with small mass $\langle u_0 \rangle$ exist globally while there are initial data with large mass for which the corresponding solution to the quasilinear Smoluchowski-Poisson system blows up in finite time \cite{cl1,Nagai}.
\end{description}
Observe that, in space dimension $n=2$, the critical diffusion is constant and a more precise description of the situation {\bf (c)} is actually available. Namely, when $a\equiv 1$, there is a threshold mass $M_*$ such that, if $\langle u_0 \rangle<M_*$, the corresponding solution is global while, for any $M>M_*$, there are initial data with $\langle u_0 \rangle=M$ for which the corresponding solution blows up in finite time \cite{JL92,Nagai,Nagai2}. The threshold mass $M_*$ is known explicitly ($M_*=4\pi$) but it is worth mentioning that for radially symmetric solutions in a ball, the threshold mass is $8\pi$. Similar results are also available for the quasilinear Smoluchowski-Poisson system in $\RR^n$, $n\ge 2$ \cite{B-C-L,ChS04,sugi,sugi2}.
Most surprisingly, the above description fails to be valid in one space dimension and we prove in particular in \cite{cl3} that all solutions are global for the diffusion $a(r)=(1+r)^{-1}$ though it is a natural candidate to be critical. We actually identify two classes of diffusion coefficients $a$ in \cite{cl3}, one for which all solutions exist globally as in {\bf (a)} and the other for which there are solutions blowing-up in finite time starting from initial data with an arbitrary positive mass as in {\bf (b)}, but the situation {\bf (c)} does not seem to occur in one space dimension. The purpose of this note is to show that the dichotomy {\bf (a)} or {\bf (b)} can be extended to larger classes of diffusion, thereby extending the analysis performed in \cite{cl3}.
\begin{theo}\label{mainth}
Let the diffusion coefficient $a\in \mathcal{C}^1((0,\infty))$ be a positive function.
\smallskip
\noindent
(i) Assume first that $a\in L^1(1,\infty)$ and one of the following assumptions is satisfied, either
\begin{equation}\label{(1)}
\gamma:=\sup_{r\in(0,1)} r\int_r^{\infty}a(s)ds<\infty,
\end{equation}
or there exist $\vartheta>0$ and $\alpha\in (\vartheta/(1+\vartheta),2]$ such that
\begin{equation}\label{decr}
\gamma_\vartheta:=\sup_{r\in (0,1)}{ r^{2+\vartheta} a(r) } < \infty \;\;\mbox{ and }\;\; C_\infty := \sup_{r\ge 1} r^\alpha a(r) < \infty.
\end{equation}
For any $M>0$, there exists a positive initial condition $u_0\in\mathcal{C}([0,1])$ such that $\langle u_0 \rangle = M$ and the corresponding classical solution to (\ref{he1})-(\ref{he3}) blows up in finite time.
\smallskip
\noindent
(ii) Assume next that $a\not\in L^1(1,\infty)$ and consider an initial condition $u_0\in\mathcal{C}([0,1])$ such that $u_0\ge m_0>0$ and $\langle u_0 \rangle = M$ for some $M>0$ and $m_0\in (0,M)$. Then the corresponding classical solution to (\ref{he1})-(\ref{he3}) exists globally.
\end{theo}
As already mentioned, Theorem~\ref{mainth} extends the results obtained in \cite{cl3}. More precisely, in \cite[Theorem~5]{cl3}, the assertion~(ii) of Theorem~\ref{mainth} is proved under the additional assumption that, for each $\varepsilon\in (0,\infty)$, there is $\kappa_\varepsilon>0$ for which
$$
a(r) \le \varepsilon\ r a(r) + \frac{\kappa_\varepsilon}{r} \;\;\mbox{ for }\;\; r\in (0,1)\,,
$$
which roughly means that $a$ cannot have a singularity stronger than $1/r$ near $r=0$. This assumption turns out to be unnecessary for global existence but nevertheless ensures the global boundedness of the solution in $L^\infty$. Under the sole assumption of Theorem~\ref{mainth}~(ii), our proof does not exclude that solution to (\ref{he1})-(\ref{he3}) becomes unbounded as $t\to\infty$. Concerning Theorem~\ref{mainth}~(i), it is established in \cite[Theorem~10]{cl3} for $a\in L^1(1,\infty)$ such that there is a concave function $B$ for which
\begin{eqnarray}
0 \le -r A(r) & \le & B(r) \;\;\mbox{ with }\;\; A(r) = - \int_r^\infty a(s)\ ds\,, \qquad r\in (0,\infty)\,, \label{bu1} \\
\lim_{r\to \infty} \frac{B(r)}{r} & = & 0\,. \label{bu2}
\end{eqnarray}
We make this criterion more explicit here by showing that the integrability of $a$ on $(1,\infty)$ and (\ref{(1)}) guarantee the existence of a concave function $B$ satisfying (\ref{bu1}) and (\ref{bu2}), see Lemma~\ref{imp} below. Let us point out here that the assumption (\ref{(1)}) somehow means that $a$ cannot have a singularity stronger that $1/r^2$ near $r=0$. However, the result remains true if $a$ has an algebraic singularity of higher order near $r=0$ which is allowed by (\ref{decr}) provided $a$ decays suitably at infinity. Observe that the second condition in (\ref{decr}) is compatible with the integrability of $a$ at infinity as $\vartheta/(1+\vartheta)<1$.
Summarizing the outcome of Theorem~\ref{mainth}, we realize that, for a given diffusion coefficient $a$ with a singularity weaker than $1/r^2$ near $r=0$, the integrability or non-integrability of $a$ at infinity completely determines whether we are in the situation {\bf (a)} or {\bf (b)} described above and excludes the situation {\bf (c)}. There is thus no critical diffusion in this class. The same comment applies to the class of diffusion coefficients satisfying (\ref{decr}) with an algebraic singularity stronger than $1/r^2$ near $r=0$. In particular there is no critical nonlinearity in the class of functions ${\cal C}([0,\infty))\cap {\cal C}^1((0,\infty))$.
The paper is organized as follows: in section~\ref{pre} we recall some statements from \cite{cl3}. Section~\ref{ftb} is devoted to proving the finite time blowup of solutions to (\ref{he1})-(\ref{he3}) when $a\in L^1(1,\infty)$. Global existence of solutions for all initial data when $a$ is not integrable at infinity is proved in the last section.
\section{Preliminaries.}\label{pre}
In this section we summarize some results and methods introduced in \cite{cl3}. Let $a\in \mathcal{C}^1((0,\infty))$ be a positive function and consider an initial condition $u_0\in\mathcal{C}([0,1])$ such that $u_0\ge m_0>0$ and $\langle u_0 \rangle = M$ for some $M>0$ and $m_0\in (0,M)$. By \cite[Propositions~2 and~3]{cl3} there is a unique maximal classical solution $(u,v)$ to (\ref{he1})-(\ref{he3}) defined on $[0,T_{max})$ which satisfies
\begin{equation}
\label{zz1}
\min_{x\in [0,1]}{u(t,x)} > 0\,, \quad \langle u(t) \rangle := \int_0^1 u(t,x)\ dx = M\,, \;\;\mbox{ and }\;\; \langle v(t) \rangle := \int_0^1 v(t,x)\ dx = 0
\end{equation}
for $t\in (0,T_{max})$. In addition, $T_{max}=\infty$ or $T_{max}<\infty$ with $\|u(t)\|_\infty\to \infty$ as $t\to T_{max}$.
We next recall the approach introduced in \cite{cl3} which will be used herein as well. Owing to the positivity (\ref{zz1}) and the regularity of $u$, the indefinite integral
$$
U(t,x) := \int_0^x u(t,z) dz\,, \quad x\in [0,1]\,,
$$
is a smooth increasing function from $[0,1]$ onto $[0,M]$ for each $t\in [0,T_{max})$ and has a smooth inverse $F$ defined by
\begin{equation}\label{zero}
U(t,F(t,y))=y\,, \qquad (t,y)\in [0,T_{max})\times [0,M]\,.
\end{equation}
Introducing $f(t,y):=\partial_y F(t,y)$, we have
\begin{equation}\label{odw}
f(t,y)\ u(t,F(t,y))=1\,, \qquad (t,y)\in [0,T_{max})\times [0,M]\,,
\end{equation}
and it follows from (\ref{he1})-(\ref{he3}) that $f$ solves
\begin{eqnarray}
\partial_t f & = & \partial_y^2 \Psi(f)-1+Mf\,, \qquad (t,y)\in (0,T_{max})\times (0,M)\,, \label{main} \\
\partial_y f(t,0) & = & \partial_y f(t,M)=0\,, \qquad t\in (0,T_{max})\,, \label{boundary} \\
f(0,y) & = & f_0(y) := \frac{1}{u_0(F(0,y))}\,, \qquad y\in (0,M)\,, \label{init}
\end{eqnarray}
where
\begin{equation}\label{Psi}
\Psi'(r):=\frac{1}{r^2}\ a\left( \frac{1}{r} \right) \;\;\mbox{for any}\;\;r>0\,, \qquad \Psi(1):=0\,,
\end{equation}
Moreover the conservation of mass (\ref{zz1}) yields
\begin{equation}\label{mas}
\int_0^M f(t,y) dy=F(t,M)-F(t,0)=1\,, \qquad t\in [0,T_{max})\,.
\end{equation}
At this point, the crucial observation is that, thanks to (\ref{odw}), finite time blowup of $u$ is equivalent to the vanishing (or touch-down) of $f$. In other words, $u$ exist globally if the minimum of $f(t)$ is positive for each $t>0$. We refer to \cite[Proposition 1]{cl3} for a more detailed description.
An salient property of (\ref{he1})-(\ref{he3}) is the existence of a Liapunov function \cite[Lemma~8]{cl3} which we recall now:
\begin{Le}\label{lem}
The function
$$
L_1(t):= \frac{1}{2} \int_0^M \left| \partial_y\Psi(f(t,y)) \right|^2\ dy+\int_0^M \left( \Psi(f(t,y)) -M\ \Psi_1(f(t,y)) \right)\ dy
$$
is a non-increasing function of time on $[0,T_{max})$, the function $\Psi_1$ being defined by
\begin{equation}
\label{gex3}
\Psi_1(1):=0 \;\;\mbox{ and }\;\; \Psi_1'(r) := r \Psi'(r) = \frac{1}{r}\ a \left( \frac{1}{r} \right) \,, \qquad r\in (0,\infty)\,.
\end{equation}
\end{Le}
\section{Finite time blowup.}\label{ftb}
In this section we prove the blowup assertion of Theorem~\ref{mainth}. To this end we first prove that the condition (\ref{(1)}) allows us to construct a concave function $B$ satisfying (\ref{bu1}) and (\ref{bu2}) so that \cite[Theorem~10]{cl3} can be applied.
\begin{Le}\label{imp}
Let $a\in {\cal C}^1((0,\infty))$ be a positive function such that $a\in L^1(1,\infty)$ and (\ref{(1)}) holds.
Then there exists a concave function $B\in {\cal C}([0,\infty))$ such that for all $r\geq 0$
\begin{equation}\label{(3)}
B(r)\geq r\int_r^{\infty}a(s)ds
\end{equation}
and
\begin{equation}\label{(4)}
\lim_{r\rightarrow\infty}\frac{B(r)}{r}=0.
\end{equation}
\end{Le}
\vspace{0.3cm}
\noindent
\textit{Proof of Lemma~\ref{imp}.} We construct $B:[0,\infty)\rightarrow[0,\infty)$ in the following way: we put
$$
b_i:=\int_{2^i}^\infty a(s)ds\,, \qquad i\ge 0\,,
$$
and notice that $\{b_i\}_{i\ge 0}$ is a decreasing sequence converging to zero as $i\to\infty$. We next define
\begin{equation}
B(r) = \left\{\begin{array}{ccl}
\displaystyle{b_0 r+\gamma}& \mbox{if} &r\in[0,2],\\
& & \\
\displaystyle{b_i r+\sum_{j=0}^{i-1}(b_j-b_{j+1})2^{j+1}+\gamma}& \mbox{if}&r\in(2^i,2^{i+1}] \;\mbox{ and }\; i\geq 1,
\end{array}\right.
\end{equation}
Clearly, $B\in \mathcal{C}([0,\infty))$ and
\begin{equation}
B'(r) = \left\{\begin{array}{ccl}
b_0 & \mbox{if} &r\in(0,2),\\
b_i & \mbox{if} &r\in(2^i,2^{i+1}) \;\mbox{ and }\; i\geq 1.
\end{array}\right.
\end{equation}
Hence $B$ is concave as a consequence of the fact that the sequence $\{b_i\}_{i\ge 0}$ is decreasing.
Furthermore, for $r\in[0,1]$, we have
\[
B(r)\geq \gamma\geq r\int_r^{\infty}a(s)ds,
\]
and for $r\in[2^i,2^{i+1}]$, $i\geq 0$,
\[
B(r)\geq b_i r=r\int_{2^i}^\infty a(s)ds\geq r\int_r^\infty a(s)ds.
\]
Therefore, $B$ satisfies (\ref{(3)}).
Finally, let $k\ge 1$. If $i\geq k+1$ and $r\in(2^i,2^{i+1}]$, then
\begin{eqnarray*}
\frac{B(r)}{r} &=& b_i+\frac{\gamma}{r}+\sum_{j=0}^{i-1}(b_j-b_{j+1})\frac{2^{j+1}}{r}\leq b_i+\frac{\gamma}{r}+\sum_{j=k}^{i-1}(b_j-b_{j+1})+\sum_{j=0}^{k-1}(b_j-b_{j+1})\frac{2^{j+1}}{r} \\
&\le & b_i+\frac{1}{r}\left(\gamma+2^k\sum_{j=0}^{k-1}(b_j-b_{j+1})\right)+(b_k-b_i)\leq b_k+\frac{1}{r}\left(\gamma+2^kb_0\right).
\end{eqnarray*}
Consequently,
\[
\limsup_{r\rightarrow\infty}\frac{B(r)}{r} \leq b_k\;\mbox{for all}\;k\geq1\,.
\]
Letting $k\rightarrow\infty$, we obtain (\ref{(4)}) since $b_k\to 0$ as $k\to\infty$ and Lemma~\ref{imp} is proved. \qed
\vspace{0.3cm}
\noindent
\textit{Proof of Theorem~\ref{mainth}~(i), Part~1}. When $a$ belongs to $L^1(1,\infty)$ and satisfies (\ref{(1)}), it follows from Lemma~\ref{imp} that the conditions (\ref{bu1}) and (\ref{bu2}) are satisfied so that Theorem~\ref{mainth}~(i) follows from \cite[Theorem~10]{cl3}. \qed
\medskip
To handle the other case, we proceed in a different way by showing an upper bound for the function $f$ defined in section~\ref{pre}. We first observe that the function $\Psi$ defined in (\ref{Psi}) satisfies
$$
\Psi(r) = \int_1^\infty \frac{1}{s^2} a\left( \frac{1}{s} \right)\ ds = \int_{1/r}^1 a(s)\ ds\,, \quad r\in (0,\infty)\,,
$$
so that, if $a\in L^1(1,\infty)$, $\Psi(r)$ has a finite limit $\Psi(0):= - \|a\|_{L^1(1,\infty)}$ as $r\to 0$. We then define
\begin{equation}\label{Psit}
\tilde{\Psi}(r) := \Psi(r) - \Psi(0) = \int_0^r \frac{1}{s^2} a\left( \frac{1}{s} \right)\ ds = \int_{1/r}^\infty a(s)\ ds\,, \quad r\in (0,\infty)\,.
\end{equation}
\begin{Le}\label{impo}
Let $a\in {\cal C}^1((0,\infty))$ be a positive function such that $a\in L^1(1,\infty)$. There exists a positive constant $\mu_M>0$ depending only on $M$ and $a$ such that, for any non-negative function $g\in H^1(0,M)$ satisfying $\|g\|_{L^1(0,M)}=1$, we have
\begin{equation}\label{boundb}
\|\tilde{\Psi}(g)\|_{L^\infty(0,M)}^2 \le 32M \mathcal{L}_1(g) + \mu_M,
\end{equation}
with
\begin{equation}\label{pim}
\mathcal{L}_1(g) := \frac{1}{2} \|\partial_y \Psi(g)\|_{L^2(0,M)}^2 + \int_0^M \left( \Psi(g) - M \Psi_1(g) \right)(y)\ dy\,,
\end{equation}
the functions $\Psi$ and $\Psi_1$ being defined in (\ref{Psi}) and (\ref{gex3}), respectively.
\end{Le}
\vspace{0.3cm}
\noindent
\textit{Proof of Lemma~\ref{impo}.} We set $G:= \|g\|_{L^\infty(0,M)}$ which is finite owing to the continuous embedding of $H^1(0,M)$ in $L^\infty(0,M)$. Assume first that $G>1$. Then, for $y\in (0,M)$ and $z\in (0,M)$, we have
$$
\tilde{\Psi}(g(y)) = \tilde{\Psi}(g(z)) + \int_z^y \partial_{x} \tilde{\Psi}(g(x))\ dx \le \tilde{\Psi}(g(z)) + M^{1/2} \|\partial_y \Psi(g)\|_{L^2(0,M)}.
$$
Integrating the above inequality over $(0,M)$ with respect to $z$ gives
\begin{eqnarray*}
M\tilde{\Psi}(g(y)) & \le & \int_0^M \tilde{\Psi}(g(z))\ dz + M^{3/2} \|\partial_y \Psi(g)\|_{L^2(0,M)} \\
& \le & \int_0^M \mathbf{1}_{[0,2/M]}(g(z)) \tilde{\Psi}(g(z))\ dz + \int_0^M \mathbf{1}_{(2/M,\infty)}(g(z)) \tilde{\Psi}(g(z))\ dz + M^{3/2} \|\partial_y \Psi(g)\|_{L^2(0,M)} \\
& \le & M \tilde{\Psi}\left( \frac{2}{M} \right) + \frac{M \tilde{\Psi}(G)}{2}\ \int_0^M \mathbf{1}_{(2/M,\infty)}(g(z)) g(z)\ dz + M^{3/2} \|\partial_y \Psi(g)\|_{L^2(0,M)} \\
& \le & M \tilde{\Psi}\left( \frac{2}{M} \right) + \frac{M \tilde{\Psi}(G)}{2} + M^{3/2} \|\partial_y \Psi(g)\|_{L^2(0,M)}\,,
\end{eqnarray*}
where we have used the property $\|g\|_{L^1(0,M)}=1$ to obtain the last inequality. Taking the supremum over $y\in (0,M)$ and using the monotonicity and non-negativity of $\tilde{\Psi}$, we deduce that
\begin{equation}\label{y4b}
\tilde{\Psi}(G) \le 2\tilde{\Psi}\left( \frac{2}{M} \right) + 2M^{1/2} \|\partial_y \Psi(g)\|_{L^2(0,M)}.
\end{equation}
We next observe that the integrability of $a$ at infinity also ensures that $\Psi_1(0)>-\infty$, so that $\tilde{\Psi}_1:= \Psi_1-\Psi_1(0)$ is well-defined and satisfies
\begin{equation}\label{bas}
\tilde{\Psi}_1(r) = \int_0^r s \Psi'(s)\ ds \le r \tilde{\Psi}(r)\,, \quad r\in (0,\infty)\,.
\end{equation}
Since $\|g\|_{L^1(0,M)}=1$, it follows from (\ref{y4b}) and (\ref{bas}) that
\begin{equation}\label{y4c}
\int_0^M \tilde{\Psi}_1(g)\ dy \le \int_0^M g \tilde{\Psi}(g)\ dy \le \tilde{\Psi}(G)\ \int_0^M g dy \le 2\tilde{\Psi}\left( \frac{2}{M} \right) + 2M^{1/2} \|\partial_y \Psi(g)\|_{L^2(0,M)}.
\end{equation}
We next infer from (\ref{y4c}) and the non-negativity of $\tilde{\Psi}$ that
\begin{eqnarray*}
\mathcal{L}_1(g) & \ge & \frac{1}{2} \|\partial_y \Psi(g)\|_{L^2(0,M)}^2 + \int_0^M \tilde{\Psi}(g)\ dy + M \Psi(0) - M \int_0^M \tilde{\Psi}_1(g)\ dy \\
& \ge & \frac{1}{2} \|\partial_y \Psi(g)\|_{L^2(0,M)}^2 + M \Psi(0) - 2M \tilde{\Psi}\left( \frac{2}{M} \right) - 2M^{3/2} \|\partial_y \Psi(g)\|_{L^2(0,M)}\\
& \ge & \frac{1}{4} \|\partial_y \Psi(g)\|_{L^2(0,M)}^2 + \left( \frac{1}{2} \|\partial_y \Psi(g)\|_{L^2(0,M)} - 2 M^{3/2} \right)^2 - 4M^3 + M \Psi(0) - 2M \tilde{\Psi}\left( \frac{2}{M} \right) \\
& \ge & \frac{1}{4} \|\partial_y \Psi(g)\|_{L^2(0,M)}^2 - 4M^3 + M \Psi(0) - 2M \tilde{\Psi}\left( \frac{2}{M} \right) ,
\end{eqnarray*}
whence
$$
\|\partial_y \Psi(g)\|_{L^2(0,M)}^2 \le 4 \mathcal{L}_1(g) + 16M^3 - 4M \Psi(0) +8M \tilde{\Psi}\left( \frac{2}{M} \right) .
$$
It then follows from (\ref{y4b}) and the above inequality that
\begin{eqnarray*}
\tilde{\Psi}(G)^2 & \le & 8 \tilde{\Psi}\left( \frac{2}{M} \right)^2 + 8M \|\partial_y \Psi(g)\|_{L^2(0,M)}^2 \\
& \le & 8 \tilde{\Psi}\left( \frac{2}{M} \right)^2 + 32M \mathcal{L}_1(g) + 128M^4 - 32M^2 \Psi(0) +64M^2 \tilde{\Psi}\left( \frac{2}{M} \right) \\
& \le & 32M \mathcal{L}_1(g) + \mu_M,
\end{eqnarray*}
with
$$
\mu_M := 1 + 128M^4 - 32M^2 \Psi(0) +64M^2 \tilde{\Psi}\left( \frac{2}{M} \right) + 8 \tilde{\Psi}\left( \frac{2}{M} \right)^2 + \Psi(0)^2 - 32M\Psi(0).
$$
We have thus shown Lemma~\ref{impo} when $G=\|g\|_{L^\infty(0,M)}>1$. To complete the proof, we finally consider the case $G\in [0,1]$ and notice that, in that case,
$$
0 \le \tilde{\Psi}(G) \le -\Psi(0) \;\;\mbox{ and }\;\; \mathcal{L}_1(g) \ge \int_0^M \tilde{\Psi}(g)\ dy + M \Psi(0) \ge M \Psi(0) ,
$$
since $\Psi_1\le 0$ in $(0,1)$ and $\tilde{\Psi}\ge 0$. Consequently,
\begin{eqnarray*}
\tilde{\Psi}(G)^2 & \le & \Psi(0)^2 = 32M \Psi(0) + \Psi(0)^2 - 32M\Psi(0) \le 32M \mathcal{L}_1(g) + \mu_M,
\end{eqnarray*}
and the proof of Lemma~\ref{impo} is complete. \qed
\medskip
As an obvious consequence of Lemmas~\ref{lem} and~\ref{impo} we have the following result:
\begin{Cor}\label{cor:y2}
Let $a\in {\cal C}^1((0,\infty))$ be a positive function such that $a\in L^1(1,\infty)$. For $t\in [0,T_{max})$ and $y\in [0,M]$, we have
$$
0 \le \tilde{\Psi}(f(t,y)) \le \left( 32M \max{\{\mathcal{L}_1(f_0),0\}} + \mu_M \right)^{1/2}.
$$
\end{Cor}
\vspace{0.3cm}
\noindent
\textit{Proof of Corollary~\ref{cor:y2}}. Clearly $\mathcal{L}_1(f(t))=L_1(t)\le L_1(0)=\mathcal{L}_1(f_0)\le \max{\{\mathcal{L}_1(f_0),0\}}$ for $t\in [0,T_{max})$ by Lemma~\ref{lem} and Corollary~\ref{cor:y2} readily follows from Lemma~\ref{impo}. \qed
\medskip
\begin{Rem}\label{rem:positivity}
Corollary~\ref{cor:y2} provides an $L^\infty$-bound on $f$ only if $\Psi(r)\to\infty$ as $r\to\infty$, that is, if $a\not\in L^1(0,1)$. In that case, it gives a positive lower bound for $u$ by (\ref{odw}).
\end{Rem}
We next turn to the proof of the second part of Theorem~\ref{mainth} for which we develop further the arguments from \cite[Theorem~10]{cl3}.
\vspace{0.3cm}
\noindent
\textit{Proof of Theorem~\ref{mainth}~(i), Part~2}. Assume now that $a\in L^1(1,\infty)$ and satisfies (\ref{decr}). We fix $M>0$, $q>2$, and $\varepsilon_M\in (0,1)$ such that
\begin{equation}\label{prunelle}
q> \max{\left\{ 3+\vartheta, \frac{5+3\vartheta}{\alpha(\vartheta+1)-\vartheta} \right\}} \;\;\;\mbox{ and }\;\;\; \frac{q(q+1)}{M^2} \int_{1/\varepsilon_M}^\infty a(s)\ ds \le \frac{1}{2}\,,
\end{equation}
the existence of $\varepsilon_M$ being guaranteed by the integrability of $a$ at infinity.
For
\begin{equation}\label{gaston}
\delta\in \left( 0,\min{\left\{ 1 , 2M ,(2M)^{-1/q} \right\}} \right)\,,
\end{equation}
we put
\begin{equation}\label{pam}
f_0(y) := \frac{2(1-M\delta^q)}{\delta^2}\ (\delta-y)_+ + \delta^q\ge \delta^q>0\,, \quad y\in [0,M]\,.
\end{equation}
Then
\begin{equation}\label{poum}
\int_0^M f_0(y)\ dy = 1, \quad \|f_0\|_{L^\infty(0,M)} = \frac{2(1-M\delta^q)}{\delta} + \delta^q\le \frac{2}{\delta}\,.
\end{equation}
Introducing next
$$
m_q(t) := \int_0^M y^q f(t,y)\ dy\,, \quad t\in [0,T_{max})\,,
$$
we have
\begin{equation}\label{boule}
m_q(0) = \left( \frac{2(1-M\delta^q)}{(q+1)(q+2)} + \frac{M^{q+1}}{q+1}\right)\ \delta^q \le C_1 \delta^q \;\;\mbox{ with }\;\; C_1 := \left( \frac{2 +(q+2) M^{q+1}}{(q+1)(q+2)} \right)\,.
\end{equation}
It follows from (\ref{main}), (\ref{boundary}), and the non-negativity of $\tilde{\Psi}$ that
\begin{eqnarray}
\frac{dm_q}{dt} & = & - q\ \int_0^M y^{q-1} \partial_y \tilde{\Psi}(f)\ dy + M m_q - \frac{M^{q+1}}{q+1}, \nonumber \\
\frac{dm_q}{dt} & \le & q(q-1)\ \int_0^M y^{q-2} \tilde{\Psi}(f)\ dy + M m_q - \frac{M^{q+1}}{q+1} \label{y12}\,.
\end{eqnarray}
We shall now estimate the integral on the right-hand side of (\ref{y12}): to this end, we split the domain of integration into three parts which we handle differently. As a preliminary step, we notice that, by (\ref{decr}),
\begin{equation}\label{y12b}
\Psi'(r) \le \gamma_\vartheta r^\vartheta \;\;\;\mbox{ and }\;\;\; \Psi(r) \le \frac{\gamma_\vartheta}{\vartheta+1}\ r^{\vartheta+1} \le \gamma_\vartheta\ r^{\vartheta+1}\,, \quad r\ge 1\,.
\end{equation}
We next define
$$
K_0:= \left( 32M \max{\{\mathcal{L}_1(f_0),0\}} + \mu_M \right)^{1/2(2+\vartheta)} > 1\,,
$$
and consider $(t,y)\in [0,T_{max})\times [0,M]$.
\begin{itemize}
\item If $f(t,y)\in (0,\varepsilon_M]$, it follows from (\ref{prunelle}) and the monotonicity of $\tilde{\Psi}$ that
\begin{equation}\label{jeden}
\tilde{\Psi}(f(t,y)) \le \tilde{\Psi}(\varepsilon_M) = \int_{1/\varepsilon_M}^\infty a(s)\ ds \le \frac{M^2}{2q(q+1)}\,.
\end{equation}
\item If $f(t,y)\in (\varepsilon_M,K_0)$, then (\ref{y12b}) and the monotonicity of $\tilde{\Psi}$ yield
\begin{equation}\label{dwa}
\tilde{\Psi}(f(t,y)) = \frac{\tilde{\Psi}(f(t,y))}{f(t,y)}\ f(t,y) \le \frac{\Psi(K_0)-\Psi(0)}{\varepsilon_M}\ f(t,y) \le \frac{\gamma_\vartheta K_0^{\vartheta+1} - \Psi(0)}{\varepsilon_M}\ f(t,y)\,.
\end{equation}
\item If $f(t,y)\ge K_0$, Corollary~\ref{cor:y2} ensures that
\begin{equation}\label{trzy}
\tilde{\Psi}(f(t,y)) = \frac{\tilde{\Psi}(f(t,y))}{f(t,y)}\ f(t,y) \le \frac{K_0^{\vartheta+2}}{K_0}\ f(t,y) \le K_0^{\vartheta+1}\ f(t,y)\,.
\end{equation}
\end{itemize}
Consequently, recalling that $K_0>1$ and $\Psi(0)<0$, we deduce from (\ref{y12}) and (\ref{jeden})-(\ref{trzy}) that
\begin{eqnarray*}
\frac{dm_q}{dt} & \le & q(q-1)\ \int_0^M y^{q-2} \tilde{\Psi}(f)\ \mathbf{1}_{(0,\varepsilon_M]}(f)\ dy + q(q-1)\ \int_0^M y^{q-2} \tilde{\Psi}(f)\ \mathbf{1}_{(\varepsilon_M,K_0)}(f)\ dy \\
& + & q(q-1)\ \int_0^M y^{q-2} \tilde{\Psi}(f)\ \mathbf{1}_{[K_0,\infty)}(f)\ dy + M m_q - \frac{M^{q+1}}{q+1} \\
& \le & \frac{(q-1)M^2}{2(q+1)}\ \int_0^M y^{q-2}\ dy + q(q-1)\ \frac{\gamma_\vartheta K_0^{\vartheta+1} - \Psi(0)}{\varepsilon_M}\ \int_0^M y^{q-2} f\ dy \\
& + & q(q-1) K_0^{\vartheta+1}\ \int_0^M y^{q-2} f\ dy + M m_q - \frac{M^{q+1}}{q+1} \\
& \le & C_2\ K_0^{\vartheta+1}\ \int_0^M y^{q-2} f\ dy + M m_q - \frac{M^{q+1}}{2(q+1)}\,,
\end{eqnarray*}
with $C_2:= q(q-1)(\gamma_\vartheta-\Psi(0)+\varepsilon_M)/\varepsilon_M$. We next use H\"older's inequality and (\ref{mas}) to conclude that
\begin{equation}\label{jules}
\frac{dm_q}{dt} \le C_2\ K_0^{\vartheta+1}\ m_q^{(q-2)/q} + M m_q - \frac{M^{q+1}}{2(q+1)}\,.
\end{equation}
It remains to estimate $K_0$ and in fact $\mathcal{L}_1(f_0)$. Since $\Psi$ is negative on $(0,1)$ and $\Psi_1$ is bounded from below by $\Psi_1(0)$, it follows from (\ref{gaston}) and (\ref{pam}) that
\begin{eqnarray*}
\mathcal{L}_1(f_0) & \le & \frac{2}{\delta^4}\ (1-M\delta^q)^2\ \int_0^\delta |\Psi'(f_0)|^2\ dy + \int_0^\delta \Psi(f_0)\ dy - M^2 \Psi_1(0) \\
& \le & \frac{2}{\delta^4}\ \int_0^\delta |\Psi'(f_0)|^2\ dy + \int_0^\delta \Psi(f_0)\ dy - M^2 \Psi_1(0) .
\end{eqnarray*}
On the one hand, we infer from (\ref{poum}), (\ref{y12b}), and the monotonicity of $\Psi$ that
$$
\int_0^\delta \Psi(f_0)\ dy \le \delta\ \Psi\left( \frac{2}{\delta} \right) \le \gamma_\vartheta 2^{\vartheta+1}\ \delta^{-\vartheta}\,.
$$
On the other hand, we have
\begin{eqnarray*}
f_0(y) & \ge & 1 \;\;\mbox{ for }\;\; y\in [0,y_\delta] \;\;\mbox{ with }\;\; y_\delta := \delta - \frac{1-\delta^q}{2(1-M\delta^q)} \delta^2>0\,, \\
f_0(y) & \in & [\delta^q,1] \;\;\mbox{ for }\;\; y\in [y_\delta,\delta]\,,
\end{eqnarray*}
so that, if $y\in [0,y_\delta]$,
$$
\Psi'(f_0(y))^2 \le \gamma_\vartheta^2 f_0(y)^{2\vartheta} \le \gamma_\vartheta 4^\vartheta\ \delta^{-2\vartheta}
$$
by (\ref{poum}) and (\ref{y12b}), while, if $y\in (y_\delta,\delta]$,
$$
\Psi'(f_0(y))^2 \le \frac{1}{f_0(y)^4} a\left( \frac{1}{f_0(y)} \right)^2 \le C_\infty^2\ f_0(y)^{2(\alpha-2)} \le C_\infty^2\ \delta^{-2q(2-\alpha)}
$$
by (\ref{decr}) since $\alpha\le 2$. Therefore,
\begin{eqnarray*}
\mathcal{L}_1(f_0) & \le & \frac{2}{\delta^4}\ \left[ \int_0^{y_\delta} \gamma_\vartheta 4^\vartheta\ \delta^{-2\vartheta}\ dy + \int_{y_\delta}^\delta C_\infty^2\ \delta^{-2q(2-\alpha)}\ dy \right] + \gamma_\vartheta 2^{\vartheta+1}\ \delta^{-\vartheta} - M^2 \Psi_1(0)\\
& \le & \gamma_\vartheta 4^{\vartheta+1}\ \delta^{-3-2\vartheta} + C_\infty^2\ \frac{1-\delta^q}{2(1-M\delta^q)}\ \delta^{-2-2q(2-\alpha)} + \gamma_\vartheta 2^{\vartheta+1}\ \delta^{-\vartheta} - M^2 \Psi_1(0)\\
& \le & \gamma_\vartheta 4^{\vartheta+1}\ \delta^{-2(2+\vartheta)} + C_\infty^2\ \delta^{-2-2q(2-\alpha)} + \gamma_\vartheta 2^{\vartheta+1}\ \delta^{-\vartheta} - M^2 \Psi_1(0) \\
& \le & C_3\ \left( \delta^{-2(2+\vartheta)} + \delta^{-2-2q(2-\alpha)} \right)
\end{eqnarray*}
with $C_3:= \gamma_\vartheta 4^{\vartheta+2} + C_\infty^2 - M^2 \Psi_1(0)$. Therefore,
\begin{equation}\label{cubitus}
K_0^{\vartheta+1} \le C_4\ \left( \delta^{-(\vartheta+1)} + \delta^{-(\vartheta+1)(1+q(2-\alpha))/(\vartheta+2)} \right)
\end{equation}
for some constant $C_4>0$ depending only on $M$ and $a$.
Combining (\ref{jules}) and (\ref{cubitus}) yields
\begin{equation}\label{senechal}
\frac{dm_q}{dt} \le \Lambda_\delta(m_q) := C_5\ \left( \delta^{-(\vartheta+1)} + \delta^{-(\vartheta+1)(1+q(2-\alpha))/(\vartheta+2)} \right)\ m_q^{(q-2)/q}\ + M m_q - \frac{M^{q+1}}{2(q+1)}
\end{equation}
for $t\in [0,T_{max})$ and some constant $C_5>0$ depending only on $M$ and $a$. At this point, we note that the monotonicity of $\Lambda_\delta$ and (\ref{senechal}) imply that $\Lambda_\delta(m_q(t))\le \Lambda_\delta(m_q(0))$ for $t\in [0,T_{max})$ if $\Lambda_\delta(m_q(0))<0$, the latter condition being satisfied for $\delta$ small enough as
$$
\Lambda_\delta(m_q(0)) \le C_1^{(q-2)/q} C_5\ \left( \delta^{q-3-\vartheta} + \delta^{(q(\alpha(\vartheta+1)-\vartheta)-3\vartheta-5)/(\vartheta+2)} \right) + M C_1\ \delta^q - \frac{M^{q+1}}{2(q+1)}
$$
by (\ref{prunelle}) and (\ref{boule}).
Summarizing, we have shown that, if $\delta$ satisfies (\ref{gaston}) and
\begin{equation}\label{semaphore}
C_1^{(q-2)/q} C_5\ \left( \delta^{q-3-\vartheta} + \delta^{(q(\alpha(\vartheta+1)-\vartheta)-3\vartheta-5)/(\vartheta+2)} \right) + M C_1\ \delta^q < \frac{M^{q+1}}{2(q+1)}\,,
\end{equation}
we have
$$
\frac{dm_q}{dt}(t) \le \Lambda_\delta(m_q(t)) \le \Lambda_\delta(m_q(0))<0\,, \quad t\in [0,T_{max})\,,
$$
an inequality which can only be true on a finite time interval owing to the non-negativity of $m_q$. Therefore, $T_{max}<\infty$ in that case and, for any $M>0$, we have found an initial condition $u_0$ given by (\ref{zero}), (\ref{odw}), and (\ref{pam}) (for $\delta$ small enough according to the above analysis) such that $\langle u_0\rangle=M$ and the first component $u$ of the corresponding solution to (\ref{he1})-(\ref{he3}) blows up in finite time. \qed
\section{Global existence.}\label{ge}
The proof of Theorem~\ref{mainth}~(ii) also relies on the study of the function $L_1$ defined in Lemma~\ref{lem}. For that purpose, we first recall another property from \cite{cl3}. We define the function $E_1$ by
\begin{equation}\label{landau}
E_1(h) := \frac{1}{2} \|\partial_y h\|_2^2 + \int_0^M \mathbf{1}_{(-\infty,0)}(h(y))\ h(y)\ dy\,, \quad h\in H^1(0,M)\,,
\end{equation}
for which we have the following lower bound.
\begin{Le}\cite[Lemma 9]{cl3}\label{lab1}
For $M>0$, we have
\begin{equation}\label{gex5}
E_1(h) \ge \frac{1}{4}\ \|\partial_y h\|_2^2 - M^3 - M \left| \Psi\left( \frac{1}{M} \right) \right|\,,
\end{equation}
and
\begin{equation}\label{gex6}
\|h\|_1\leq M^{3/2} \|\partial_y h\|_2 + M \left| \Psi\left( \frac{1}{M} \right) \right|
\end{equation}
for every $h\in H^1(0,M)$ satisfying
\begin{equation}
\label{gex7}
\int_0^M \Psi^{-1}(h)(y)\ dy = 1\,.
\end{equation}
\end{Le}
We now show that the non-integrability of $a$ at infinity allows us to show that $T_{max}=\infty$. To this end, we use the alternative formulation (\ref{main})-(\ref{init}) as in \cite{cl3} and prove that $f$ cannot vanish in finite time.
\vspace{0.3cm}
\noindent
\textit{Proof of Theorem~\ref{mainth}~(ii).} Owing to (\ref{init}) and the assumptions made on $u_0$, we have
\[
0<f_0(y)\leq \frac{1}{m_0}\,, \quad y\in [0,M].
\]
Introducing $\Sigma(t):=M^{-1}+e^{Mt}\left(m_0^{-1}-M^{-1}\right)$ for $t\geq 0$ we have
\[
\partial_t\Sigma-\partial_y^2\Psi(\Sigma)-M\Sigma+1=M\left(\Sigma- \frac{1}{M}\right)-M\Sigma+1=0,
\]
\[
\Sigma(0)=\frac{1}{m_0}\geq f_0(y), \quad y\in(0,M),
\]
and the comparison principle warrants that
\begin{equation}\label{por}
f(t,y)\leq\Sigma(t),\quad (t,y)\in [0,T_{max})\times[0,M].
\end{equation}
We now follow the strategy of the proof of \cite[Theorem 5]{cl3} and first use the properties of $\Psi$, $\Psi_1$, and (\ref{por}) to estimate the function $L_1$ defined in Lemma~\ref{lem} from below. Indeed, since $\Psi\geq 0$ on $(1,\infty)$ and $\Psi_1\leq 0$ on $(0,1)$ we arrive at
\begin{eqnarray*}
L_1(0)\ge L_1(t) &=& \frac{1}{2}\left\|\partial_y\Psi(f(t))\right\|_{L^2(0,M)}^2+\int_0^M \mathbf{1}_{(0,1)}(f(t,y))(\Psi-M\Psi_1)(f(t,y))\ dy \\
& + & \int_0^M \mathbf{1}_{(1,\infty)}(f(t,y))(\Psi-M\Psi_1)(f(t,y))\ dy \\
& \geq & \frac{1}{2}\left\|\partial_y\Psi(f(t))\right\|_{L^2(0,M)}^2+\int_0^M{\bf 1}_{(-\infty,0)}(\Psi(f(t,y)))\Psi(f(t,y))\ dy \\
& - & M\int_0^M{\bf 1}_{(1,\infty)}(f(t,y))\Psi_1(f(t,y))\ dy \\
& \geq & E_1(\Psi(t))-M^2\Psi_1(\Sigma(t)),
\end{eqnarray*}
where $E_1$ is defined in (\ref{landau}) and we have used (\ref{por}) to obtain the last inequality. Next, by Lemma~\ref{lab1} and (\ref{mas}), we have
$$
L_1(0)\geq \frac{1}{4}\left\|\partial_y\Psi(f(t))\right\|_{L^2(0,M)}^2-M^3-M\left|\Psi\left(\frac{1}{M}\right)\right|-M^2\Psi_1(\Sigma(t)),
$$
whence
\begin{equation}\label{prandtl}
\frac{1}{4}\left\|\partial_y\Psi(f(t))\right\|_{L^2(0,M)}^2\leq L_1(0)+M^3+M\left|\Psi\left(\frac{1}{M}\right)\right|+M^2\Psi_1(\Sigma(t)).
\end{equation}
Using again Lemma~\ref{lab1}, we have
\begin{eqnarray*}
\left\|\Psi(f(t))\right\|_{L^1(0,M)} & \leq & M^{3/2}\left\|\partial_y\Psi(f(t))\right\|_{L^2(0,M)}+M\left|\Psi\left(\frac{1}{M}\right)\right| \\
& \leq & 2M^{3/2}\left(L_1(0)+M^3+M\left|\Psi\left(\frac{1}{M}\right)\right|+M^2\Psi_1(\Sigma(t))\right)^{1/2}+M\left|\Psi\left(\frac{1}{M}\right)\right|.
\end{eqnarray*}
Combining the previous inequality with (\ref{prandtl}) and the Poincar\'e inequality leads us to the bound
\begin{equation}\label{last}
\left\|\Psi(f(t))\right\|_{H^1(0,M)}\leq C_6(T)\,, \quad t\in [0,T]\cap [0,T_{max})\,,
\end{equation}
for all $T>0$. Together with the continuous embedding of $H^1(0,M)$ in $L^\infty(0,M)$, (\ref{last}) gives
\[
-C_7(T)\leq\Psi(f(t,y))\leq C_7(T)\,, \quad (t,y)\in ([0,T]\cap [0,T_{max})) \times [0,M]\,.
\]
Since
$$
\lim_{r\rightarrow 0}\Psi(r)=-\infty
$$
due to $a\not\in L^1(1,\infty)$, the above lower bound on $\Psi(f)$ ensures that $f(t)$ cannot vanish in finite time, from which Theorem~\ref{mainth}~(ii) follows as already discussed in section~\ref{pre}. \qed
\noindent\textbf{Acknowledgements.} This work was done while the first author held a post-doctoral position at the University of Z\"{u}rich and is also partially supported by the Polish Ministry of Science and Higher Education under grant number NN201 396937 (2009-2012).
|
\section{Introduction}
Gamma-ray bursts (GRBs) are extremely bright and highly variable
sources of gamma-ray radiation. Their isotropic equivalent emitted
energy can be as large as $10^{55}$~erg (GRB080916C; Abdo et
al. 2009). GRBs last between a fraction of a second and several
thousand seconds and can be divided in two classes based on their
duration and spectral characteristics (Kouveliotou et al. 1993). Short
bursts last less than 2~s, while long bursts last between 2 and
several thousand seconds. Even though some bursts seem hard to
classify within this simple scheme (e.g., GRB~060505 and GRB~060614;
Fynbo et al. 2006, Gal-Yam et al. 2006, Della Valle et al. 2006;
GRB~060121, de Ugarte Postigo et al. 2006) it is widely believed that
a substantial fraction of, if not all, long-duration GRBs are
associated with the death of massive, rapidly spinning stars (Woosley
1993; MacFadyen \& Woosley 1999; Stanek et al. 2003; Hjorth et
al. 2003; Woosley \& Bloom 2006).
Within the overall duration of the prompt phase, fast variability is
commonly observed, with the shortest spikes lasting only a fraction of
a millisecond (Schaefer \& Walker 1999, Walker et al. 2000).
Characterizing the burst variability has proved challenging, with each
GRB seeming to have its own individual pattern. This is particularly
frustrating since the variability can in principle carry information
on the workings of the central engine, the energy dissipation
processes, and the radiation mechanisms involved in the release of the
burst emission. Fenimore \& Ramirez-Ruiz (1999) showed that the
variability time scale of GRB light curves does not evolve from the
beginning to the end of the prompt emission. Their work placed a
strong constraint on the dissipation process and proved that the
variability in the GRB prompt emission is not due to the interaction
of the fireball with the external medium (see Dermer et al. 1999; 2000
for an alternative interpretation). Beloborodov et al. (1998, 2000)
analyzed the composite power spectrum of BATSE light curves, finding
that it is well described by a power-law of the shape $PDS(f)\propto
f^{-5/3}$, reminiscent of the Kolmogorov spectrum of fully developed
turbulence.
The discovery of the association of GRBs with massive stars puts burst
variability in a new light. There are two possible sources of
variability in the outflow: the engine itself (MacFadyen \& Woosley
1999; Aloy et al. 2000; Ouyed et al. 2003; Proga et al. 2003; McKinney
\& Narayan 2007; McKinney \& Blandford 2009) and the interaction of
the flow with the progenitor star material (Aloy et al. 2002; Gomez \&
Hardee 2004; Aloy \& Obergaulinger 2007; Morsony et al. 2007). How
these two sources of variability combine and interact to give rise to
the variability in the light curve is unknown but of fundamental
importance if any conclusion is to be drawn from the temporal
properties of GRB light curves.
In this paper we present the results of 2D axisymmetric simulations of the
propagation of baryonic GRB jets through their progenitor star
material. We show simulations in which the engine is either constant
or variable and compare the structure of the jets as they emerge from
the progenitor star. This paper is organized as follows: in \S~2 we
describe our simulations, in \S~3 we describe our results, and in \S~4
we discuss their potential implications.
\section{The simulations and the injected variability}
We present the results of four simulations, each of which has the same
progenitor star and average jet properties, but with different
injected luminosity histories. The first simulation has uniform
injected luminosity and is analogous to that already presented in
Morsony et al. (2007, hereafter MLB07) and Lazzati et al. (2009
hereafter LMB09). We call this simulation {\it uniform}. The second
and third simulations ({\it variable entropy} and {\it variable baryon
load}) represent an inner engine that releases a luminosity that varies
in time. The time history was simulated as a random series of numbers
convolved with a Gaussian, to yield a flat power spectrum (white
noise) out to a cutoff at $\sim0.1$s.
This represents a randomly varying energy input, with a cutoff such that all the variability is well resolved.
The energy input in a real GRB could, of course, have a different power spectrum, but for our purpose of determining how the variability is modified during propagation, this input is an adequate example.
The difference between the
simulation that we call {\it variable entropy} and the one that we
call {\it variable baryon load} is that in the former the luminosity
is varied by changing the entropy $\eta=L/\dot{m}c^2$ holding
$\dot{m}$ constant, while in the latter the luminosity is changed by
changing the baryon load ${\dot{m}}$. Finally, a simulation was run
with an on-off engine with a period of 0.2s (0.1s on, 0.1s off), in
which the luminosity was changed by changing the baryon load ${\dot{m}}$. We identify
this last simulation as {\it step}.
The represents an extreme case of strong, fast variability but on a time scale that is well resolved in our simulation.
A 3s section of the injected luminosity for the four simulations
is shown in Figure~\ref{fig:lum}.
In all simulations, the progenitor star is model
16TI from Woosley \& Heger (2006, see its density profile in
Fig.~\ref{fig:16ti}), the jet is injected at a distance
$R_0=10^{9}$~cm from the center of the star and has an initial opening
angle $\theta_0=10^\circ$ and an initial Lorentz factor
$\Gamma_0=5$. In all simulations self-gravity is neglected, and
therefore the star expands slightly under the effect of its internal
pressure during the 50 seconds of our simulation. The expansion is
however negligible (less then 1\% of the radius) and does not affect
the dynamics of the jet. In all simulations the surrounding medium is
uniform, with a density $\rho=10^{-9}$~g~cm${-3}$ (for the big box
simulation discussed below, the density was reduced to
$\rho=10^{-13}$~g~cm${-3}$, LBM09). Further simulations
will be performed to study the effect, if any, of a wind environment
on the jet evolution outside the stellar progenitor.
The simulations were performed in 2D cylindrical coordinates with the adaptive mesh refinement (AMR)
special relativistic hydrodynamic code FLASH (Fryxell et al. 2000; see
MLB07 for extensive testing of the special relativistic module), as
modified by the authors (MLB07). The simulations were carried out for
50 seconds for a maximum number of refinements of 13 in the stellar
core, corresponding to a maximum resolution of
$7.8125\times10^{6}$~cm. Outside $5\times10^{9}$~cm from the center
of the star, the maximum resolution is $3.125\times10^{7}$~cm,
corresponding to 11 levels of refinement. This is 4 times the
resolution used in MLB07 and LMB09.
The temporal resolution of these simulations is $1/100$th of a second.
Figure~\ref{fig:panels} shows
logarithmic density contours for the four simulations at $t=5.5$~s. In
all panels, comoving density is shown. In all cases, a very narrow
low-density jet is visible surrounded by a highly structured,
turbulent cocoon of shocked stellar material.
\section{Results}
\subsection{Effect of a variable input on the jet propagation inside the
star}
We first concentrate on the way in which the jets propagate through
the material of their stellar progenitor. Extensive work on this topic
has been done in the past (MacFadyen \& Woosley 1999; Aloy et
al. 2000; MacFadyen et al. 2001; Zhang et al. 2003, 2004; MLB07,
Mizuta \& Aloy 2009) but only in one case (Aloy et al. 2000) were
variable engines considered. Aloy et al. concluded that a variable
engine has a faster propagation time through the stellar progenitor.
Figure~\ref{fig:jh} shows the propagation of the jet head through the
progenitor star for our four simulations. The jet head was identified
as the location of the bow shock where the outward velocity of the
material exceeds 1\% of the speed of light along the jet
axis. Contrary to previous results, we find that the shortest
propagation time is observed for a uniform injected luminosity, with
the head of the jet breaking out of the star's surface at about 6.2s,
corresponding to an average velocity $v_{\rm{jh}}=0.21c$. The {\it
step} simulation has a slightly longer propagation time of 6.25~s,
while the {\it variable entropy} and {\it variable baryon load}
simulations have longer breakout times of 6.8 and 7.1~s,
respectively. It is not straightforward to understand the origin of
the difference between our findings and those of Aloy et al. (2000).
One possibility is that thanks to the AMR capabilities of the FLASH
code, the resolution of our simulation is higher, especially as the
jet approaches the stellar surface. On the other hand,
Figure~\ref{fig:jh} shows that the effect of the engine variability is
not simple. In the core of the star, out to approximately 20 per cent
of its radius, the {\it uniform}, {\it variable entropy}, and {\it
variable baryon load} jets propagate at the same speed, while the
{\it step} simulation (the one most similar to the Aloy et al. 2000
set up) is ahead. At larger radii, the behaviors differentiate, with
the uniform simulation showing the fastest velocity and reaching the
stellar surface first. It is therefore possible that the difference
between our result and the result of Aloy et al. (2000) is due to a
different structure of the progenitor star. Alternatively, the
difference may be due to the different prescription used in Aloy et
al. (2000) for the jet injection: while we injected an outflow with
net momentum as a boundary condition, Aloy et al. (2000) injected
energy with no net momentum in a conical region in the core of the
star
, allowing a jet to develop rather than be injected.
The lower boundary of the grid used by Aloy et al. (2000) is also much farther inside the star, at $2\times10^{7}$~cm rather than $10^{9}$~cm. Differences in jet propagation may arise in this inner region that our simulations are unable to capture.
Finally, it is possible that the speed of the jet head depends on
the duration of the on and off phases or the frequency at which the
variability is injected. The difference between our result and Aloy et
al. (2000) may simply be due to the fact that we investigate
variability at different frequencies.
The complexity and the sensitivity to details of the interaction of
the jet head with the star is also shown in Figure~\ref{fig:sh}. The
figure shows the Lorentz factor of the jet material along the jet axis
at the moment of breakout for the four simulations. The {\it uniform}
simulation has the most regular structure, with strong recollimation
shocks separated by acceleration phases in which the Lorentz factor
grows linearly with radius. All the {\it variable} simulations show a
more complex evolution of the Lorentz factor, with milder but more
numerous shocks. In particular the {\it step} simulation shows a very
complex evolution of the Lorentz factor in which strong recollimation
shocks are absent. This is probably due to the fact that a strong
shock cannot be long lived since it is lost when it propagates from a
low luminosity phase to a high luminosity one.
\subsection{Variability properties of the jets outside the star \label{section3.2}}
While the propagation of the jet inside the progenitor star is
strongly dependent on details, the jet emerging at the surface of the
star and propagating out of it is fairly insensitive to those same
details. Figure~\ref{fig:oa} shows the evolution of the opening angle
of the jet measured at a distance of $R=10^{11}$~cm from the center of
the progenitor star (or at 2.5 stellar radii). The opening angle is
measured as the geometric angle of the plasma parcel moving at
$\Gamma\ge10$ that is located farthest from the jet axis. As described
in MLB07, the opening angle is initially very large, due to the
quasi-isotropic cocoon material (see also Lazzati et al. 2010). Such a
phase, visible in Figure~\ref{fig:oa} as the spike in angle at
$t\sim10$~s, is followed by the shocked jet phase, in which the
opening angle of the jet is roughly constant and about half the size
at injection. This is the phase in which the jet interaction with the
progenitor star is the strongest and lasts, in all cases,
approximately 10 to 15 seconds. After that, the interaction with the
star becomes weak, and the jet gradually expands to its injection
opening angle\footnote{Note that if an injection opening angle is not
assumed, such as in Aloy et al. 2000, the asymptotic opening angle
is set by the jet and progenitor properties.}
$\theta_0=10^\circ$. Since the jet is injected with a moderate lorentz
factor $\Gamma_0=5$, the opening angle eventually expands to
$\theta_{\rm{lim}}=\theta_0+1/\Gamma_0\sim21\degr$ (MLB07).
Even though minor differences are present, all four simulations follow
the same qualitative trend, with the three phases clearly identified
and lasting approximately the same time, and with approximately the same opening angle evolution.
This indicates that the structure of the jet that emerges from the progenitor star is similar regardless of the short-timescale variability of the energy injection.
\subsection{Light-power curves}
Another way in which the structure of the jet outside the star can be
investigated is through the calculation of light-power curves. The
light-power curve at a radius $R$ is computed as:
\begin{equation}
L_R(t)=c \int_{\Sigma_R} d\sigma \left[(4p+\rho c^2) \Gamma^2 -\rho
c^2 \Gamma\right]\delta^2
\label{eq:lum}
\end{equation}
where $\Sigma_R$ is a spherical surface of radius $R$ centered on the
GRB engine, $p$ and $\rho$ are the pressure and comoving density of
the jet, respectively, and $\delta=[\Gamma(1-\beta\cos\theta)]^{-1}$
is the Doppler factor. The luminosity in Eq.~\ref{eq:lum} is formally
the luminosity that the observer would detect if all the energy were
released as radiation at the radius $R$ with a 100\% efficiency. The
light-power curves have properties of light curves, since they contain
the Doppler factor to beam photons in the direction of motion, and of
power-curves, since they assume 100\% efficiency and are computed at a
fixed radius. In addition, we included in the computation only
material that has a minimum Lorentz factor
$\Gamma_{\infty,\rm{min}}=10$, since slower material is not supposed
to contribute to the keV-MeV prompt light curve of GRBs.
Using the light power curve as a proxy for the light curve is a rough
approximation of the processes that lead to the generation of the
light curve in a GRB outflow. However, given the impossibility of
following the propagation of the jet in 2D out to the radiative
radius, an assumption has to be made. The rationale behind our choice
is that the brightest spikes in the light curve are produced by the
most energetic parts of the outflow, and the duration of a spike
should be correlated to the radial thickness of the energetic part of
the outflow. Such close connection is true for the photospheric
scenario for the prompt emission (Rees \& Meszaros 2005; LBM09) and
for the internal shocks scenario, at least in first approximation
(Kobayashi et al. 2007). In other words, we believe our light power
curves have the right number of pulses with correct durations. If,
however, we analyze the details of the light curve of individual
pulses, our approximation is likely to break down, since the spectral
evolution is completely missed in our approximation (see, e.g., Daigne
\& Mochkovitch 2002). In addition, it has been shown that second
generation shells - shells produced by collisions between other shells
- are complex and would produce more complex pulses than original
shells (Mimica et al. 2005, 2007). So, we expect that our
approximation does miss some of the fine structure in the light curves,
but we also expect the overall qualitative behavior to be accurate.
Figure~\ref{fig:lc1} shows the light-power curve computed at a radius
$R=2.5\times10^{11}$~cm (the edge of our simulation box\footnote{To
make sure the light-power curves are not affected by edge effects,
we computed also light power curves at a radius
$R=2\times10^{11}$~cm, obtaining fully consistent results}.) for the
{\it uniform} and the {\it variable baryon load} simulations. Several
important conclusions can be drawn from the figure. First, even if the
central engine releases a constant luminosity, the power curve shows
marked variability as a result of the interaction between the jet and
the progenitor star. This is not a new result (MacFadyen \& Woosley
1999; Aloy et al. 2000, 2002; Zhang et al. 2003, 2004; MLB07;
LMB09). The comparison between the two curves shows, however, a more
interesting and new conclusion. The characteristics of the long
time-scale variability are similar in the two light curves, with an
initial phase characterized by peaks with a width of several seconds
followed by a broad bump with a duration of approximately 20
seconds. Some differences are however apparent, with a second peak
appearing in the {\it variable baryon load} light power curve that is
not observed in the {\it uniform} simulation.
Another interesting comparison is presented in Figure~\ref{fig:lc2},
which shows the luminosity injected at the base of the jet (red line)
compared to the light-power curve measured well outside the star at
radius $R=2.5\times10^{11}$~cm for the same {\it variable baryon load}
simulations shown in Fig.~\ref{fig:lc1}. The injected curve has been
shifted in time by adding the light propagation time from the inner
boundary of the simulation out to the radius at which the light power
curves have been computed. In this way, the jet propagation time was
removed. The left panel shows the comparison during the first three
seconds of the light-power curve while the right panel shows the same
comparison at a later time, still during the shocked jet phase
(according to the definition of MLB07). In the left panel, fast
variability is observed in the light-power curve, but it is almost
completely uncorrelated with the injected variability at the base of
the jet. On the other hand, at the later time shown in the right
panel, the variability of the engine and that observed in the
light-power curve are almost identical. A further analysis shows that
the strong correlation appears at about $\sim4$~s from the trigger and
lasts until $t=30$~s, the beginning of the unshocked jet phase
(MLB07). The same analysis was performed with the {\it variable
entropy} and the {\it step simulations}. We found that in both cases
the strong correlation is observed. However, the {\it variable
entropy} light power curve is sometimes shifted in time, likely due
to different propagation times of shells with different entropy and
therefore different asymptotic Lorentz factor. In the {\it step}
simulations, instead, the strong correlation is observed at all times,
likely due to the unphysically strong character of the fluctuations
injected in that model.
This suggests that there are two sub-phases within the shocked-jet
phase. An initial short phase, lasting a few seconds, has the
capability of modifying the variability injected by the inner engine,
while a second phase, lasting several tens of seconds until the end of
the shocked-jet phase, leaves the engine variability unaffected. A
closer inspection of our simulations reveals that the initial short
phase is characterized by perpendicular shocks, propagating along the
jet similarly to internal shocks, while the latter phase is
characterized by tangential shocks that propagate from the sides to
the jet axis and vice versa. The fact that the variability injected by
the central engine is preserved during the shocked jet phase comes as
a surprise. Both Zhang et al. (2003) and MLB07 had predicted, based on
simulations of constant engines, that the variability injected by the
central engine, if any, would have been visible only ``at very late
times when the star has exploded'' (Zhang et al. 2003). We find here,
from the analysis of a simulation with a variable inner engine, that
the transverse collimation shocks are not able to erase the
short-timescale (fraction of a second) variability injected at the
base of the jet by the inner engine. This finding makes us much more
optimistic that the characteristics of the central engine can be
studied by observing the light curves of GRBs. Further analysis is
however needed to confirm the robustness of this result for different
progenitors, injected luminosities, and jet properties.
The inner engine variability is also visible during the unshocked jet
phase, as described in MLB07. This final phase, however, is much
dimmer than the shocked jet phase and it is unclear whether it could
be observed at all in GRB light curves.
\subsection{Polar jet structure}
The computation of light-power curves allows us to compare the amount
of energy and the peak luminosity as a function of the off-axis angle
$\theta_o$. Previous studies on the distribution of energy in the
fireball with respect to $\theta_o$ have resulted in controversial
results. Zhang et al. (2004) found a somewhat universal slope
$dE/d\Omega\propto\theta^{-3}$, while MLB07 and Mizuta \& Aloy (2009)
found that the slope depends on the properties of the stellar
progenitor, and on the limiting Lorentz factor $\Gamma_{\rm{min}}$
that is adopted. In this paper, we compute the value of $dE/d\Omega$
from the light-power curves, rather than computing it locally from the
kinetic energy in the flow. As a result, the values of $dE/d\Omega$
that we obtain are smoother
than those in MLB07
because they include contributions from
material pointing in different directions with respect to the line of
sight.
This is similar to the method used in Mizuta \& Aloy (2009) for including contributions from material away from the line of sight.
Each panel of Figure~\ref{fig:azi} shows the results from one of our
simulations. We find that the isotropic equivalent energy distribution
$dE/d\Omega$ drops with angle smoothly in all cases, with a slope well
represented by a $\theta^{-3}$ power-law, analogously to that found by
Zhang et al. (2004). Only for the {\it step} simulation, a deviation
is observed at $\theta_o\sim20^{\circ}$ off-axis.
Mizuta \& Aloy (2009) found that the slope ranges from -2.7 for high mass progenitors to -3.7 for low mass progenitors. However, the progenitor model in their paper most similar to the model we use is model 16OC (from Woosley \& Heger 2006). For this model, Mizuta \& Aloy (2009) find a
slope of -2.95, consistent with our results.
Figure~\ref{fig:azi}
also shows the isotropic equivalent peak luminosity distribution
$dL_{\rm{pk}}/d\Omega$, characterized by a much more varied
behavior. At angles less than $\theta_o\sim5^\circ$, the peak
light-power curve tracks the isotropic equivalent energy curve. At
larger angles, however, it becomes roughly constant, while the
isotropic equivalent energy keeps dropping. The peak luminosity
remains constant out to an angle that depends on the simulation, but
is roughly a few tens of degrees; it eventually drops for very large
off-axis angles.
The angular dependence of the peak luminosity curve is also much more
sensitive than that of the isotropic energy curve to details of the
variability model. For small off-axis angles ($\theta_o<5^\circ$) ,
increasing the angle decreases the luminosity throughout the duration
of the burst. For larger off-axis angles, however, the loss of total
energy is compensated by the fact that the luminosity is concentrated
into a shorter time (usually the first pulse in the light-power
curve), so that the peak luminosity is kept constant. To check this,
we computed two measures for the duration of the prompt emission,
$T_{50}$ and $T_{1/2}$. $T_{50}$ is the time during which 50\% of
the total energy is observed, while $T_{1/2}$ time is the time during
which the light curve (in our case the light-power curve) has a
luminosity greater than one half of the peak luminosity. These
quantities are shown in Figure~\ref{fig:t90}. It is interesting to
note that $T_{50}$ and $T_{1/2}$ have very different behaviors,
demonstrating the extent to which the variability properties of the
light-power curves change with observing angle.
\subsection{Power density spectra}
To study the variability of GRB light curves more quantitatively, we
compute the power density spectrum (PDS), i.e., the square modulus of
the Fourier transform of the signal. The PDS of BATSE GRB light curves
is characterized by a power-law slope $PDS(f)\propto f^{-5/3}$ between
cutoffs at $f\sim0.01$~Hz and $f\sim1$~Hz
(Beloborodov et al. 1998, 2000).
In this section we introduce a new simulation, that we call {\it
extended uniform}, which is analogous to the {\it uniform}
simulation but has $1/4$th the resolution ($1.25\times10^{8}$~cm) at
the stellar surface and extends to a much larger radius:
$R=2.5\times10^{12}$~cm. This is the simulation used by LMB09 to
derive their on-axis photometric light curve. Here we use the
simulation to compute light-power curves at various off-axis angles at
a radius $R=2.5\times10^{12}$~cm, closer to the radius at which the
radiation is actually released.
At this radius, the typical spatial resolution is $2\times10^{9}$~cm, although it is allowed to be as high as $5\times10^{8}$~cm.
The temporal resolution of this simulation is $1/15$th of a second, giving a maximum resolved frequency of $7.5$~Hz.
Light-power curves were computed at
off-axis angles $\theta_o=1$, 2, 3, 4, 5, 6, 7, 8, 9, and
$10\degr$. For each of these curves, the power density spectrum was
computed and the 10 spectra were averaged and binned to increase the
signal-to-noise ratio.
The resulting PDS is shown in Figure~\ref{fig:pds1}. The spectrum has
been multiplied by $f^{5/3}$ to emphasize the slope of the power-law
and in an attempt to reproduce, as closely as possible, Fig. 2 of
Beloborodov et al. (1998), whose data are shown as a thin line in the
figure for comparison. The frequencies of Beloborodov et al. (1998)
have been multiplied by a factor 2 to take into account the average
redshift of BATSE GRBs. The two spectra show a remarkable qualitative
similarity. Both spectra are consistent with a power-law slope
$PDS(f)\propto f^{-5/3}$ between a low- and a high-frequency cutoff.
The locations of the cutoffs are also compatible with observations for
the assumed average redshift of BATSE GRBs.
The rough agreement of the low-frequency cutoff is not surprising
since it is due to the overall duration of the burst, which has been
set in the simulations to be analogous to the average observed GRB
duration. The similarity between the observed and simulated
high-frequency cutoffs is more intriguing,
although it could be due to either numerical effects or the evolution of the jet.
The time scale associated
with the $\sim3$~Hz cutoff in the synthetic spectra is $\delta
t\sim0.33$~s. The jet at the surface of the star has an opening angle
of $\sim4^\circ$ (see Figure~\ref{fig:oa}) which corresponds to a
transverse size $R_\perp=2.8\times10^9$~cm. For a relativistic sound
speed $c_s=c/\sqrt{3}$, the transverse dimension corresponds to a
crossing time of 0.16~s.
One possible interpretation is that the cutoff at $\sim3$~Hz can be
identified as the signature of the jet-crossing time of a disturbance
that propagates from the side to the axis of the jet, such as a
recollimation shock.
The difference between the $\sim6$~Hz cutoff predicted and the $\sim3$~Hz
cutoff seen in our simulations could be because either the ``real''
opening angle of the jet may be large than the value we use based on
our definition in Sect. \ref{section3.2}, or the characteristic radius
at which a shock is crossing the jet may be larger than the original
stellar radius.
It is also possible that the high-frequency cutoff is a numerical artifact.
The cutoff in our data is at $\sim3$~Hz cutoff, while the maximum
frequency we can sample in the {\it extended uniform} simulation
is $7.5$~Hz, close enough that it is possible the cutoff is a result
of resolution effects in the simulation.
The consistency of the observed PDS slope with the synthetic one is,
however, more challenging to explain in terms of the physics of the
jet-star interaction. It is well known that fully developed turbulence
has a 5/3 spectrum (Beloborodov 1998, 2000; Kumar \& Narayan 2009;
Narayan \& Kumar 2009). Turbulent motions are observed in the
simulations, but are concentrated in the cocoon material and not in
the outflow. Even though the collimation shocks are due to the cocoon
pressure on the side of the jet, it is hard to imagine how the
turbulence spectrum can be imprinted directly on the jet light-power
curves.
The qualitative similarity of the observed PDS with the synthetic one is
suggestive that the long time-scale variability in GRBs (down to
tenths of seconds) may be due to the jet interaction with the star and
not to the inner engine variability. However, GRB light curves show
fast spikes with duration of milliseconds or less that cannot be
explained as due to the interaction with the stellar progenitor
(Walker et al. 2000). It seems therefore that there are two sources of
variability in the light curves: the intrinsic variability of the
engine, likely operating on the ms time scale since the engine is a
compact object, and the star-induced variability, with a cutoff at
several tenths of a second, possibly associated with the sound crossing
time of the jet at the star's surface. To study if and how these two
variability sources interfere with each other, we compared the PDS of
the light-power curve from {\it uniform} and {\it variable}
simulations.
Figure~\ref{fig:pds2} shows the comparison of the PDS from the
light-power curves of the various simulations. The power spectrum of
the input variability in the {\it variable} simulations is also
shown. The comparison of the various spectra shows that the two
sources of variability have no significant interaction, with the {\it
variable} spectra being qualitatively similar to the sum of the {\it
uniform} spectrum - which contains only the star-induced variability
- and the input spectrum. The two {\it variable} spectra are also
qualitatively similar, indicating that the variability of the jet
depends only on the variability of the total luminosity of the engine
and not on whether the variability is in the entropy or in the baryon
loading of the outflow.
The PDS of the {\it step} simulation also appears similar to the
{\it uniform} spectrum, with a series of
strong resonant peaks at multiples of $5$~Hz, the frequency of the
on-off cycle, added at higher frequencies.
Again, further simulations will be required to
confirm the generality of the conclusions against different
progenitors and different jet properties.
\section{Discussion and conclusions}
We have carried out high-resolution hydrodynamic simulations of the
propagation of baryonic GRB jets through their progenitor stars and
beyond. For the first time, we explored the consequences of a variable
injected luminosity with AMR simulations. Our simulations allowed us
to draw a number of conclusions about the role played by the
progenitor star material in shaping the final appearance of the prompt
light curve:
\begin{itemize}
\item The propagation time of the jet in the progenitor star depends
on the variability properties of the engine in a complex way. The
jet propagation inside the star is affected by the detailed
structure of turbulent eddies inside the cocoon, the development of
which is highly non-linear and depends on the resolution of the
simulation. The propagation time is a very important quantity in any
GRB model since it is the quantity that establishes whether the
engine is active for a time long enough to power a GRB or not. If
the engine is active for a time shorter than the propagation time,
it is more likely that an ``engine powered'' supernova is produced
(Soderberg et al. 2010) rather than a successful GRB event. The
propagation time also sets the amount of jet energy that is
dissipated in the cocoon, available to power a precursor
(Ramirez-Ruiz et al. 2002; Lazzati \& Begelman 2005) or possibly a
short duration GRB (Lazzati et al. 2010). High resolution 3D
simulations are required to nail down this important aspect of the
jet propagation in the progenitor star.
\item Outside their progenitor stars, jets show a much more
predictable evolution that does not depend on details. All jets
show the three phases of evolution described in MLB07: (i) a
wide-angle cocoon phase, (ii) a narrow shocked jet phase lasting 10
to 15 seconds, and (iii) an expanding jet phase. This behavior and
its timing seem to be fairly independent of the simulation
resolution as well as the details of the central engine.
\item Contrary to what was previously thought (Zhang et al. 2003;
MLB07), short time-scale variability injected by the central engine
is preserved in the jet, even during the shocked phase. The
variability injected by the central engine is lost only during a
fast initial phase, lasting a few seconds. On the other hand, the
interaction of the jet material with the surrounding stellar
material creates long time-scale variability even for jets injected
with no variability whatsoever. When a variable jet is injected, we
find that the progenitor induced variability is not affected, at
least at the qualitative level. For the range of simulations studied
here, we find that the properties of the variability at time-scales
of seconds depend on the progenitor structure and are insensitive to
the properties of the outflow injected by the central engine. This
is an important finding because it implies that the fast variability
in GRB light curves, observable during the bright shocked-jet phase,
may depend only on the properties of the GRB engine. We can therefore
study the nature of the GRB engine (still clouded in mystery) by
studying the properties of the fast variability in GRB light curves.
\item Even though we inject an uniform outflow in the center of the
star, with no dependence of the jet properties on the polar angle,
the jet propagating in the circumstellar material is highly
structured. The total observed energy decreases with the off-axis
angle as $dE/d\Omega\propto\theta^{-3}$. The peak luminosity, on the
other hand, is constant between about $5^\circ$ and $20^\circ$. This
behavior is due to the marked changes in the temporal properties of
the jet at different off-axis angles. The diversity of light curves
observed in GRB catalogs can therefore be partly attributed to the
viewing geometry.
\item We computed the power density spectrum (PDS) of light-power
curves from an extended simulation (reaching distances from the
progenitor star ten times larger than our standard simulations). It
shows two important similarities with the PDS of BATSE light curves
(Beloborodov et al. 1998; see Fig.~\ref{fig:pds1}): a high frequency
cutoff at a frequency of few hertz and a power-law
behavior consistent with a slope $PDS(f)\propto{}f^{-5/3}$ at lower
frequencies. The high frequency cutoff
may tentatively
be interpreted as the
result of the timescale of the propagation of perturbations across
the jet,
although the cutoff is near the resolution limit of the simulation
and may be a numerical artifact.
If the presence of the high-frequency cutoff can be confirmed,
the coincidence of the simulated cutoff to the
observed one may be the first direct confirmation that long duration
GRBs are associated with massive stars at all redshifts and not only
at $z<1$. The coincidence between the simulated and observed
power-law slope suggests instead that the variability that
characterizes GRB light curves is mostly due to the interaction of
the jet with the progenitor star and not to the GRB engine. Such a
result has been obtained under the assumption that the radiation is
released at a constant radius and at constant efficiency. In
addition, the PDS is steeper, with a slope $PDS(f)\propto{}f^{-2}$
if the initial 5 seconds are removed from the light curve. Since the
structure of the head of the jet is affected by the dimensionality
of the simulations (cfr. Zhang et al. 2004), we consider this result
suggestive and not conclusive.
\item A PDS analysis was also performed for the {\it variable}
simulations (Fig.~\ref{fig:pds2}). The low frequency part of the PDS
of {\it variable} simulations show a power-law behavior
qualitatively similar to the one of the {\it uniform}
simulation. The analogy suggests that, as already discussed in the
light curve analysis, the long time-scale variability is a result of
the jet-star interaction that does not depend on the details of the
engine properties. On the other hand, the high frequency part of the
{\it variable} PDS has the same qualitative shape of the PDS of the
luminosity injected by the engine. This suggests that at high
frequencies the propagation of the jet through the star has no
impact on the jet structure. The high frequency features are
transported unmodified by the relativistic outflow. If our results
are confirmed by simulations with different progenitors and jet
properties, the analysis of fast variability in GRB light curves
could give us important clues to the nature of the GRB engine, while
the analysis of the long time-scale variability can yield important
clues to the nature of the progenitor star.
\end{itemize}
There are several potential sources the could create variability in the inner $10^{9}$~cm of the star, including instabilities in the propagating jet, instabilities in the region of jet formation, an inherent variability in the energy available from the central engine (i.e. from a variable accretion rate onto a black hole or a varying spin down rate of a proto-magnetar). Disentangling the sources that give rise to short-timescale variability in observations is likely to be quite challenging. However, whatever variability does arise in the innermost region of the progenitor is preserved during propagation and remains imprinted on the jet that emerges from the stellar surface.
Based on all the evidence gathered in this project, we advance the
possibility that variability has two origins, with engine and
propagation having comparable roles in shaping the light curve in
different regions of the frequency domain. This possibility is shown
graphically in Figure~\ref{fig:1606} that shows the BATSE light curve
of GRB 920513 with a thick line underlying the envelope of long
time-scale variability due to the jet interaction with the progenitor
star. Faster variability is present in the light curve and that is
interpreted as variability in the luminosity injected by the central
engine at the base of the jet. Observations of high frequency
variability, therefore, may offer important clues about the nature and
evolution of the central engine.
\acknowledgements The software used in this work was in part developed
by the DoE-supported ASC / Alliance Center for Astrophysical
Thermonuclear Flashes at the University of Chicago. This work was
supported in part by NASA ATP grant NNG06GI06G and Swift GI program
NNX06AB69G and NNX08BA92G. We thank NCSA and the NASA NAS for the
generous allocations of computing time.
|
\section{Introduction}
When L\'evy \cite{le} introduced his famous curve in 1938, he also
constructed fractal surfaces in a similar way. 70 years later, we
have plenty of papers on fractal sets in the plane, and a number
of general statements and constructions which hold in every
dimension, but very few studies on the geometry of fractals in
dimension $\ge 3.$
Visualization is certainly one reason to prefer plane sets: they
can be easily shown on a computer screen. In dimension 3,
visualization is more difficult, even though there are good
ray-tracing programs, like chaoscope \cite{de} and IFS builder
\cite{km}, which allow to look at three-dimensional fractals from
any chosen viewpoint, with prescribed light sources. Figure
\ref{twin6} shows two views of one of our main examples. Does it
have interior points? If so, is the interior connected? Is the
figure simply connected? Unlike in two dimensions, the answer to
such questions can hardly be guessed from the pictures - this
requires tools which we develop here.
There are several mathematical difficulties in dimension $\ge 3.$
We avoid the problem that linear maps do usually not commute, by
concentrating ourselves to fractals generated by a single matrix.
But there is another problem which will concern us: proper
self-similarity is rather an exceptional case. The moduli of all
eigenvalues of a $3\times 3$ matrix do rarely coincide while for a
$2\times 2$ matrix they coincide as soon as the eigenvalues are
complex. This problem is addressed in Section 2, and Theorem 2.3
implies that in ${\mathbb R}^3$ there is essentially only one type
of `integer' self-similar tile with 2 up to 7 pieces.
\begin{figure}
\begin{center}
\includegraphics[width = 150mm]{twin6.eps}
\caption{\label{twin6} Two views of the twindragon \CC , made by chaoscope \cite{de}.
Is this set homeomorphic to a ball? Is the interior connected?
We develop concepts and tools to answer such questions.}
\end{center}
\end{figure}
Other difficulties concern the topology of three-dimensional
fractals. Our main aim here is to study the topology of
self-affine tiles, and of their boundaries which are fractal
surfaces. In particular, we would like to know which of these
tiles are homeomorphic to a closed ball. In two dimensions, the
corresponding question for disk-like tiles was discussed in a
large number of papers \cite{at1,at2,bg,bw,gg,lrt,ll,st,t} (more
references can be found in \cite{ll,v,w}), and Jordan curve
arguments were heavily used. Among others, it is easy to see that
connectedness of the interior of a self-affine tile in the plane
is sufficient for its disk-likeness. In ${\mathbb R}^3$ this is
not the case. The topology of the boundary of a tile can be much
more complicated than in ${\mathbb R}^2.$ Even the connectedness
of a tile is not so easy to verify \cite{ag,klr}.
The geometric phenomena which we encounter in fractal tiles are
not caused by knots or wild spheres \cite{bc}. It is rather the
self-affine fibre structure of the boundary which makes the
topology complicated. Apparently, the different eigenvalues of the
generating matrix produce long and thin fibres which can pierce
the interior of neighboring tiles, or distort the boundary
structure.
We develop algebraic tools which describe the geometry of a
self-affine tile $T$ in arbitrary dimension, in a similar way as
homotopy and homology groups describe the geometry of manifolds.
Our tools use self-affinity in a specific way, and they yield
quite detailed information about the geometry. The basic concept
is the {\it neighbor graph} $G=(V,E).$ It can be considered as a
blueprint of $T$ which contains all information about the topology
of $T$ in a simple scheme. In the case where $T$ tiles by
translations in a unique way, the vertices $v\in V$ are those
translations for which $T+v$ belongs to the tiling and intersects
$T.$ Such translations $v$ will be called neighbor maps, and they
do not only represent the neighboring piece $T+v$ but also the
boundary set $T\cap (T+v)$ of $T.$ This set can be a face, an
`edge' or just a point of $T,$ or a more complicated fractal
boundary set. For a ball-like tile we expect that faces are
homeomorphic to disks and edges are homeomorphic to intervals, and
we have to check this idea.
Neighbors in lattice tilings were introduced by Indlekofer,
K\'atai and Racsko \cite{ikr}, and neighbor maps for self-similar
sets were defined by Bandt and Graf \cite{bg}. For a self-affine
tile, the boundary sets have themselves a self-affine structure:
each face or edge is the union of smaller copies of other boundary
sets. This is indicated by the edges of the graph $G.$ The
definition of $G$ in Section 3 implies that the boundary $\partial
T$ is a self-affine graph-directed construction in the sense of
Mauldin and Williams \cite{mw}, and the graph which directs this
construction is $G$ itself. This fact was stated by Scheicher and
Thuswaldner \cite{st} for lattice tiles, and used in many other
papers to determine the Hausdorff dimension and topological
structure of $\partial T$ for plane self-similar tiles
\cite{gi,sw,klsw,lrt,ll,t,ve}. Some authors
\cite{dkv,st,at1,at2,al} prefer the contact matrix defined by
Gr\"ochenig and Haas \cite{gh} which describes a subgraph of $G.$
A first application of the neighbor graph concerns the existence
of self-affine lattice tilings, a problem treated in \cite{gh} and
in a series of papers by Vince \cite{v1,v4,v}. Theorem 5.1 shows
that the existence of such tilings is equivalent to the
compatibility of all neighbors, which can be easily checked with
$G.$
In sections 7 to 12 we study the topology of the boundary of $T.$
In section 8 we define the hierarchy of different boundary sets
like faces, edges, points etc. which corresponds to the hierarchy
of sets with different Hausdorff dimensions in Mauldin-Williams
constructions. An important problem is to determine all {\it
faces} -- those boundary sets which cover an open set on the
boundary. Two methods are developed for this purpose. Theorem 9.2
gives an algorithmic approach which only uses the combinatorics of
the neighbor graph. Theorem 10.1 provides an analytic approach
based on recent work by He and Lau \cite{hl} and Akiyama and
Loridant \cite{al}, connected with eigenvalues of adjacency
matrices and Hausdorff dimension. In section 11 we define a
polyhedral structure of fractal tiles by the intersection sets of
two and more faces. There are corresponding neighbor intersection
graphs $G^2,G^3,...$ which allow to calculate details of the
polyhedral structure and compare it with the structure of ordinary
polyhedra.
In contrast to the contact matrix, all these tools can also be
applied to tiles with incompatible neighbors as well as to
fractals which are not tiles, and neighbors can be related by
arbitrary isometries of ${\mathbb R}^3$ instead of translations.
To illustrate our methods, we selected a small class of examples,
which is introduced in section 6: the {\it twindragons.} These are
the self-affine lattice tiles with two pieces. The one-dimensional
twindragon is the interval. In the plane, there are three
examples: rectangle or parallelogram, the twindragon and the tame
twindragon. They are all disk-like \cite{ikr,bg}. In
three-dimensional space, there are seven examples, which we denote
by letters $\mathcal A$\ } \def\BB{$\mathcal B$\ } \def\CC{$\mathcal C$\ to $\mathcal G$\ . While $\mathcal A$\ } \def\BB{$\mathcal B$\ } \def\CC{$\mathcal C$\ is conjugate to a cube, \FF and
$\mathcal G$\ have an extremely intricate structure. We mention some of
their properties without proof. For $\mathcal D$\ } \def\EE{$\mathcal E$\ } \def\FF{$\mathcal F$\ and \EE we determine the
boundary faces and their intersections, and we show that their
interior is not simply connected. \BB and \CC seem to be
homeomorphic to a ball. In section 12 we determine their exact
polyhedral structure. We show that they have faces which are not
homeomorphic to a disk, but \CC has connected interior.
This paper focusses on concepts rather than algorithmic questions
although it seems clear that the neighbor graph and related finite
automata are well-suited for computerized evaluation. Section
\ref{algo} gives a short account of algorithmic aspects.
\section{Self-affine lattice tiles}
{\bf Basic concepts. } The following definitions are standard
\cite{lw,lw1,v,w}. A {\it lattice} $L$ in $\R^n$ is the set of
integer linear combinations of $n$ linearly independent vectors
$e_1,...,e_n.$ Usually we take $L=\Z^n.$ A regular-closed set $T$
is called a {\it tile} if it admits a tiling of $\R^n.$ A {\it
tiling} by the tile $T$ is a countable union $\bigcup_k T_k$ which
covers $\R^n,$ such that each $T_k$ is isometric to $T,$ and no
two tiles have interior points in common. We have a {\it lattice
tiling} by $L$ if the tiles $T_k$ are just the translates $T+k$
with $k\in L.$
A tile $T$ is a {\it self-affine lattice tile} if there is an
affine expanding mapping $g$ and a lattice $L$ such that $g$
preserves $L$ and maps $T$ to a union of tiles $T+k_i$ with
$k_i\in L.$ With respect to the basis vectors $e_i,$ the map $g$
has a matrix representation $g(x)=Mx$ where $M$ is an integer
matrix. Expanding means that all eigenvalues of $M$ have modulus
greater one. $T$ is determined by the map $g$ and by the lattice
coordinates $k_1,...,k_m$ of the tiles which form the supertile
$g(T).$
\begin{equation} g(T)=MT=\bigcup_{j=1}^m T+k_j \label{selfa}
\end{equation}
For $T$ to become a tile, we must have $m=|\det M|,$ and the set
$K=\{k_1,...,k_m\}$ must fulfil some condition. A sufficient
condition is \cite{b5}
\begin{equation} g(L)=\bigcup_{j=1}^m L+k_j=L+K \label{selfb}
\end{equation}
in which case $K$ is called a {\it standard digit set} or {\it
complete residue system.} The latter name comes from the fact that
for $|\det g|=m$, the subgroup $g(L)$ of the additive group $L$
has $m$ residue classes, so by (\ref{selfb}) each class is
represented by exactly one $k_i.$ See Lagarias and Wang
\cite{lw,lw1} who also studied non-standard digit sets. In this
paper, we consider only standard digit sets.
The self-affine tile $T$ is called {\it self-similar} if the
mapping $g$ is a similarity mapping with respect to the Euclidean
metric. In this case, we must distinguish the standard basis which
defines the Euclidean distance, and the basis for the lattice $L$
which need not be orthonormal.
{\bf Conjugacy of tiles. } $T$ is said to be {\it conjugate to a
self-similar tile} if there is a linear map $h$ so that
$\tilde{T}=h(T)$ is a self-similar tile. The expanding map for
$\tilde{T}$ is $\tilde{g}=hgh^{-1},$ the lattice is
$\tilde{L}=h(L),$ and the digits are $\tilde{k}_i=h(k_i),$ for
which (\ref{selfa}) and (\ref{selfb}) are easily checked.
\begin{Exa} In the plane, we consider the digits
$k_1={0\choose 0}, k_2={1\choose 0}.$ The parallelogram $T$ with
vertices ${0\choose 0}, {1\choose 0}, {1\, \choose 1}, {2\choose
1}$ is a self-similar tile with respect to $M={1\ 1\choose 1\,
-1}$ and the lattice $L=\Z^2.$ The rectangle $T'=[0,1]\times
[0,\sqrt{2}]$ is also a self-similar lattice tile, with $M'={0\
\sqrt{2} \choose \sqrt{2}\ 0}$ and $L'=\{ {m\choose\sqrt{2}\, n}\,
|\, m,n\in\Z\} .$ The unit square $T''=[0,1]^2$ is a self-affine
tile with respect to the matrix $M''={0\, 2 \choose 1\, 0}$ and
$L''=\Z^2,$ but not self-similar. However, $T''$ is conjugate to
the self-similar tile $T'$ since $h(x)={1\,\ 0 \choose
0\,\sqrt{2}}x$ fulfils $h(T'')=T'.$ Note that $\tilde{h}(T)=T'$
for $\tilde{h}(x)={1\, -1 \choose 0\,\sqrt{2}}x.$ \end{Exa}
We shall identify conjugate tiles so that $T,T',T''$ are
considered as different versions of the same tile, which is
essentially self-similar. {\it Considering tiles up to conjugacy
means that we focus on the essential data: the characteristic
polynomial of $g$ instead of different matrix representations.} We
shall also assume that $k_1={0\choose 0}$ since this can be
obtained by conjugacy with a translation.
\begin{Prop} (The class of self-similar tiles)\\
A self-affine lattice tile $T$ with respect to $g(x)=Mx$ is
conjugate to a self-similar tile if and only if all eigenvalues of
$M$ have the same modulus.
\end{Prop}
\begin{Proof} Necessity of the condition follows from the fact
that eigenvalues are not changed under conjugacy. On the other
hand, if the eigenvalues have equal modulus, there is a matrix $B$
such that $\tilde{g}(x)=BMB^{-1}x$ is a similarity map. Thus
$h(x)=Bx$ maps $T$ to a self-similar tile.\end{Proof}
\begin{Thm} (Very few self-similar lattice tiles in dimension 3)\\
If in $\R^3$ a self-affine lattice tile $T$ with respect
to $g(x)=Mx$ is conjugate to a self-similar tile, then one of the
following two conditions is true.
\begin{enumerate}\item[ (i)] $m=|\det
M|$ is a cubic number - in particular $m\ge 8,$
\item[(ii)] $M$ is conjugate to $\tilde{M}=\left(\begin{array}{lcr}0&0&\pm m\\
1&0&0\\ 0&1&0\end{array}\right) .$
\end{enumerate}
\end{Thm}
\begin{Proof} With respect to the lattice base, $g$ has an integer
matrix, and so the characteristic polynomial $p(\lambda)$ of $M$
has integer coefficients. We express them in terms of the
eigenvalues $\lambda_i.$
\[ p(\lambda)=-\lambda^3 +(\lambda_1+\lambda_2+\lambda_3)\cdot\lambda^2
-(\lambda_1\lambda_2+\lambda_1\lambda_3+\lambda_2\lambda_3)\cdot\lambda
+\lambda_1\lambda_2\lambda_3 \] where
$\lambda_1\lambda_2\lambda_3=\det M=\pm m.$ Since $T$ is conjugate
to a self-similar tile, all $\lambda_i$ have equal modulus:
$|\lambda_i|=r>1.$
If the eigenvalues are real, then $\lambda_i=\pm r$ and $m=\pm
r^3.$ In this case $\lambda_1+\lambda_2+\lambda_3\in \{\pm r, \pm
2r, \pm 3r\},$ so $r$ must be an integer and $m$ a cubic number.
Now let us assume $\lambda_1,\lambda_2$ are complex eigenvalues.
Then $\lambda_1=r(\cos\alpha +i\sin\alpha)$ and
$\lambda_2=r(\cos\alpha -i\sin\alpha)$ for some $\alpha ,$ so
\[ p(\lambda)=-\lambda^3 +2r(\cos\alpha +\frac{s}{2})\cdot\lambda^2
-(r^2+2sr^2\cos\alpha)\cdot\lambda +r^3s
\] where $s=\pm 1$ is the sign of $\lambda_3.$ The coefficient of
$\lambda$ can be written as $-2sr^2(\cos\alpha +\frac{s}{2}).$
Since the coefficients are integers, either $r$ must be a rational
number - and hence an integer, and $m$ a cubic number. Or
$\cos\alpha +\frac{s}{2}$ must be zero. In this case,
$p(\lambda)=-\lambda^3+sm$ which is the characteristic function of
$\tilde{M}.$
\end{Proof}
\begin{Rem}
The orthogonal part of $\tilde{M}$ is a rotation around $120^o$
for $s=+1,$ and a rotation around $60^o$ composed with a
reflection at the plane of rotation for $s=-1.$ Even in the case
where $r$ is an integer and $m$ a cubic number, $2r\cos\alpha$
must be an integer and $\alpha$ can assume only few values. Thus
there are really very few self-similar lattice tiles in $\R^3.$
\end{Rem}
\section{Neighbors of tiles and self-affine sets}
When we want to study the boundary $B$ of a lattice tile $T,$ it
is quite natural to consider the {\it neighbors} in the tiling -
those tiles $T+k$ for which $B_k=T\cap (T+k)$ is non-empty. The
$B_k$ can be considered as the faces of $T,$ and $B$ is the union
of the $B_k.$ Obviously, the number of neighbors is finite. This
idea was introduced by Gilbert \cite{gi} and Indlekofer, K\'atai
and Racsko \cite{ikr} and used in many other papers.
\smallskip
{\bf Neighbors in fractals. } There is a related concept of
neighbor in self-similar fractals which appeared first in
\cite{bg}. Suppose we have a fractal $A$ which consists of two
copies or itself: $A=f_1(A)\cup f_2(A),$ where the $f_i$ denote
contracting similarity maps. Then the geometry and topology is
determined by the structure of the intersection set $C=f_1(A)\cap
f_2(A).$ Only at points of $C$ it makes sense to zoom into the
picture. If $x$ is in $f_1(A),$ say, and $U$ is a neighborhood of
$x$ in $A$ which does not intersect $C,$ then the magnified copy
$f_1^{-1}(U)$ of $U$ does already exist in $A.$
The sets $f_1^{-1}(C)$ and $f_2^{-1}(C)$ can be considered as
boundary sets of $A,$ where $A$ intersects a {\it potential
neighbor} $f_1^{-1}f_2(A)$ or $f_2^{-1}f_1(A).$ These need not be
the only boundary sets, however, since $C$ consists of
intersections of smaller pieces like $f_1f_2(A)\cap f_2f_2(A),$
and these may be at other positions when they are pulled back to
$A.$ This argument shows that the boundary sets have a
self-similar structure. The number of boundary sets obtained by
going to smaller and smaller pieces need not be finite, but in
most common examples it is.
Let us define neighbors in this general sense. Let $f_1,...,f_m$
denote contractive affine mappings on $\R^n.$ There is a unique
compact non-empty set $A$ with
\begin{equation}A=\bigcup_{j=1}^m
f_j(A),\label{selfc}\end{equation} the {\it self-affine set} with
respect to the $f_i,$ cf. \cite{fa}. For each integer $q$ the set
$A$ splits into $m^q$ small copies $f_{{\bf j} } \def\ii{{\bf i} }(A)=A_{{\bf j} } \def\ii{{\bf i} },$ where
${\bf j} } \def\ii{{\bf i} =j_1...j_q$ denotes a word of length $q$ from the alphabet
$I=\{ 1,...,m\},$ and $f_{\bf j} } \def\ii{{\bf i} =f_{j_1}\cdot ...\cdot f_{j_q}.$
A potential {\it neighbor of $A$} has the form
$h(A)=f_{\ii}^{-1}f_{{\bf j} } \def\ii{{\bf i} }(A)$ where $\ii,{\bf j} } \def\ii{{\bf i} \in I^*$ are any words
over $I$ such that the pieces $A_{\ii}=f_{\ii}(A)$ and
$A_{{\bf j} } \def\ii{{\bf i} }=f_{{\bf j} } \def\ii{{\bf i} }(A)$ intersect. The idea is that $f_{\ii}^{-1}$
maps $A_{\ii}$ to $A$ and ${A_{\bf j} } \def\ii{{\bf i} }$ to $h(A),$ a neighbor set
which intersects $A$ and should be of comparable size. $h$ is
called a {\it neighbor map} and the set $B=A\cap h(A)$ the
corresponding {\it boundary set} of $A.$ To get `comparable size',
we confine ourselves to words $\ii,{\bf j} } \def\ii{{\bf i} $ of equal length and make
some assumptions concerning the $f_i.$
\begin{Prop} (Neighbor maps which are isometries)\\
Let $A\subset \R^n$ be a self-affine set
of the form (\ref{selfc}). Let $\ii,{\bf j} } \def\ii{{\bf i} $ be words of the same
length $q,$ and $h=f_{\ii}^{-1}f_{{\bf j} } \def\ii{{\bf i} }.$
\begin{enumerate} \item[ (i)] If all $f_j$ are similarity maps with the
same similarity factor $r,$ then $h$ is a Euclidean isometry.
\item[(ii)] If $f_j(x)=M^{-1}(x +k_j)$ where $M$ is an expanding matrix
and $k_1,...,k_m\in\R^n,$ then $h$ is a translation.
\end{enumerate}
\end{Prop}
\begin{Proof} (i): If the $f_j$ are similarity maps with factor
$r$ then $f_{{\bf j} } \def\ii{{\bf i} }$ and $f_{\ii}^{-1}$ are similarity maps with
factor $r^q$ and $r^{-q},$ respectively. Thus $h$ is an
isometry.\\
(ii) is first proved for $q=1.$ Since $f_i^{-1}(y)=My-k_i,$ we
have
\begin{equation}f_i^{-1}f_j(x)=x+k_j-k_i \label{neibeg}\end{equation}
Moreover, if $h(x)=x+v$ is a translation then
\begin{equation}
f_i^{-1}hf_j(x)=x+Mv+k_j-k_i \label{induc}
\end{equation}
Now induction on $q$ shows that for $h=f_{\ii}^{-1}f_{{\bf j} } \def\ii{{\bf i} },$
\begin{equation} h(x)=f_{i_q}^{-1}...f_{i_1}^{-1}
f_{j_1}...f_{j_q}(x)=x+M^{q-1}(k_{j_1}-k_{i_1})+M^{q-2}(k_{j_2}-k_{i_2})
+...+(k_{j_q}-k_{i_q}) \label{neivec}
\end{equation}
\end{Proof}
{\bf Lattice tiles as a special case. } The case (ii) includes our
self-affine lattice tiles since the defining equation
(\ref{selfa}) can be rewritten as
\begin{equation} T=\bigcup_{j=1}^m g^{-1}(T)+g^{-1}(k_j)
=\bigcup_{j=1}^m M^{-1}(T+k_j) \label{selfd}
\end{equation}
We get $f_j(x)=M^{-1}(x+k_j)$ as contracting maps for the
self-affine set $T.$
At this point it is possible to explain the concept of standard
digit set $k_1,...,k_m$ which says that $k_{j_q}-k_{i_q}$ is not
in $ML$ for $j_q\not= i_q$ (see section 2). This property implies
that the translation vector in (\ref{neivec}) can never be zero if
$j_q\not= i_q$ since $k_{j_q}-k_{i_q}$ cannot cancel with the
other terms in (\ref{neivec}). As a consequence, the inequality
$h\not= id$ then holds whenever $\ii\not={\bf j} } \def\ii{{\bf i} .$ From the results
in \cite{bg} then follows the so-called open set condition which
says that the pieces $f_{\ii}(T)$ and $f_{{\bf j} } \def\ii{{\bf i} }(T)$ have no
interior points in common. (In \cite{b5}, this was shown by
Baire's category theorem.) In other words, $T$ and a translate
$h(T)$ in (\ref{neivec}) have no common interior points.
\section{The neighbor graph}
{\bf Self-similarity of boundary sets. } To explain the
self-similar structure of boundary sets, we return to the general
case. Consider intersecting pieces $A_{\ii}$ and $A_{{\bf j} } \def\ii{{\bf i} }$ in the
self-affine set $A.$ Since all pieces divide into $m$ subpieces,
like $A$ in (\ref{selfc}),
\[ A_{\ii}\cap A_{{\bf j} } \def\ii{{\bf i} } =\bigcup_{i,j=1}^m A_{\ii i}\cap A_{{\bf j} } \def\ii{{\bf i} j}
.\] Now consider the boundary sets $B=f_{\ii}^{-1}(A_{\ii}\cap
A_{{\bf j} } \def\ii{{\bf i} })$ and $B_{ij}=f_{\ii i}^{-1}(A_{\ii i}\cap A_{{\bf j} } \def\ii{{\bf i} j}).$
Then
\[ B=f_{\ii}^{-1}\left(\bigcup_{i,j=1}^m A_{\ii i}\cap A_{{\bf j} } \def\ii{{\bf i} j}\right)
=\bigcup_{i,j=1}^m f_if_i^{-1}f_{\ii}^{-1}(A_{\ii i}\cap A_{{\bf j} } \def\ii{{\bf i}
j}) =\bigcup_{i,j=1}^m f_i (B_{ij}) . \] At $B,B_{ij}$ we
suppressed the subscripts $\ii,{\bf j} } \def\ii{{\bf i} $ since such an equation holds
for each possible boundary set. In other words, {\it the
subdivision of pieces induces self-similar representations of the
type} (\ref{selfc}) {\it for all boundary sets} - not with one
type of set, but with several types. We must assume now that we
have only finitely many possible boundary sets, which is true for
lattice tiles with standard digit sets. The unions on the right
contain a lot of empty terms, so we introduce a graph which better
describes the system of equations.\smallskip
{\bf Concept and properties of neighbor graph. } The {\it neighbor
graph} $G=(V,E)$ of a self-affine fractal $A$ has as vertex set
all neighbor maps $h=f_{\ii}^{-1}f_{{\bf j} } \def\ii{{\bf i} },$ and a directed edge
marked with $i$ goes from $h$ to
$h'=f_i^{-1}f_{\ii}^{-1}f_{{\bf j} } \def\ii{{\bf i} }f_j.$ Loops and multiple edges
(even with the same label) are possible \cite{bm}. For a formal
definition, we focus on self-affine tiles.
\begin{Def}\label{defng}
Let $T=\bigcup f_j(T)$ be a self-affine lattice tile with
$f_j(x)=M^{-1}(x+k_j)$ so that the neighbor maps have the form
$h(x)=x+k.$ We identify $h$ with the translation vector $k,$ and
asopciate it with the boundary set $B_k=T\cap (T+k).$ An edge from
$k$ to $k'$ with label $i$ is drawn when relation (\ref{induc})
holds:
\[ G=(V,E) \quad\mbox{ with }\quad
V=\{ k\, |\, T\cap (T+k)\not=\emptyset\} \qquad \mbox{ and }\]
\begin{equation} E=\bigcup_{i=1}^m E_i\quad\mbox{ with }\quad
E_i=\{ (k,k',i)\, |\, k'=Mk+k_j-k_i \}
\label{neigra}
\end{equation}\end{Def}
Here $E_i$ denotes the set of edges with label $i.$ It should be
mentioned that $V$ contains a root vertex $k=0$ (or $h_0=id$ in
the more general notation) which does not correspond to a boundary
set. There are edges $(0, k_j-k_i,i)$ from the root which
correspond to the first step (\ref{neibeg}) in the calculation of
neighbor maps. The loops $(0,0,i)$ will not be drawn, however. For
the calculation of $G,$ see section \ref{algo}, \cite{bm,br} and
the example below.
\begin{Prop} (The boundary equations)\\
Let $T=\bigcup_{j=1}^m f_j(T)$ be a self-affine lattice tile with
neighbor graph $G=(V,E).$ The boundary sets $B_k$ corresponding to
the $k\in V\setminus \{0\}$ fulfil the following equations.
\begin{equation}
B_k=\bigcup\, \left\{ f_j(B_{k'}\, |\, j\in\{ 1,...,m\} ,\,
(k,k',j)\in E\right\} \label{gradir}\end{equation}
\end{Prop}
This is an immediate consequence of Definition \ref{defng} and of
the discussion above. Mauldin and Williams \cite{mw} called
families of such fractals $B_k$ {\it graph-directed
constructions}, and proved that the $B_k$ are uniquely determined
by the graph $G$ and the maps $f_j.$ At the end of section 10 we
briefly discuss the open set condition of (\ref{gradir}).
\begin{Prop} (The basic symmetry of $G$)\\
If $h$ is a neighbor map, then $h^{-1}$ is also a neighbor map.
For translation this means that to every $k$ there is a $-k,$ and
each boundary set $B_k=T\cap (T+k)$ has an opposite set
\[ B_{-k}=(T-k)\cap T=B_k -k.\] This symmetry of $V$ will extend as
a graph isomorphism to the edges, but not to the labels. If
$k'=Mk+k_j-k_i$ then $(k,k')$ has label $i$ and $(-k,-k')$ has
label $j.$
\end{Prop}
In a previous publication with M. Mesing \cite{bm}, we wrote both
labels $i,j$ at the edge $(k,k').$ This is not necessary here
since we can always recover the second label from the opposite
edge.
{\bf A two-dimensional example. } In our examples, neighbors will
be denoted by lower-case Roman letters, and $-\bb$ will denote the
opposite vertex of b. We start with a two-dimensional tile
\cite{b5,v} which can be considered as a modification of the
square, and as an extension of the Sierpi\'nski gasket. The
construction of neighbor graphs for three-dimensional tiles
proceeds in the same way as demonstrated here.
\begin{Exa}
Let $f_j(x)=(x+k_j)/2$ for $j=1,...,4$ where $k_1={0\choose 0},
k_2={1\choose 0}, k_3={0\choose 1},$ and $k_4={-1\choose -1}.$ The
tile $T$ is shown in Figure \ref{siertile}. Clearly, $T\subset
[-1,+1]^2.$ Thus a translation $x={x_1\choose x_2}$ with $T\cap
(T+x)\not=\emptyset$ must fulfil $|x_1|\le 2$ and $|x_2|\le 2.$
Moreover, $L$ is the integer lattice, so $x_1,x_2\in\{
-2,-1,0,1,2\} .$ In this way we prove that for an arbitrary
self-affine lattice tile, the set $V$ is finite.
The tile $T$ has six neighbors which intersect $T$ in a single
point, and six neighbors which intersect $T$ in an uncountable set
which is in fact a Sierpi\'nski gasket. \end{Exa}
\begin{figure}
\begin{center}
\includegraphics[width = 150mm]{siertile.eps}
\caption{\label{siertile} A two-dimensional tile
and its neighbor translations.}
\end{center}
\end{figure}
\begin{Proof} We start with the translation vectors $k=k_j-k_i$
where $i\not= j,$ see the right-hand part of Figure
\ref{siertile}. If $k$ is in $V,$ there is the edge $(0,k,i).$
Since our matrix is $M={2\ 0\choose 0\ 2},$ all these points do
belong to $V,$ with a loop $(k,k,j)$ at each point $k,$ because
$k_j-k_i=2(k_j-k_i)+k_i-k_j.$
Moreover, these will be the only points of $V.$ For the other
points $x$ with $|x_1|\le 2, |x_2|\le 2$ the recursion
(\ref{induc}) will lead to vectors $k'$ which are outside the
range of $x_1,x_2$ and thus are no neighbor maps. A simple
calculation \cite{bm} shows that once we are outside the range, we
can never come in again by the recursion (\ref{induc}).
Thus the translations $\pm{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} =\pm {1\choose 1}, \pm\bb=\pm
{-1\choose 0}, \pm\cc=\pm {0\choose -1}$ marked by $\bullet ,$ and
$\pm {2\choose 1}, \pm {1\choose 2}, \pm {1\choose -1}$ marked by
$\circ$ are the vertices of our neighbor graph. The last six
vertices have only the loop and no further outgoing edges. These
six vectors are the translations along the sides of the big
triangle in Figure \ref{siertile}. They translate $T$ to a
neighbor $h(T)$ which has a single point as intersection with $T.$
These neighbors are called {\it point neighbors.}
To find point neighbors, we need only look at the graph, not at
the picture. The loop at $k={2\choose 1}$ with label $2$ says that
the boundary set has address $2222...=\overline{2},$ and the
opposite vector ${-2\choose -1}$ has address $\overline{4}.$ This
means that $T\cap (T+k)$ is exactly one point, which has address
$\overline{2}$ in $T$ and address $\overline{4}$ in $T+k.$ Indeed
the intersection is the point $k_2={1\choose 0}$ which is the
fixed point of $f_2,$ and $T+k$ will meet this point with the
translate of $k_4,$ the fixed point of $f_4.$ The translate of
$k_3$ along the vector $k'={-1\choose 1}$ is also $k_2.$ This way
it turns out that two of the six point neighbors will meet $T$ at
each vertex of the triangle.
\begin{figure}
\begin{center}
\includegraphics[width = 60mm]{siergra.eps}
\caption{\label{siergra} The reduced neighbor graph for
the Sierpi\'nski tile.}
\end{center}
\end{figure}
We now forget the point neighbors $\circ .$ The graph of the
remaining six $\bullet$ neighbors $\pm{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} , \pm\bb, \pm\cc$ and
the root, shown in Figure \ref{siergra}, will be called the
reduced neighbor graph \cite{bm}. On the left of Figure
\ref{siertile}, the corresponding boundary sets are indicated. As
mentioned above, each vertex $k=k_j-k_i$ has a loop with label
$i.$ The edge $(-\bb,{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} )$ has label 2 and the opposite edge has
label 3 because ${\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} =2\cdot(-\bb)+k_3-k_2.$ In a similar way, all
other edges were determined.
The graph in Figure \ref{siertile} is irreducible which means that
all corresponding boundary sets have a similar structure. This
also follows from the equations of the graph-directed system
(\ref{gradir}): \[ B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} =f_1(B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} )\cup f_2(B_{-\cc})\cup
f_3(B_{-\bb}), \ B_{-\bb}=f_2(B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} \cup B_{-\bb}\cup B_\cc), \
B_{-\cc}=f_3(B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} \cup B_\bb\cup B_{-\cc}) .
\] All these boundary sets are in fact small Sierpi\'nski gaskets,
as can be seen in the figure. The theorem below shows that they
are uncountable.
\end{Proof}
{\bf Topology and address map. } A sequence ${\bf s} } \def\tt{{\bf t} =s_1s_2...\in
I^\infty$ is called address of the point $x$ in a self-affine tile
$T$ if it is contained in the pieces $T_{s_1...s_n}$ for
$n=1,2,...$ (cf. \cite{bar}, Chapter IV, \cite{fa}, Section 9.1).
The points of the intersection sets $T_i\cap T_j$ have two
addresses, one starting with $i$ and the other one with $j.$ The
address map $\pi: I^\infty\to T$ is a quotient map, it determines
the topology of $T.$ For that reason it is important to know which
addresses will be identified. The neighbor graph provides this
information, as can be easily shown (\cite{bm}, sections 4 and 5).
A path in the directed graph $G$ is a finite or infinite sequence
$e_1e_2...$ such that the terminal vertex of $e_k$ coincides with
the initial vertex of $e_{k+1}.$ A finite path $e_1...e_n$ is a
cycle if the terminal vertex of $e_n$ is the initial vertex of
$e_1.$
\begin{Thm}(Neighbor graph and addresses of points, see \cite{bm})\\
Let $T$ be a self-affine tile with the address map $\pi:
I^\infty\to T$ and the neighbor graph $G=(V,E).$
\begin{enumerate} \item[ (i)] Two sequences ${\bf s} } \def\tt{{\bf t} =s_1s_2...$ and
$\tt=t_1t_2...$ are addresses of the same point if and only if
there is a path $e_1e_2...$ starting in the root with labels
$s_1s_2...,$ such that the opposite path -$e_1$-$e_2...$ has
labels $t_1t_2...$
\item[(ii)] The addresses of the points of a boundary set $B_k$
coincide with label sequences of paths starting in the vertex $k.$
They form a regular language $L_k.$
\item[(iii)] If only one cycle can be reached from the vertex $k$
by a directed path, then $B_k$ is a singleton.
\item[(iv)] If two different cycles can be reached from $k,$ and
these cycles can be reached from each other, then the set $B_k$ is
uncountable. \end{enumerate}
\end{Thm}
Note that a cycle with a diagonal, or with a loop at one of its
points counts as two different cycles which can reach each other.
\section{Existence of self-affine lattice tilings}
Before we can go on, we must settle one difficulty with neighbors.
They need not be compatible: $h(T)$ and $\tilde{h}(T)$ can be
neighbors of $T,$ but not of each other. In that case, different
tiles in a tiling will have different neighborhoods.
It turns out that the compatibility of neighbors is connected with
another question which was studied by Gr\"ochenig and Haas
\cite{gh} and in several papers of Vince \cite{v1,v4,v} as a
``central'' problem in the field. A self-affine lattice tile with
standard digit set will always admit self-affine tilings. It is a
non-trivial result that it will also admit lattice tilings
\cite{lw2} but these need not be self-affine in the sense that
tiles assemble to form supertiles (\cite{v}, Example 4.2). Vince
(\cite{v4},\cite{v}, Theorem 4.3) gave 10 equivalent conditions
for the existence of a self-affine lattice tiling. Two of them are
of an algorithmic nature.
Here we extend Vince's list by proving that a self-affine lattice
tiling exists if and only if all neighbors are compatible with
each other, and we show that this property can be easily decided
with the neighbor graph. The following theorem also shows that all
self-affine sets with mappings $f_j(x)=M^{-1}(x+k_j)$ which are
not lattice tiles must have incompatible neighbors.
\begin{Thm} (Compatible neighbors and self-affine lattice tilings)\\
Let $A\subset \R^n$ be a connected self-affine set with respect to
mappings $f_j(x)=M^{-1}(x+k_j), j=1,...,m$ where $M$ is an
expanding matrix and the open set condition holds. Then the
following conditions are equivalent.
\begin{enumerate}
\item[ (i)] All potential neighbors appear together at one piece $A_{\ii}.$
\item[(ii)] $A$ is a self-affine lattice tile which admits a
self-affine tiling by a lattice.
\item[(iii)] In the neighbor graph, each vertex $k\not= 0$ has at least
one incoming edge with each of the labels $j=1,...,m.$
\end{enumerate}\end{Thm}
\begin{Proof} (i)$\Rightarrow$(ii): We assume $A$ is a self-affine set, and
$A_{\ii}$ has all possible neighbors. Then the number of neighbors
must be finite, at most $m^n$ if $\ii=i_1...i_n.$ Moreover,
$f_{\ii}^{-1}(A)$ consists of copies of $A$ which include all
possible neighbors $h(A).$ Since the maps are translations
$h(x)=x+k$ by Proposition 3.1, let again $K$ denote the set of all
these translation vectors. We define an infinite pattern of
translates of $A$ as a union of increasing compact sets: \[
\bigcup_{q=1}^\infty f_{\ii}^{-q}(A) .\] All translates $A'$ of
$A$ in this pattern have the form $f_{\ii}^{-q}(A_{\ii{\bf j} } \def\ii{{\bf i} })$ for
some number $q$ and some word ${\bf j} } \def\ii{{\bf i} .$ This implies that they all
have the same neighbors as $A.$ To see this, consider the
(potential) boundary of a subpiece $A_{\ii j}$ with $j\in\{
1,...,m\},$ defined as the union of all its intersections with
potential neighbors. This boundary is contained in the union of
the boundary of $A_{\ii}$ and the intersections $A_{\ii j}\cap
A_{\ii i}$ with $i=1,...,m, i\not= j.$ So by assumption the
boundary of $A_{\ii j}$ is contained in $A,$ and by induction on
the length of ${\bf j} } \def\ii{{\bf i} $ this is proved for subpieces $A_{\ii{\bf j} } \def\ii{{\bf i} }.$
Thus all `atoms' $A'$ of our infinite pattern have a maximal set
of neighbors which covers their boundary completely. But there is
only one maximal set: the complete set of translates $A'+k$ with
$k\in K.$
Let $L$ denote the lattice generated by $K.$ Since $A$ was
connected, any two pieces of the same level are connected by a
chain of neighboring pieces (cf. \cite{br}, 8.2.1), and each atom
has the form $A'=A+k$ with $k\in L.$ Moreover, $L$ has the same
rank as the linear subspace generated by $A,$ which we will now
assume is $\R^n.$
Since all $A'$ have the neighbor vectors $K,$ the infinite pattern
coincides with $A+L.$ As a consequence, the pattern must be a
tiling of $\R^n:$ If there was a small open set $U$ outside $A+L$
then there would be arbitrary large copies $f_{\ii}^{-q}(U)$
outside $A+L$ which would contradict the finite mesh size of $L.$
Thus we have constructed a self-affine lattice tiling by copies of
$A.$
\smallskip
(ii)$\Rightarrow$(iii): Let $\R^n=A+L$ be a self-affine lattice
tiling by the tile $A.$ Then each tile of the tiling has the same
set of neighbor translations $K.$ Thus if $k$ is in $K,$ then $k$
appears as a neighbor translation of any piece $T_j$ of a
supertile $T,$ where $j\in\{ 1,...,m\}.$ If the neighbor is
another piece $T_i$ of $T,$ there is the edge $(0,k,j)$ in $G.$ If
the neighbor of $T_j$ is in another supertile $T'$ then we have
the edge $(k',k,j)$ where $k'$ is the neighbor map between $T$ and
$T'.$ Thus for each vertex $k$ of the vertex set $K,$ there is an
incoming edge with label $j.$\smallskip
(iii)$\Rightarrow$(i): We assume that each vertex $k$ in the
neighbor graph $G=(V,E)$ has incoming edges with each label $j.$
We have to find a word $\ii=i_1...i_N$ such that $A_\ii$ has all
neighbors $k\in V.$ For each $k$ there must be a suffix
$i_{n(k)}...i_N$ of $\ii$ which consists of labels of a path from
$0$ to $k.$
It is not difficult to construct such a word $\ii$ by induction,
going paths {\it backwards} (against the direction of edges)
towards 0. Start with a vertex $k_1$ and let ${\bf j} } \def\ii{{\bf i} _1=j_1...j_{q_1}$
be the labels of a backward path from $k_1$ to 0. According to our
assumption, there are also backward paths with label sequence
${\bf j} } \def\ii{{\bf i} _1$ from all other vertices. Their endpoints, which are
different from 0, will form a set $V_1.$
Next, take a $k_2\in V_1$ and a backward path
${\bf j} } \def\ii{{\bf i} _2=j_{q_1+1}...j_{q_2}$ from $k_2$ to 0. Also take backward
paths from all other $k\in V_1$ with label sequence ${\bf j} } \def\ii{{\bf i} _2$ and
denote the set of their non-zero endpoints by $V_2.$ Continue with
$k_3\in V_2,$ and so on. Since $V_{n+1}$ has less points than
$V_n,$ we will have $V_n=\emptyset$ for some $n.$ The sequence
$\ii=j_{q_n}...j_{q_{n-1}}...j_{q_1}...j_1$ will have all required
suffixes.
\end{Proof}
All examples considered in this paper will fulfil the condition of
theorem 5.1. For Example 4.4 this can be seen in Figure
\ref{siergra}, for the twindragons in the figures below.
\section{The seven three-dimensional twindragons}
{\bf Twindragons and their symmetry. } We now define the family of
examples we are going to study here. A twindragon is a self-affine
lattice tile with $m=2$ pieces. The expanding matrix $M$ has
determinant $\pm 2.$ In dimension 1, a twindragon is an interval.
In $\R^2,$ there are three examples: the rectangle (Example 2.1),
the ordinary twindragon and the tame twindragon, and they are all
disk-like \cite{bg}.
Since we identified conjugate tiles, we could take $k_1=0$ in
section 2. For twindragons in $\R^n,$ we can also choose $k_2$
arbitrarily, as long as $k_2$ is not in an invariant linear
subspace of $M.$ For the basis $\{ k_2, g(k_2),...,g^{n-1}(k_2)\}$
the matrix $M$ of $g$ will always be the same, determined by the
characteristic polynomial of $g$ (see \cite{gg1} and the proof of
Theorem 6.2 below). For our twindragons in $\R^3,$ we choose
another normalization:
\begin{equation}\label{knorm} k_1=(-\frac12 ,
0,0)', \quad k_2=(\frac12 ,0,0)' \,
\end{equation}
This has the effect that 0 is the symmetry center of $T,$ and in
the neighbor graph, an edge with label 1 passes from the root to
${\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} =(1,0,0)'.$
\begin{Prop} (Symmetry of twindragons)\\
Each twindragon $T$ in $\R^n$ has a symmetry center at
$c=\frac12(k_1+k_2).$ The point reflection at $c$ interchanges the
two pieces $T_1,T_2,$ and each boundary set $B_k$ with the
opposite set $B_{-k}.$ Moreover, each boundary set $B_k$ of $T$
also has a symmetry center at $c_k=c+\frac{k}{2}.$
\end{Prop}
\begin{Proof} We can take $k_1,k_2$ from (\ref{knorm}) since
the symmetry is not changed by an affine conjugacy. Thus
$f_1(x)=M^{-1}x+k_1, f_2(x)=M^{-1}x-k_1.$ With $\phi(x)=-x$ we
obtain $\phi f_1=f_2\phi$ or $f_i=\phi f_{3-i}\phi^{-1}$ since
$\phi^{-1}=\phi.$ For any word $\ii=i_1...i_n$ on $I=\{ 1,2\},$ we
have $f_\ii=\phi f_{{\bf j} } \def\ii{{\bf i} }\phi^{-1},$ where each $j_q=3-i_q$ is the
opposite symbol. Since the point of $T$ with address
${\bf s} } \def\tt{{\bf t} =i_1i_2...$ can be represented as $x=\lim_{n\to\infty}
f_{i_1}f_{i_2}...f_{i_n}(z)$ where $z$ is any starting point
\cite{fa}, this implies that $\phi$ transforms this point into the
point with the opposite address. In particular $\phi(T)=T$ and
$\phi(T_i)=T_{3-i}$ from which it follows that $\phi$ maps
$T_1\cap T_2$ onto itself.
For each boundary set $B_k=T\cap (T+k)$, the map $\phi$ transforms
$T+k$ to $\phi(T)-k=T-k,$ thus $\phi (B_k)=B_{-k}.$ On the other
hand we had $B_{-k}=B_k-k,$ so $\psi(x)=-x+k$ maps $B_k$ to
itself. $\psi$ is the point reflection at $c_k=\frac{k}{2}.$
\end{Proof}
{\bf The seven twindragons. } The last proposition shows that the
tile generated by $f_1,f_2$ can also be generated by $f_1\phi$ and
$f_2\phi,$ where $\phi$ is the point reflection at $c.$ Since in
$\R^3$ the map $\phi$ has determinant -1, this has the consequence
that we need only consider the matrices $M$ with $\det M=2.$ With
-2 we get the same twindragons.
\pagebreak
\begin{figure}
\begin{center}
\includegraphics[width = 150mm]{twin7.eps}
\caption{\label{twin7} Twindragon \BB, shown first in \cite{gg1}}
\end{center}
\end{figure}
\begin{Thm} (List of three-dimensional twindragons, cf.
\cite{klr,ag})\\
Up to conjugacy, there are exactly seven twindragons, each of
which is uniquely determined by the pair of coefficients $(a,b)$
of its characteristic polynomial
\[ p(\lambda)=-\lambda^3+a\lambda^2+b\lambda+2\, :\]
\begin{center}\begin{tabular}{|l|c|c|c|c|c|c|c|}\hline
Twindragon &$\mathcal A$\ } \def\BB{$\mathcal B$\ } \def\CC{$\mathcal C$\ &\BB&\CC&$\mathcal D$\ } \def\EE{$\mathcal E$\ } \def\FF{$\mathcal F$\ &\EE&\FF&$\mathcal G$\ \\ \hline
Parameter&$(0,0)$&(-1,1)&(1,-1)&$(0,1)$&(2,-2)&$(1,0)$&$ (0,2)$\\
\hline \end{tabular}
\end{center}
\end{Thm}
\begin{Proof}
These are the only integers $(a,b)$ for which $p(\lambda)$ has
only roots of modulus $>1$ \cite{klr,ag}. For $k_1=0,
k_2=(1,0,0)'$ and the basis $\{ k_2, Mk_2, M^2k_2\}$ the affine
map
\[ g(x)=\left(\begin{array}{lcr}0&0&2\\
1&0&b\\ 0&1&a\end{array}\right)\cdot x \] has characteristic
polynomial $p(\lambda ),$ and yields a twindragon. The
standardization (\ref{knorm}) translates this twindragon by
$(-\frac12 , 0,0)'$ so that its center is 0.
\end{Proof}
\begin{Exa} ( The non-fractal twindragon $\mathcal A$\ } \def\BB{$\mathcal B$\ } \def\CC{$\mathcal C$\ )\\
$\mathcal A$\ } \def\BB{$\mathcal B$\ } \def\CC{$\mathcal C$\ , the self-affine cube with 2 pieces, is the three-dimensional
analogue of Example 2.1. When considered as a rectangular
parallelepiped with side lengths $1, \sqrt[3]{2},\sqrt[3]{4},$ it
is the self-similar exception mentioned in Theorem 2.3, (ii). The
similarity map $g$ then is a $120^o$ rotation composed with a
homothety by the factor $\sqrt[3]{2}.$ Of course, this tile has 26
neighbors: 6 faces, 12 edge neighbors and 8 point neighbors.
\end{Exa}
\begin{figure}
\begin{center}
\includegraphics[width = 150mm]{twin2.eps}
\caption{\label{twin2} Twindragon \EE }
\end{center}
\end{figure}
{\bf Number of neighbors of twindragons. } The twindragons form a
nice little family with seven examples on which we can test our
neighbor methods. However, the six fractal examples are truly
self-affine, which makes their topology intricate. As a first
overview, we compare the moduli of eigenvalues of $M$ - the
inverses of the contraction factors of the $f_j$ along different
directions - and the number of neighbors. If the contraction
factors differ much, there are lots of neighbors. We also
determine the number of infinite boundary sets, by deleting from
the neighbor graph the vertices representing finite boundary sets,
as was explained in Example 4.4 and Theorem 4.5. It turns out that
all infinite boundary sets are uncountable. The number of faces
will be determined in the sections below.
\smallskip
\begin{Prop} (First neighbor calculation for the seven twindragons)
\begin{center}\begin{tabular}{|l|c|c|c|c|c|c|c|}\hline Twindragon &$\mathcal A$\ } \def\BB{$\mathcal B$\ } \def\CC{$\mathcal C$\ &\BB&\CC&$\mathcal D$\ } \def\EE{$\mathcal E$\ } \def\FF{$\mathcal F$\ &\EE&\FF&$\mathcal G$\ \\ \hline
Parameter&(0,0)&(-1,1)&(1,-1)&(0,1)&(2,-2)&(1,0)&(0,2)\\
\hline Complex eigenvalues'
modulus&1.26&1.29&1.22&1.15&1.14&1.09&1.06\\ \hline
Real eigenvalue&1.26&1.21&1.35&1.52&1.54&1.70&1.77\\
\hline Number of neighbors&26&18&20&34&34&48&76\\ \hline
Uncountable boundary sets&18&14&12&14&32&48&76\\ \hline
Faces&6&14&12&14&12&16&$\ge 22$\\ \hline
\end{tabular}\end{center}
\end{Prop}\smallskip
For \BB , \CC and $\mathcal D$\ } \def\EE{$\mathcal E$\ } \def\FF{$\mathcal F$\ the number of infinite boundary sets is
smaller than for the cube, so that we can expect them to be more
or less ball-like. \EE has medium complexity, and the last two
examples have many infinite neighbor sets -- they could be
three-dimensional counterparts of Example 4.4.
When we compare with the figures, we see that the twindragons
possess a kind of fibre structure along the direction of the
smaller eigenvalues of $M,$ that is, the larger eigenvalues of
$M^{-1}.$ For \BB , the pieces drawn in Figure \ref{twin7} are
stretched along the direction of the real eigenvalue of $M^{-1}$
which is the largest contraction factor. So roughly the pieces
look like `cigars'. In all other examples, the larger contraction
factor is given by the complex eigenvalues, and the pieces look
like leaves or plates. For \CC in Figure \ref{twin6}, the leaves
still have some thickness, while for \EE illustrated in Figure
\ref{twin2}, the plates are already very thin. For the other three
twindragons, the plates look similar or still thinner.\smallskip
{\bf How to draw neighbor graphs of twindragons. } Let us mention
some other nice properties of our family and fix some notation for
the neighbor graphs. First, all twindragons are connected. For two
pieces this follows from $T_1\cap T_2\not=\emptyset$ (see Barnsley
\cite{bar}, 8.2.1). If this relation would not hold, $T$ would be
a Cantor set and would have no interior points.
Next, the only possible differences $k_j-k_i$ are
$k_2-k_1=(1,0,0)', k_1-k_2=(-1,0,0)',$ and $k_i-k_i=0.$ The
corresponding labels $i$ of the edges are 1 and 2 and $i$ where
the last case means two edges labelled with 1 and 2. We shall draw
edges for the last case as double arrows, and we distinguish
labels 1 and 2 by assigning a fat tip to the arrows with label 2.
This way we save the labelling of edges. The label of the opposite
edge is $j=3-i$ for simple arrows, and $j=i$ for double arrows.
In view of the symmetry, we can simplify the neighbor graph
further by drawing only one vertex from each pair of opposite
vertices $k,-k$ and introducing the convention that a wavy arrow
$\leadsto$ to $k$ denotes an arrow to $-k.$ This will also
simplify the equation systems (\ref{gradir}) for the boundary
sets. It really matters whether one has to work with 20 or 40
vertices! Note that the opposite vertex of $k$ is $-k$ not only as
a vector, but by Proposition 6.1 also as a boundary set:
$B_{-k}=-B_k.$
For sake of brevity, we shall provide no details for the
complicated twindragons \FF and $\mathcal G$\ , but we give full arguments
for the other four examples.
To verify the condition of Theorem 5.1, we check in Figures
\ref{ng6}, \ref{ng7}, \ref{ng4}, and \ref{ng2} that each vertex is
reached either by a double arrow, or by two ordinary arrows with
different tips indicating label 1 and 2, or by an ordinary and a
wavy arrow with equal tips. In the last case the wavy arrow leads
to $-k,$ so we have to count the opposite edge.
\begin{figure}
\begin{center}
\hfill\begin{minipage}[b]{60mm}\includegraphics[width =
55mm]{ng6.eps}
\end{minipage}
\hfill
\begin{minipage}[b]{70mm}\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|}\hline
a&b&c&d&e&f&g&h&i&j\\ \hline 1&0&-1&0&0&1&1&-1&1&2\\ \hline
0&1&1&0&-1&-1&0&0&-2&-1\\ \hline 0&0&0&1&1&1&1&1&1&0\\ \hline
\end{tabular}\end{minipage}
\caption{\label{ng6} The neighbor graph for \CC contains all information
about the topology of Figure \ref{twin6}. The translation
vectors, which we shall not need,
indicate the position of the center of the neighbor.}
\end{center}
\end{figure}
\begin{Exa} We derive the equations (\ref{gradir}) of the boundary sets from
Figure \ref{ng6}. We neglect all point neighbors, given by
vertices below the thin line.
\[ B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} =f_1(B_\bb)\cup f_2(B_\bb)\cup f_2(B_\cc),\quad
B_\bb=f_1(B_{\rm d})\cup f_2(B_{\rm d}), \quad B_{\rm
d}=f_2(B_{\rm f}),\]
\[ B_\cc=f_1(B_{\rm e})\cup f_2(B_{\rm e})\cup
f_1(B_{\rm f}),\quad B_{\rm e}=f_2(-B_\cc ), \quad B_{\rm
f}=f_2(B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} )\, .\] By substitution we reduce the system to two
equations:
\[ B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} =f_2(B_\cc)\cup\bigcup_{i,j=1}^{2} f_{ij22}(B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} )\, ,\qquad
B_\cc=f_{12}(B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} )\cup\bigcup_{i=1}^{2} f_{i2}(-B_\cc)\, .\]
\end{Exa}\smallskip
\begin{figure}
\begin{center}
\includegraphics[width = 65mm]{ng7.eps}
\caption{\label{ng7} Neighbor graph for
the twindragon \BB. There are two point neighbors: i with address
$\overline{2}$ and h with address $1\overline{2}$.}
\end{center}
\end{figure}
\section{Identifying faces: the simple case}
{\bf Defining faces. } The topological boundary of a set $T$ will
be denoted by $\partial T= T\setminus{\rm int}\, T.$ If $T$ is a tile in
a tiling, then $\partial T$ must be covered by the boundary of
neighboring tiles: $\partial T=\bigcup_{k\in V} B_k.$
Theorem 4.5 says that we can determine the cardinality of a
boundary set $B_k$ by counting the number of paths starting from
$k$ in the neighbor graph $G.$ It is much more complicated to
determine the topology of $B_k.$ In three-dimensional self-affine
tiles, we would like to distinguish boundary sets which are
two-dimensional fractals, one-dimensional curves, Cantor sets or
just finite sets.
Let us say that the boundary set $B_k=T\cap (T+k)$ is a {\it face}
of $T$ if there is a point $x\in B_k$ and an open neighborhood $U$
of $x$ with $U\cap\partial T\subset B_k$ and $U\subset T\cup
(T+k).$ In that case, all boundary points of $U\cap {\rm int}\, T$ must
belong to $B_k.$ Thus $B_k$ has topological dimension 2 since the
boundary of each open set in $\R^n$ has dimension $n-1.$ The
following statement shows that we have six faces in Example 4.4.
\begin{Prop} Let $k\not= 0$ be a vertex in $G$ such that for each
other vertex $k'\not= 0,$ either $k'$ or the opposite vertex $-k'$
can be reached by a path from $k.$ Then $B_k$ is a face of $T.$
\end{Prop}
\begin{Proof} Each tile must have faces. By (\ref{gradir}),
any path from $k$ to $k'$ implies the existence of a small copy of
$B_{k'}$ in $B_k.$ Thus our assumption implies that $B_k$ contains
copies of faces, so it must be a face (cf. Proposition 8.1).
\end{Proof}
For twindragons, there is always an edge with label 1 from the
root to the first other vertex ${\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} =k_2-k_1=(1,0,0)'.$ and this
must be a face by Proposition 7.1. It turns out that for the
twindragons \BB, \CC and $\mathcal D$\ } \def\EE{$\mathcal E$\ } \def\FF{$\mathcal F$\ , all infinite boundary sets fulfil
the assumption of the proposition since from their vertices in
$G,$ there is a path to ${\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} .$ The neighbor graphs are shown in
Figures \ref{ng6}, \ref{ng7} and \ref{ng4}. The point neighbors
are separated from the faces by a thin line. They are found by
Theorem 4.5, (iii).
For \BB this can be seen in Figure \ref{ng7}. For \CC we have the
loop at g with address $\overline{2},$ and the 2-cycle $\{{\rm i}
, {\rm h}\}$ in Figure \ref{ng6} which denotes a 4-cycle $\{ {\rm
i, -h, -i, h}\}$ in the complete neighbor graph, with address
$\overline{2112}.$ (The wavy arrow indicates that we go to the
opposite sets, and there the labels 1 and 2 have to be
interchanged. A second wavy arrow reestablishes the original
labels.) In Figure \ref{ng4} we have a similar 4-cycle $\{ {\rm
q,s,j,n}\}$ which denotes an 8-cycle of point neighbors in the
complete neighbor graph with address $\overline{21221211}.$ Also h
and m are point neighbors, while i,r,p and k describe finite
boundary sets with more than one point. We summarize:
\begin{Prop} (Faces of \BB, \CC and $\mathcal D$\ } \def\EE{$\mathcal E$\ } \def\FF{$\mathcal F$\ )\\
The twindragon \CC has 12 faces and 8 point neighbors.
\BB has 14 faces and 4 point neighbors. The twindragon $\mathcal D$\ } \def\EE{$\mathcal E$\ } \def\FF{$\mathcal F$\ has 14
faces, 12 point neighbors and 8 further finite boundary sets.
\end{Prop}
\begin{figure}
\begin{center}
\includegraphics[width = 55mm]{ng4.eps}
\caption{\label{ng4} Neighbor graph for
the twindragon $\mathcal D$\ } \def\EE{$\mathcal E$\ } \def\FF{$\mathcal F$\ . The neighbors i,r,p and k describe
finite boundary sets with more than one point.}
\end{center}
\end{figure}
\section{Different types of boundary sets}
{\bf The partial order for boundary sets. } To compare different
infinite boundary sets, we define an order on the vertices of $G.$
For $u,v\in V$ we write $u\succ v$ if there is a directed path
from $u$ to $v.$ We say that the vertices are equivalent, and
write $u\sim v$ if either $u=v$ or both $u\succ v$ and $v\succ u.$
For the boundary sets, $u\succ v$ means that $B_u\supset f_\ii
(B_v)$ for some word $\ii=i_1...i_q.$
\begin{Prop} If $u\succ v$ then $\dim B_u\ge \dim B_v$ holds for
the topological dimension, the Hausdorff dimension and the box
dimension. Moreover, if $B_v$ is a face, then also $B_u$ will be a
face.
\end{Prop}
\begin{Proof} The dimension concepts mentioned here (and also some
others \cite{fa}) have the following properties.
\begin{enumerate}
\item[ (i)] If $A\subseteq B$ then $\dim A \le \dim B .$
\item[(ii)] If $h$ is an affine map then $\dim h(B)=\dim B.$
\end{enumerate}
Together with the remark above this proves the first assertion.
\\ If $B_v=T\cap T+v$ is a face, there is $x\in B_v$ with an open
neighborhood $U\subset T\cup T+v .$ Now $u\succ v$ means that
$B_u\supset f_\ii (B_v)$ for some word $\ii$ where $T+u\supset
f_\ii (T+v).$ So $y=f_\ii (x)$ and its neighborhood $f_\ii (U)$
can be taken to show that $B_u$ is a face.
\end{Proof}
{\bf Strong components of the neighbor graph. } We can now
classify the vertices of $G$ in a similar way as states of a
Markov chain. Clearly $\sim$ is an equivalence relation. The
topological and Hausdorff dimension, as well as the property of
describing a face, are invariant with respect to $\sim .$ Let
$\tilde{V}$ be the set of equivalence classes with respect to
$\sim$ which are called strong components of $G.$ Then we get a
new graph $\tilde{G}$ by drawing an edge from
$(\tilde{u},\tilde{v})$ if there is $u\in\tilde{u}$ and
$v\in\tilde{v}$ such that $(u,v)$ is an edge in $G.$
The graph $\tilde{G}$ compresses the structure of $G.$ If two
vertices $u,v\in G$ belong to the same class, this means that
$B_u$ contains a copy of $B_v$ and conversely, so both boundary
sets have essentially the same structure. The root $0\in V$ would
also be the root of $\tilde{G}$ but it has no meaning and is
omitted.
The relation $\succ$ is a partial order on $\tilde{V}.$ If
$\tilde{u}\succ\tilde{v}$ then all boundary sets $B_u$ contain
copies of all boundary sets $B_v,$ and $\dim B_u\ge\dim B_v$ for
all choices $u\in\tilde{u}, v\in\tilde{v}.$ Thus the largest
classes of boundary sets are the faces, then we have
one-dimensional sets, Cantor sets, and as terminal classes
countable and finite sets. Boundary sets from classes
$\tilde{u},\tilde{v}$ with $\tilde{u}\succ\tilde{v}$ can also have
equal dimensions, however.
\begin{Rem}
Consider a class $\tilde{u}\in \tilde{V},$ as a subset of $V.$
There are three cases.
\begin{enumerate} \item[ (i)] $\tilde{u}$
contains only one vertex. Then the boundary set $B_u$ has the same
dimension as the successor vertex $v$ (in case of several
successors, the union of the successor boundary sets). Namely
$B_u=f_i(B_v).$
\item[(ii)] $\tilde{u}$ is a cycle in $G,$ without loops and diagonals.
In that case the boundary sets $B_u$ with $u\in \tilde{u}$ are
countable unions of copies of the successor boundary sets.
\item[(iii)] $\tilde{u}$ contains two cycles. Only in this case $\tilde{u}$
produces its own Cantor structure, and the $B_u$ can have larger
dimension than the successor sets.
\end{enumerate}
\end{Rem}
In Figures \ref{ng6}, \ref{ng7}, and \ref{ng4} we have one class
of faces, a number of singleton classes and a terminal cycle which
is a class of two, one or four points. In Figure \ref{ng2} we have
a class of six faces, and a class of eight infinite neighbors h..r
on the right, as well as three singleton classes. The vertex s
describes a point neighbor, the vertices g and p correspond to
infinite boundary sets with equal dimensions as h..r, by the above
remark, (i). We now have to decide whether g,p and h..r are faces.
\section{Identifying faces: the general case}
{\bf An algorithmic approach to find the faces. } When we consider
one equivalence class of the neighbor graph which contains a face,
then it consists of faces only.
\begin{Prob} Must all faces belong to one equivalence class of $G$
?\end{Prob}
This problem also arises when we work with the contact matrix. In
\cite{al}, irreducibility was just assumed. We present methods to
decide the problem algorithmically. We start with a simple
topological observation.
\begin{Prop} The faces cover the boundary of a self-affine lattice
tile $T.$
\end{Prop}
\begin{Proof} Let $x$ be a boundary point of $T$ and $U$ an open
neighborhood of $x.$ Let $E=U\cap\partial T$ and let $T_k,
k=1,...,q$ denote the neighbor tiles of $T.$ The sets $T\cap T_k$
cover $E,$ and by the Baire category theorem one of these sets,
say $T\cap T_1,$ must contain an open subset $V$ of $E.$ Each
$y\in V$ has a neighborhood $W$ in $\R^n$ which is contained in
$T\cup T_1.$ So $T\cap T_1$ is a face. Since this holds for any
$U,$ the point $x$ is an accumulation point of the faces. Since we
have only finitely many faces, $x$ must belong to one of them.
\end{Proof}
\begin{Thm} (The neighbor graph decides which boundary sets are faces)\\
Let us assume that all neighbors of the tile $T$ are compatible.
Then a boundary set $B_u$ is a face if it is not contained in the
union of all other boundary sets.\\
For the neighbor graph this means that there is a word $\ii$ which
belongs to the language $L_u$ of labels of paths starting in $u,$
but not to $L_v$ for any other $v\not= u.$
\end{Thm}
\begin{Proof} Consider the union of the tile $T$ and all its
neighbor tiles. If $B_u$ is not a face, then by the proposition it
is contained in the union of the remaining boundary sets. If
$B_u$ is a face, there is $x\in B_u$ and a neighborhood $U\subset
T\cup T+u$ of $x.$ Then $U$ cannot intersect any other neighbor
$T+v$ since neighbors must have no common interior points. The
first part of the theorem is proved.\\
Now $x$ has an address: $x=\pi ({\bf s} } \def\tt{{\bf t} )$ where $\pi$ is the
projection from the symbol space $I^\infty$ to $T.$ Since $\pi$ is
continuous \cite{br,fa}, there is a prefix $\ii$ of $x$ such that
all addresses $\tt$ which begin with $\ii$ will fulfil $\pi
(\tt)\in U.$ By Theorem 4.5 (iii), this shows that for a face
$B_u$ there is a word $\ii$ which is contained in $L_u$ and in no
other $L_v.$\\
Conversely, suppose $B_u$ is not a face and $\ii\in L_u.$ There is
an address ${\bf s} } \def\tt{{\bf t} $ which begins with $\ii$ such that $\pi({\bf s} } \def\tt{{\bf t} )=x$ is
not on the intersection of different pieces $f_j(T).$ (It is known
that the intersections have Lebesgue measure 0 \cite{bg}, and
since the normalized Lebesgue measure on $T$ can be considered as
the image measure of the product measure
$\{\frac{1}{m},...,\frac{1}{m}\}$ on $I^\infty,$ no cylinder can
be mapped completely into the intersections.) Since $x$ is also in
a face $B_v$ and has only one address, $L_v$ contains the sequence
${\bf s} } \def\tt{{\bf t} $ starting with $\ii .$
\end{Proof}
{\bf Examples of address calculations for boundary sets. } We
shall apply our theorem to Figure \ref{ng2}. Address calculations
are boring and best left to the computer, but they can provide
valuable information about the intersection of pieces. We start
with simple examples.
\begin{figure}
\begin{center}
\includegraphics[width = 75mm]{ng2.eps}
\caption{\label{ng2} Neighbor graph of
twindragon \EE. The strong component of a..f represents
faces, s is a point neighbor. Vertices h..r form a strong component
representing infinite boundary sets. g and p are components with
boundary sets of the same dimension as h..r.}
\end{center}
\end{figure}
\begin{Exa} ( Point neighbors in \CC and \BB)\\
Which faces contain the point neighbors of \CC and $ \mathcal B ?$
In Figure \ref{ng6} we find
\[ B_{\rm g}=\pi(\overline{2})=B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} \cap B_\bb \cap B_\cc \cap
B_{\rm d}\, , \qquad B_{\rm h}=\pi(\overline{2211})\subseteq
B_\cc\cap B_{\rm d}\cap B_{\rm e}\quad \mbox{ and} \]
\[ B_{\rm i}=\pi(\overline{2112})=B_{-\cc}\cap B_{\rm e}
\cap B_{\rm f}\, , \qquad B_{\rm
j}=\pi(2\overline{1221})\subseteq B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} \cap B_{-\cc}\] so g,h,i
seem proper vertices but j seems to be on an edge. On the other
hand, both point neighbors of Figure \ref{ng7} coincide with the
intersection of only two faces:
\[ B_{\rm i}=\pi(\overline{2})=B_{\cc}\cap B_{\rm e}, \qquad
B_{\rm h}=\pi(1\overline{2})=B_{{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} }\cap B_{\cc}\ .
\]
\end{Exa}
Already these calculations show the strange polyhedral structure
of twindragons. For Figure \ref{ng2} the boundary sets look much
more complicated than for ordinary polyhedra where an edge is
covered by two faces. The proof of the next proposition will also
reveal hidden symmetry behind the apparently orderless structure
of $G.$
\begin{Prop} ( Boundary structure of \EE)\\
The vertices {\rm g, p, h..r} in the neighbor graph of the
twindragon \EE do not represent faces. The boundary set {\rm i} is
contained in the union of the four faces {\rm e, -f, c, -a} and
not in a union of three of them.
\end{Prop}
\begin{Proof}
We determine the language $L_{\rm i}$ from the three elementary
cycles which lead in $G$ from i to either i or -i (cf. section 7
and Example 9.4).
\begin{equation}\label{Li} L_{\rm i}= *2111 L_{\rm i} \cup
12 L_{\rm -i} \cup ***1122 L_{\rm -i}= C_1L_{\rm i}\cup C_2L_{\rm
i}\cup\bigcup_{j=0}^4 D_jL_{\rm -i}
\end{equation} where $*$ is considered as a wildcard for both 1
and 2, and $C_1=12111, C_2=22111, D_0=12, D_1=1*11122,
D_2=1*21122, D_3=2*11122, D_4=2*21122.$ We now consider the set of
faces $J=\{ {\rm e, -f, c, -a}\} .$ We will show
\begin{equation}\label{li}
L_{\rm i}\subset L_J=L_{\rm e}\cup L_{\rm -f}\cup L_{\rm c}\cup
L_{\rm -a}.
\end{equation}
First we study the action of $C_1$ and
$C_2$ on the vertices of $J.$ In the following table $+$ means
there is path from x to y with label $C_1,$ and $*$ means the same
for $C_2.$
\begin{center}\begin{tabular}{|l|c|c|c|c|}\hline
x\ $|$y&e&-f&c&-a\\ \hline e&$*$&&$*$&\\ \hline -f&&$+$&&$+$\\
\hline c&&$*$&&$*$\\ \hline -a&$+$&&$+$&\\ \hline \end{tabular}
\end{center}
{\it Claim 1: } any word $C\in \{ C_1,C_2\}^*$ is in $L_J,$ and it
can be realized with any prescribed terminal point in $J.$\\
This claim can be proved by induction on the number $t$ of terms
in $C=C_{q_1}^{n_1}...C_{q_t}^{n_t},$ using the table. The
starting point always depends on the word. For instance, it must
be -f for $C=C_1^3.$
Now we study how the $D_j, j=0,..,4$ lead from $J$ to $-J.$ In $G$
it can be seen that $D_0$ leads from e to -e, -f to f, c to a, and
-a to -c, respectively. Moreover, $D_1$ leads from -a to all
vertices of $-J,$ and from no other vertex of $J$ to any vertex of
$-J.$ Similarly, $D_2$ leads exclusively from -f to all of $-J,$
and $D_3$ leads exclusively from e to all of $-J,$ and $D_4$ leads
exclusively from c to all of $-J.$ From this observation it
follows that all vertices of $J$ are needed to obtain $L_{\rm
i}\subset L_J.$ Moreover, it is clear that
{\it Claim 2: } any word $D_j, j=0,..,4$ can be realized by a path
from $J$ to $-J$ with any prescribed terminal point in $-J,$
provided we can select the starting point.
Applying Claims 1 and 2 alternatingly, to $J$ and its opposite set
$-J,$ we complete the proof of (\ref{li}). The rest follows from
Theorem 9.3.
\end{Proof}
\section{Dimensions and eigenvalues}
{\bf The modified Hausdorff dimension. } In this section we give
an alternative proof for the fact that \EE has only 12 faces. It
is based on recent work of He and Lau \cite{hl} and of Akiyama and
Loridant \cite{al}. Let $H$ denote the adjacency matrix of the
graph $G\setminus \{ 0\}.$ For two non-zero vertices $u,v$ the
entry $h_{uv}$ denotes the number of edges leading from $u$ to
$v.$ It is also possible to take the adjacency matrix of our
simplified graphs, where only one vertex from each pair $\{
k,-k\}$ is taken into account and $h_{uv}$ counts the edges from
$u$ to $\{ v,-v\}.$ Since $H$ is a non-negative matrix it has a
real eigenvalue of maximum modulus which we call $\lambda .$
When the matrix $M$ defining the affine maps $f_j$ is conjugate to
a similarity matrix with expanding factor $R$ (see Proposition
2.2) it was shown by Mauldin and Williams \cite{mw} that the
Hausdorff dimension of the sets $B_k$ from (\ref{gradir}) fulfils
\begin{equation}\label{dimb}
\dim B_k \le \frac{\log\lambda}{\log R}\ ,
\end{equation}
and equality holds if an open set condition is satisfied, see the
discussion at the end of this section.
For self-affine sets, however, calculation of dimensions is much
more complicated then for self-similar ones. However, for our case
where all $f_j$ are defined by one matrix $M,$ He and Lau
\cite{hl} have defined a pseudo-norm $w$ on $\R^n$ which turns the
$f_j$ into similarity mappings, with factor $r=1/\sqrt[n]{\det
M}.$ Moreover, $w$ fulfils a weak type of triangle inequality
which is sufficient to develop Hausdorff measure and dimension as
usual. And $w$ generates the Euclidean topology even though it
drastically distorts the geometry of $\R^n.$ The value of the
modified Hausdorff dimension $\dim^w B_u$ was used in \cite{hl} to
give estimates of the real Hausdorff dimension in terms of
eigenvalues of $M.$ Akiyama and Loridant \cite{al} applied the
modified Hausdorff dimension to boundary sets of tiles. Here we
shall compare it with the topological dimension.
\begin{Thm} (Upper estimate of topological dimension of boundary sets)\\
Suppose that $T$ is a self-affine tile with $m$ pieces in $\R^n,$
and all neighbors are compatible. Let $W$ denote a strong
component of the graph $G,$ and let $\lambda_W$ denote the largest
eigenvalue of the adjacency matrix $H_W$ of $W$ as a subgraph of
$G.$ If for some integer $q<n,$
\[ \frac{n\log\lambda_W}{\log m}< q \]
then the topological dimension of all boundary sets $B_v$ with
$v\in W$ is strictly smaller than $q.$
\end{Thm}
\begin{Proof}
The matrix $H_W$ is irreducible, and $W$ is non-periodic (i.e. the
g.c.d. of the lengths of cycles from a point $v\in W$ to itself is
1) because of the compatibility of neighbors. So $\lambda_W$ is
the unique largest eigenvalue. The Hausdorff dimension of the sets
$B_v$ with $v\in W$ with respect to the He-Lau pseudo-norm is $\le
\frac{\log\lambda_W}{\log R}=\frac{n\log\lambda_W}{\log m}$ since
$R=\sqrt[n]{m}$ \cite{al,hl}. The pseudo-norm generates the
Euclidean topology, and the topological dimension is always
smaller than the Hausdorff dimension.
\end{Proof}
\begin{Exa} ( Boundary sets of the twindragon \EE)\\
For \EE we determine the dimension of the boundary sets $B_v$ with
$v\in\{\rm g, p, h..r\} .$ We obtain $\lambda_W \approx 1.554.$ A
small polynomial for $\lambda_W$ can be obtained from equation
(\ref{Li}) for $B=B_{\rm i}$ which expresses $B$ as a self-affine
set
\[ B= \bigcup_{i=1}^2 f_{i2111}(B)\cup
f_{12}(-B)\cup\bigcup_{i,j,k=1}^2 f_{ijk1122}(-B)\ . \] The
characteristic equation is $1=2z^5 +z^2+8z^7,$ which for
$\lambda=1/z$ becomes \ $\lambda^7-\lambda^5-2\lambda^2-8=0.$
The modified Hausdorff dimension is $\log\lambda_W
/\log\sqrt[3]{2} \approx 1.908.$ Thus the topological dimension of
the $B_v$ is $\le 1.$ So the $B_v$ cannot be faces. A calculation
for the faces {\rm a..f} gives modified dimension 2.67.
\end{Exa}
For the complicated twindragons \FF and $\mathcal G$\ we also computed the
strong components and their modified dimensions. For $\mathcal G$\ , the
above theorem does not apply since beside the irreducible
component of 22 faces with modified dimension 2.87, there is
another component of 20 boundary sets, plus 2 singleton
components, all with modified dimension 2.13. All the remaining 32
boundary sets have modified dimension 1.52. This boundary seems
really complicated! The other example is less intricate. We omit
the details of the following example.
\begin{Exa} ( Boundary sets of the twindragon \FF )\\
For \FF, the neighbor graph $G\setminus\{ 0\}$ has an irreducible
component $V_1$ of 16 faces with modified dimension 2.78, and the
other 32 neighbors are not faces. 10 of them form a cycle without
diagonals as second component $V_2,$ and there are two singleton
components between $V_1$ and $V_2.$ Since the cycle $V_2$ consists
of six double arrows and 4 simple arrows, $\lambda_2=2^{3/5}$ and
$3*\log\lambda_2/\log 2= 1.8.$ Due to the theorem, the topological
dimension is at most 1.
Moreover, there is another 10-cycle $V_3$ with only two double
arrows and $\lambda_3=2^{1/5},$ and 10 singleton components
between $V_2$ and $V_3.$ So the 20 corresponding boundary sets
have modified dimension $3*\log\lambda_3/\log 2= 0.6$ and by
Theorem 10.1 must be Cantor sets.
When one considers the labels along the cycles $V_2$ and $V_3,$
one obtains the languages
\[ L_2=\overline{*1**1*2**2}\quad\mbox{ and }\quad
L_3=\overline{11*1122*22} \] and their cyclic permutations as
address sets of the corresponding boundary sets. From this it is
easy to conclude that each Cantor set associated to a vertex in
$V_3$ is the intersection of two respective boundary sets in
$V_2.$\end{Exa}
We did not use the open set condition of the boundary equations
(\ref{gradir}), and this condition need not always be fulfilled.
In the cube (Example 6.3), for instance, some of the equations
(\ref{gradir}) contain equal terms since an edge is counted two
times. However, the open set condition will be true if we remove
such double sets and if the neighbors are all compatible. In
particular it is true when we restrict ourselves to faces
\cite{al}. When the graph is a cycle, as $V_2$ and $V_3$ in the
last example, the open set condition also holds.
\section{Polyhedral structure of tiles}
{\bf Homeomorphism with a polyhedron. } We have seen that the
structure of boundary sets can be quite complicated. We are
interested in the presence of simple structure, however. Let us
say that $T\subset\R^3$ {\it has the structure of polyhedron} if
$T$ is homeomorphic to a ball in $\R^3,$ the faces are
homeomorphic to disks in $\R^2,$ and their non-empty intersections
are either singletons or homeomorphic to an interval - in such a
way that $T$ together with its boundary structure becomes
homeomorphic to a polyhedron in Euclidean $\R^3.$\smallskip
This condition is stronger than just requiring that $T$ is
homeomorphic to a ball. However, this is the structure which
crystallographers expect of a tiling: ``We define a tiling as a
periodic subdivision of space into bounded, connected regions
without holes, which we call tiles. If two tiles meet along a
surface, we call the surface a face. If three or more faces meet
along a curve, we call the curve an edge. Finally, if at least
three edges meet at a point, we call that point a vertex''
\cite{fdhkm}.\smallskip
{\bf Definition of edges and vertices. } We follow this quotation
and define a polyhedral structure on every self-affine tile. We
remove from $G$ all vertices which do not represent faces. Then we
define the graph $G^2$ of edges - that is, of intersection sets
$B_k\cap B_{k'}.$ For these, it is not so difficult to check
whether they are homeomorphic to intervals \cite{bm}. If
necessary, we can also determine the graph $G^3$ of vertices of
the tiling. Euler's polyhedra formula can then be used to check
whether $T$ is homeomorphic to an ordinary polyhedron. The method
is quite fast in showing that certain tiles are {\it not}
ball-like.
A definition similar to the following was given by Thuswaldner and
Scheicher \cite{st} but it was not applied to examples.
\begin{Def}\label{defng2}
Let $T\subset \R^n$ be a self-affine lattice tile with neighbor
graph $G=(V,E=\bigcup_{i=1}^m E_i).$ For $\ell =2,3$ the graph of
$\ell$-intersections of boundary sets of $T$ is
\[ G^\ell =(V^\ell ,E^\ell ) \quad\mbox{ with }\quad
V^\ell =\left\{ K=\{k_1,...,k_\ell\}\, |\mbox{ all }k_i\mbox{
different, } B_{k_1}\cap ...\cap B_{k_\ell}\not=\emptyset\right\}
\]
\begin{equation}\mbox{ and }\qquad E^\ell =\bigcup_{i=1}^m E^\ell_i\quad
\mbox{ with } \label{neigra2}\end{equation}
\[ E^\ell_i=\{ (K,K',i)\, |\mbox{
there is a 1-1-map }\phi:K\to K' \mbox{ with }(k,\phi(k),i)\in
E_i\mbox{ for }k\in K\} \]\end{Def}
The idea is that a neighbor intersection $(T+k)\cap (T+k')$ has a
piece which is at the boundary of $T_i$ if and only if both
neighbor sets have such a piece, and the pieces do intersect. The
algorithmic check for non-empty intersection is the same as for
$G.$ We start with all $\ell$-subsets of $V,$ and step by step we
remove those which have no outgoing vertices to the remaining
vertex set. Since for $\ell=2,3$ we must have non-empty
intersections, we end with a graph where each set has outgoing
vertices, and hence `infinite paths' of edges (counting repeated
use of cycles). The properties of $G^\ell$ are similar to those of
$G.$
\begin{Prop} $G^\ell$ provides a graph-directed fractal structure
on the $\ell$-intersec\-tions of neighbors. For each vertex $K\in
V^\ell,$ the labels of infinite paths in $G^\ell$ with starting
point $K$ are the addresses of the corresponding boundary
intersection $\bigcap_{k\in K} B_k .$ The cardinality of this
intersection can be determined as in Theorem 4.5.
\end{Prop}
\begin{Exa}
We continue Example 4.4 by determining $G^2$ for the six faces of
the Sierpi\'nski tile, see Figure \ref{siertile}. Since the
outgoing edges from $-{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} , -\bb, -\cc$ all start with $4,2,$ and
3, respectively, the corresponding boundary sets have pairwise
empty intersection. The intersection of a boundary set with its
opposite is also empty. We get edges
\[\{{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} , -\bb\}\stackrel{2}{\to}\{{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} , -\cc\}\stackrel{3}{\to}
\{{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} , -\bb\}\] so $B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} \cap B_{-\bb}$ is a singleton with
address $2323...=\overline{23}$ and $B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} \cap B_{-\cc}$ has
address $\overline{32}.$ The vertices ${\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} ,\bb,\cc$ all have a
loop with label 1, so the corresponding three boundary sets meet
in the point with address $\overline{1}.$ Altogether, $G^2$
consists of three cycles of length 2, and three isolated loops.
$G^3$ has only one vertex with a loop, representing the center
point of the tile.
\end{Exa}
\section{Polyhedral structure of twindragons}
{\bf Intersections between antipodal surfaces. } The first
information which the face intersection graph $G^2$ provides is a
list of non-empty intersections of pairs of faces. As in Theorem
4.5, it is easy to decide whether such an intersection consists
just of one point.
\begin{Prop}
Two faces $B_u, B_v$ with $u,v\in V$ intersect each other if and
only if $\{ u,v\}$ is a vertex in $G^2.$ Moreover, if there is
only one infinite directed path in $G^2$ starting in the vertex
$\{ u,v\} ,$ then $B_u\cap B_v$ is one point, and the address of
this point can be read from the labels of the path.
\end{Prop}
For our twindragons we check whether there is a face $B_k$ which
intersects its opposite face $B_{-k}.$ If such intersections
exist, it is unlikely that our tile has the structure of a
polyhedron. Actually, this happens for four of our twindragons and
provides an argument to show that $\mathcal D$\ } \def\EE{$\mathcal E$\ } \def\FF{$\mathcal F$\ and \EE are not homeomorphic
to a ball.
A set $A\subset \R^n$ is called simply connected if it is
connected and its homotopy group is trivial. The latter means that
$A$ has ``no holes'' -- any closed curve $C\subset A$ can be
contracted within $A$ to a point.
\begin{Thm} (Tiles for which the interior is not simply
connected)\\
Let $T$ be a connected self-affine lattice tile which has a
symmetry center $c$ which is not an interior point of $T.$ Then
the interior of $T$ is not simply connected.\\
In particular, the interior of $\mathcal D$\ } \def\EE{$\mathcal E$\ } \def\FF{$\mathcal F$\ and \EE is not simply
connected. These twindragons are therefore not homeomorphic to a
ball.
\end{Thm}
\begin{Proof} We can assume $c=0$ and thus $T=-T$ by changing the maps as in
Proposition 6.1. If 0 is not an interior point of $T,$ there is
some tile $T+k$ in the tiling which contains 0, and by symmetry
also $-(T+k)=T-k$ does not intersect ${\rm int}\, T$ and contains 0.
Take a big sphere $S$ around 0 which contains the three tiles in
its interior. Let $z\in S$ be of the form $z=t\cdot k, t>0.$ There
is a line segment $\alpha$ from $z$ to the point $y\in T+k$ which
maximizes the scalar product $<y,k>.$ Since $T$ is connected, it
is arcwise connected, and there is an arc $\beta\subset T+k$ from
$y$ to 0. The union $\varepsilon$ of the arcs $\alpha , \beta ,
-\beta, -\alpha$ is an arc in $S$ from $z$ to $-z$ which does not
intersect ${\rm int}\, T.$
Now assume ${\rm int}\, T$ is connected and take a point $x\in
{\rm int}\, T.$ The opposite point $-x$ is also in ${\rm int}\,
T.$ So there is a connecting arc from $x$ to $-x.$ Covering this
arc by finitely many open balls inside ${\rm int}\, T,$ we see
that we can replace it by a union $\gamma$ of finitely many line
segments, and we can choose them so that no two segments are
parallel. Then the union $\gamma\cup -\gamma$ is a closed curve
from $x$ through $-x$ to $x$ with a finite number of
self-intersections. If $x_1$ is the first self-intersection point,
we find a closed arc from $x_1$ through $-x_1$ to $x_1$ with two
self-intersection points less. After a finite number of steps we
get a simple closed curve $\delta$ in ${\rm int}\, T$ from some
$x_n$ through $-x_n$ to $x_n.$ By construction, $\delta$ surrounds
the arc $\varepsilon$ and thus cannot be contracted within ${\rm
int}\, T$ to a single point. So ${\rm int}\, T$ is not simply
connected. The second assertion will be proved in the example
below.
\end{Proof}
\begin{Prop}
If an infinite path in the neighbor graph $G$ of a twindragon
starts in the root and contains no double arrow, it defines an
address of the center 0 of $T.$
\end{Prop}
\begin{Proof}
For a path of single arrows which addresses a point $x,$ the
opposite path is obtained by just interchanging labels 1 and 2. In
the proof of Proposition 6.1, it was shown that the address of the
opposite path corresponds to the point $-x.$ For paths starting
in the root of $G,$ Theorem 4.5 (i) says that both addresses
belong to the same point. $x=-x$ implies $x=0.$
\end{Proof}
\begin{Exa} ( Faces intersecting their opposite face in $\mathcal D$\ } \def\EE{$\mathcal E$\ } \def\FF{$\mathcal F$\ and
\EE )\\
For $\mathcal D$\ } \def\EE{$\mathcal E$\ } \def\FF{$\mathcal F$\ , Figure \ref{ng4} shows only one path from the root
without double arrows, given by the cycle {\rm acfd} which
describes an 8-cycle in the complete neighbor graph. The
associated address $1\overline{22211112}$ corresponds to the
center $0.$ As can be directly seen in Figure \ref{ng4}, this
address is also obtained from a path starting in vertex \cc, going
through the cycle {\rm cefg} and the opposite vertices. Thus the
center 0 belongs to the boundary set $B_\cc,$ hence not to the
interior of $T.$
For \EE, two root addresses of the center $0$ are visible in
Figure \ref{ng2}: $121\overline{2}$ from {\rm abs}, and
$122\overline{1221}$ from {\rm abim.} The second address is also
obtained when we start at {\rm -e} and go {\rm -ebcef-e-fef... }
Thus $0\in B_{\rm -e},$ the center is not in ${\rm int}\, T.$
\end{Exa}
The addresses were found from a calculation of $G^2$ although $G$
is sufficient to check them. We determined $G^2$ also for the more
complicated twindragons and found that for \FF there are three
faces $B_k$ which intersect their opposite faces $B_{-k},$ not in
a point, but in a Cantor set with $\dim^w B_k\cap B_{-k}=0.6.$ It
is not clear whether any of the $B_k$ contains $c,$ so the above
argument fails, but it is obvious that $T$ has not the structure
of a polyhedron. In an ordinary polyhedron, given topologically as
a planar map on the sphere, at most one pair of opposite faces can
intersect. For $\mathcal G$\ the situation is similar: we have even five
pairs of opposite faces $B_k$ which intersect each other, not in a
point, but in an uncountable set.\smallskip
{\bf The polyhedral structure of \BB and \CC. } Only the two
twindragons shown in Figures \ref{twin6} and \ref{twin7} can still
have the structure of a polyhedron. We shall prove that this is
not quite the case but we expect that
\begin{Conj} The twindragons \BB and \CC are homeomorphic to a
ball.\end{Conj}
It will not be possible to settle this question here, but we shall
establish the polyhedral structure which leads us to the
conjecture. We start with the list of non-empty intersections of
faces which was established by computing the graph $G^2.$
\smallskip
\begin{Exa} ( Polyhedral structure of \CC)\\
In \CC, the following faces have uncountable intersection:\\
{\rm a \emph{with} b,f,-c,-e \qquad b \emph{with} a,c,d,-e\qquad c
\emph{with} b,d,-a,-f\\ d \emph{with} b,c,e,f\qquad\quad e
\emph{with} d,f,-a,-b \qquad f \emph{with} a,d,e,-c\\}
Moreover, there are one-point intersections\\
{\rm a,-a \emph{with} d,-d, \qquad b,-b \emph{with} f,-f\quad
\emph{ and }\quad c,-c \emph{with} e,-e\\ } and two-point
intersections {\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} with \cc and {\rm -a} with {\rm -c}. The two
points coincide, however, since the corresponding addresses are
identified by Theorem 4.5 (i).
\end{Exa}
\begin{Proof} The proof is a simple but lengthy calculation.
We sketch some facts which can be seen directly from inspection of
$G$ in Figure \ref{ng6}. The address $\overline{2}$ belongs to
$L_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} , L_\bb, L_{\rm d}, L_{\rm f}$ and it is clear that it is
the only address in $L_a\cap L_{\rm d}$ as well as in $L_\bb\cap
L_{\rm f}.$ For the opposite faces we have address $\overline{1}.$
Moreover, $L_\cc\cap L_{\rm e}=\overline{2211}=L_{\rm d}\cap
L_{\rm -a}$ and $L_\cc\cap L_{\rm -e}=\overline{1221}=L_{\rm
f}\cap L_{\rm -b}.$ We can consider the languages restricted to
the vertices of faces, due to Proposition 9.2. Nevertheless, we
see that these addresses also appear for the point neighbors g, h
and -j, respectively, so they should really represent corner
points in the tiling by $T.$
It is obvious from $G$ that $L_{\rm f}\cap L_{\rm -f}=L_{\rm
f}\cap L_{\rm -e}=\emptyset .$ From this we conclude
\[ L_\bb\cap L_{\rm e}=2(L_{\rm d}\cap L_{\rm -c})=22(L_{\rm
f}\cap L_{\rm -f})\cup 22(L_{\rm f}\cap L_{\rm -e})=\emptyset .\]
Now we can determine
\[ L_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} \cap L_\cc= *(L_\bb\cap L_{\rm e})\cup 2(L_\cc\cap L_{\rm
e})\cup 1(L_\bb\cap L_{\rm f})=2\overline{2211} \cup
1\overline{2}\, ,\] and these two addresses are equivalent since
they label opposite paths from the root of $G.$
\end{Proof}
\begin{figure}
\begin{center}
\includegraphics[width = 77mm]{twin6po.eps}\quad
\includegraphics[width = 67mm]{twin7po.eps}
\caption{\label{twinpo} Polyhedral structure of
\CC (left) and \BB (right)}
\end{center}
\end{figure}
\begin{Prop} ( Equivalent convex polyhedron for \CC)\\
The polyhedron on the left of Figure \ref{twinpo} realizes exactly
all intersections of the faces of \CC, except for $B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} \cap
B_\cc$ and $B_{\rm -a}\cap B_{\rm -c}.$ It is an octahedron
truncated at four vertices, in the middle of the corresponding
edges, and its faces are rhombi.\\
Nevertheless, \CC has not the structure of an ordinary polyhedron
because the interior of the face $B_\bb$ is not connected and so
$B_\bb$ is not homeomorphic to a disk.
\end{Prop}
\begin{Proof}
The first part is done by checking all pairs of faces. For the
second part becomes clear when we prove that $B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} \cap B_\cc$
lies on the surface of $T,$ in $B_\bb .$ However, it can be seen
in $G$ that $1\overline{2}$ is an address in $L_\bb .$\\
Here is another argument. From $G$ we have the equation
$B_\bb=f_1(B_{\rm d})\cup f_2(B_{\rm d}).$ On the other hand
$f_1(T)\cap f_2(T)=f_1(B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} )=f_2(B_{\rm -a}).$ We have seen that
$B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} $ and $B_{\rm -a}$ intersect $B_{\rm d}$ in a single point,
and this also holds for their images under $f_1$ or $f_2.$ Thus
$f_1(B_{\rm d})$ and $f_2(B_{\rm d})$ have only one common point.
So $B_\bb$ consists of two isometric pieces which intersect in a
singleton, and cannot be homeomorphic to a disk.
\end{Proof}
It should be mentioned that the polyhedron in Figure \ref{twinpo}
describes \CC only topologically - the metric structure is
completely different. We have $B_{\rm f}=f_2(B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} )$ and $B_{\rm
d}=f_2(B_{\rm f})$ so these faces are of different size. The
intersection of the faces $B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} ,B_\bb, B_{\rm d}, B_{\rm f}$ is
not a corner as in Figure \ref{twinpo}: the point $\pi
(\overline{2})=(\frac12,0,\frac12)'=\frac{{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} +{\rm
d}}{2}=\frac{\bb+{\rm f}}{2}$ is on the middle between the centers
of opposite adjoining faces (see the table at Figure \ref{ng6}). A
similar calculation shows that $B_{\rm a} } \def\bb{{\rm b} } \def\cc{{\rm c} \cap B_\cc$ is in fact the
center point of $B_\bb .$ Now let us study \BB in the same way as
\CC .\smallskip
\begin{Exa} ( Polyhedral structure of \BB)\\
The following faces have uncountable intersection in \BB :\\
{\rm a \emph{with} c,d,e,g,-b,-f \qquad b \emph{with}
c,f,-a,-d\qquad c \emph{with} a,b,e,f,g,-d\\ d \emph{with}
a,e,f,-b,-c,-g \qquad e \emph{with} a,c,d,f \qquad f \emph{with}
b,c,d,e,-a,-g\qquad g \emph{with} a,c,-d,-f.\\}
Moreover, there are one-point intersections\\
$B_\cc\cap B_{\rm -e}=\pi (\overline{12}),$ $B_\cc\cap B_{\rm
-a}=\pi (2\overline{21}),$ and their opposites.
\end{Exa}
\begin{Prop} ( Equivalent convex polyhedron for \BB)\\
The polyhedron on the right of Figure \ref{twinpo} realizes
exactly all intersections of the faces of \BB, except for the
point neighbors. It is an octahedron truncated at all vertices, at
one third of the corresponding
edges. The faces are regular hexagons and squares.\\
Nevertheless, \BB has not the structure of an ordinary polyhedron
because the interior of the faces $B_\bb$ and $B_{\rm d}$ is not
connected and so these faces are not homeomorphic to a disk.
\end{Prop}
This is proved similarly as for \CC . Note that faces of the
truncated octahedron never meet in a single point. Using Figure
\ref{ng7} we see that $B_\cc\cap B_{\rm -a}$ is on $B_\bb$ and
$B_\cc\cap B_{\rm -e}$ on $B_{\rm -d}.$
\section{Connectedness of the interior}
In this section, we go a small step towards proving our
conjecture, by proving that the interior of \CC is connected.
\begin{Prop}
Suppose $T=\bigcup_{j=1}^m f_j(T)$ is a self-affine tile and there
is a connected set $E\subset{\rm int}\, T$ such that $E\cap
f_j(E)\not=\emptyset$ for $j\in I=\{ 1,...,m\}.$ Then the interior
of $T$ is connected.
\end{Prop}
\begin{Proof} $D_1=\bigcup_{j\in I} f_j(E)$ is a connected subset of
${\rm int}\, T.$ By induction we show that $D_n=\bigcup_{|w|\le n}
f_w(E)$ is connected, where $|w|$ denotes the length of the word
$w.$ Then $D=\bigcup_{n=1}^\infty D_n=\bigcup_{w\in I^*} f_w(E)$
is connected, and this set is dense in ${\rm int}\, T.$ Hence ${\rm int}\, T$
is connected.
\end{Proof}
\begin{Prop} ${\rm int}\, $\CC is connected. \end{Prop}
\begin{Proof} A piece $T_w$ belongs to ${\rm int}\, T$ if it is surrounded
in $T$ by all possible neighbors. $T_w$ has all possible neighbors
within $T$ if a path labelled with a suffix of $w$ leads from the
root to each vertex of $G$ (cf. proof of Theorem 5.1). When we
require this for faces only, $w=121212$ and $v=2121121$ fulfil
this condition, as one can check with Figure \ref{ng6}. We put
\[ E=T'_w\cup T'_v\cup T'_{\tilde{w}}\cup T'_{\tilde{v}}\, ,\qquad
\tilde{w}=212121, \tilde{v}=1212212\] where $T'_w$ means $T_w$
minus its one-point boundary sets. Then $E\subset {\rm int}\, T.$ Since
$E$ contains $\pi (\overline{12})$ and $\pi (\overline{21}),$ the
set $E$ intersects $f_1(E)$ and $f_2(E).$ To apply Proposition
13.1, it only remains to verify that $E$ is connected. Since
twindragons are connected (cf. section 7), each $T'_u$ is
connected. Moreover, there are pairs of opposite paths from the
root in $G$ with labels $w1$ and $v,$ hence also with $\tilde{w}2$
and $\tilde{v},$ and also with $v$ and $\tilde{v},$ which all end
in a vertex of $G$ corresponding to a face. Thus $T_{w1}$ and
$T_v,$ $T_{\tilde{v}}$ and $T_{\tilde{w}2},$ as well as $T_v$ and
$T_{\tilde{v}}$ have a face in common. So $E$ is connected, and
the proposition applies.
\end{Proof}
$w$ and $v$ were obtained in a straightforward way. $w$ is the
shortest word for which $T_w$ has all possible face neighbors (we
used a computer search). $v$ is the shortest word with all face
neighbors which contains the center point 0 by Proposition 12.3
(this is easy: the address of 0 is $1\overline{212}$). In fact, 0
is an interior point of $T_v\cup T_{\tilde{v}}.$ For \BB we also
determined such $v.$ The address of 0 is $1\overline{122},$ which
gives $v=1122122122.$ This indicates that the interior of \BB is
more fragmented than that of \CC and, to let the above method
work, we would need a longer chain of pieces.
\section{Remarks on algorithms}\label{algo}
{\bf Use of computer. } We have shown that the topology of
self-affine tiles can be studied by rather simple methods: finite
automata and regular languages. To answer concrete questions,
however, we have to perform quite a number of elementary logical
operations. Computer assistance seems necessary and convenient to
study more complicated examples.
Although all results in this paper have been checked by hand,
interactive computer work was essential to obtain them. Let us
briefly document the main algorithms we have used. \smallskip
{\bf Construction of the neighbor graph. } To construct $G,$ three
steps are performed, as indicated in Example 4.4.
\begin{enumerate}
\item[ (i)] Lower and upper bounds $l_q,b_q$ for $\{ x_q|\, x\in T\}$ are
derived for each coordinate $q=1,2,3.$ They need not be sharp, but
should be taken rather generously (we work with integers anyway).
If $b_q$ is too large, the computer will work a little longer --
if it is too small, only part of the neighbor graph will be
determined. For examples like \BB and \CC, 100000 points of an IFS
algorithm \cite{br} are completely sufficient to find $l_q$ and
$b_q$. For \FF and $\mathcal G$\ there are rare outliers on the thin fibres,
and an exact estimate is needed.
\item[(ii)] Starting with $k=0,$ a list of new vectors is
calculated by the recursion \[ k'=Mk+k_j-k_i \] where $i,j$ take
all values in $\{ 1,..,m\}.$ The index of $k$ is listed together
with the labels $i,j$ at $k'$ as the predecessor of $k',$ and the
number of successors of $k$ is updated at $k.$ Of course, $k'$ is
only processed if it is within the bounds $l_q$ and $b_q.$
Moreover, it must always be checked whether $k'$ has already been
listed before. In that case, the new predecessor is added at the
old place. Since the number of possible $k$ is finite, the whole
list will be completed after finite time, and describes a graph
containing $G.$
\item[(iii)] All $k$ without successors are removed from
the list, and for each of their predecessors, the number of
successors is updated. This step is repeated until each $k$ has at
least one successor (or, in case of a Cantor set, $G$ is empty).
\end{enumerate}
When the resulting $G$ has predecessors of the root, the open set
condition is not fulfilled. This cannot happen if the $k_j$ form a
complete residue system for $M.$ \smallskip
{\bf Other operations with graphs. } For constructing the
intersection graph $G^2$ from $G,$ we have to consider $G\times
G,$ identify $(u,v)$ with $(v,u)$ and define the edges according
to Definition 11.1. Then we apply step (iii) as above, to exclude
empty intersections.
To identify point neighbors in $G,$ we use the adjacency matrix
$H$ defined in section 10. If $q$ is the number of non-zero
vertices in $G,$ then $k$ is a point neighbor if the sum of the
row associated with $k$ in $H,H^2,...,H^q$ always equals 1. When
we repeatedly apply a procedure like (iii) to the point neighbors,
we also remove the finite boundary sets. One-point intersection
sets in $G^2$ are found in the same way as point neighbors in $G.$
The partial order of boundary sets (section 8) is also determined
with $H.$ Let the matrix $N$ be the maximum of $H+H^2+...+H^q$ and
1, taken in each cell. Then $n_{uv}=1$ if and only if $u\succ v.$
In the row of $u$ there are the elements which can be reached from
$u,$ in the column of $u$ there are the $v$ from which $u$ can be
reached. The strong components of $G$ can be found by properly
ordering the rows and columns of $N.$
To deal with languages $L_u,$ we introduce adjacency matrices
$H_i$ for the edges with label $i.$ Then the matrix $H_1H_2H_2,$
for instance, tell us between which vertices there is a path
labelled $122.$ This fact was used to find the proof of
Proposition 9.5.
There are many similar tools waiting to be developed and tested to
help reveal the geometric structure of fractal tiles. \pagebreak
|
\section{Introduction}
River flow arises as a result of interactions between climatic inputs and landscape characteristics over a wide range of spatial and temporal scales. Depending upon the nature of the climatic inputs (e.g. rainfall, temperature) and that of the landscape (e.g. basin area, slope, land use), river flow process can have regular or irregular behaviors. The degree of regularity or irregularity, i.e. \emph{complexity} (loosely speaking), of the river flow process serves as an important evaluator of its predictability. Study on the dynamic nature and predictability of river flow process has always been a key research topic in the field of hydrology and water resources and in related fields, as such play vital roles both in undertaking short-term water emergency measures and in devising long-term water management strategies.
Linearity and nonlinearity are two of the fundamental properties of river flow process that ultimately dictate its level of complexity. Although the inherent nonlinear nature of river flow process has been known for long \cite{Izzard66, Amorocho67, Amorocho71}, much of early hydrologic research (especially during 1960s-1980s) was constrained to linear approaches \cite{Harms67, Yevjevich, Klemes78, Salas81}; the lack of data and computational power largely contributed to this situation. Linear approaches continue to be prevalent in river flow and other hydrologic studies. However, advances in computational power and data collection during the last two decades or so have also facilitated proposal of nonlinear approaches as viable alternatives. The nonlinear approaches that have found widespread applications in river flow studies include artificial neural networks, data-based mechanistic models \cite{Young94}, and chaos theory \cite{Sivakumar00}, among others. The present study is concerned with chaos theory.
In the nonlinear science literature, the term \emph{chaos} is normally used to refer to situations where complex and random-looking behaviors arise from simple nonlinear deterministic systems with sensitive dependence on initial conditions \cite{Lorenz63}, and the converse also applies \cite{Goerner}. The three fundamental properties inherent in this definition, (i) nonlinear interdependence; (ii) hidden determinism and order; and (iii) sensitivity to initial conditions, are highly relevant in river systems and processes. For example: (i) components and mechanisms involved in the river system (e.g. rainfall, flow, sediment transport) act in a nonlinear manner and are also interdependent; (ii) annual cycle in river flow possesses determinism and order; and (iii) particle transport in rivers largely depend upon the time (i.e. rainy or dry season) at which the particles were released at the source, which themselves are often not known. The first property represents the 'general' nature of river flow, whereas the second and third represent its 'deterministic' and 'stochastic' natures, respectively.
In view of the obvious relevance of its fundamental concepts to river systems, chaos theory has been gaining considerable interest in river flow and related studies \cite{Sivakumar05, Koutsoyiannis06} (and Ref. therein). The outcomes of these studies are encouraging, as they reveal that chaos theory offers new avenues to study the inherent nonlinear and complex nature of the river flow process.
The major findings in this direction are related to both chaotic and stochastic nature in the flow dynamics: it ranges from a less complex (deterministic) to a more complex (stochastic) behavior by varying the scale of aggregation. Thus, apparently controversial results may emerge by employing analyses on river flow, because of the observed transitions from determinism to stochasticity with increasing time scale \cite{Sivakumar09}. For an extensive review on chaos theory applications to river flow (and other hydrologic) processes see Ref. \cite{Sivakumar00, Sivakumar04, Sivakumar09} (and references therein).
Hence, another important aspect in regards to the complexity of river flow process is \emph{scale}. Climatic inputs and landscape characteristics often vary in space and time scales and accordingly influence the river processes at various scales. Our general perception is that aggregation in scale averages out the variations and reduces the level of complexity. Such a perception, however, may not always stand good, since averaging may occur only within a limited range of scales, which is often defined by the processes themselves. Further, larger spatial and temporal scales may bring their own complexities, such as additional terrain types in space and climatic scenarios in time. Our knowledge and experience indicate that, for example, daily flow in a small river basin often exhibits a higher level of complexity than that of monthly flow, but the opposite is often the case when the basin size is large. The last few decades have witnessed numerous studies into the scaling properties of river networks and river flows \cite{Sivakumar04, Rinaldo06} (and Ref. therein). It is clear, from the above observations, that reliable assessment of the complexity and predictability of river flow process and identification of the appropriate models for predictions and engineering applications requires careful investigation of both the nonlinear (especially chaotic) and scaling properties. However, a close examination of the literature reveals that studies have, in general, investigated only the chaotic property or the scaling property, but not both.
Linear analysis, by means of autocorrelation and power spectrum, and nonlinear analysis in the framework of multifractal theory have been extensively used as standard methods to investigate only the scaling feature of river flow and rainfall. It has been shown that models known as multiplicative cascades, are able to simulate river flow scaling behaviour \cite{koscielny2006long}. The range and the nature of scaling have been studied on several streamflow data from the real world, identifying basic regimes, scale breaking, universal multifractal parameters and their independence on basin size, reflecting the space-time multiscaling of both the runoff and the rainfall processes \cite{tessier1996multifractal, pandey1998multifractal}. Hence, space-time multifractals appear to be the natural framework for analyzing the scaling features of geodynamical processes including river flows, but it is worth remarking that all of these features have been investigated by using river flow and rainfall data.
In fact, the nature of multiplicative cascades is controversial. Multifractal cascades have been successfully introduced as simple stochastic mechanisms for understanding the self-similar and intermittent behaviour of turbulent processes \cite{parisi1985turbulence, meneveau1987simple}. However, it has been shown that the deterministic variant of multiplicative cascade models can be chaotic, preserving the scaling feature \cite{biferale1994chaotic}. When, as in our case, there is no information about rainfall in the same period of river flow data, a deeper comprehension of the dynamics requires the analysis of chaotic features as well as multifractal ones: this provides the motivation for the present study to investigate both these properties in river flow.
In this paper, first, we review the main methods for investigating the chaotic behavior and the scaling properties from the time series of a process. The investigation on chaos is made by employing a multi-dimensional phase space reconstruction, using the embedding theorem \cite{Packard80, Takens81}, to obtain preliminary information on possible patterns and extent of complexity; the correlation dimension method \cite{GrassbProc83} and the largest Lyapunov exponent method \cite{Rosenstein} are used to investigate the geometry of the phase space. The scaling properties are examined through the Hurst exponent estimation \cite{Hurst51, Hurst56} and the R\'enyi dimension analysis \cite{Renyi61}. Second, as a case study, flow dynamics of the Karoon River in Iran is investigated and daily flow data, observed during 1999-2004, are analyzed by applying these methods.
\section{Methods}
In this study, the investigation of the presence/absence of chaos and scaling behaviors in the river flow series is made using a host of methods. We chose to use more than one method to avoid spurious results related to possible drawbacks of each procedure. Phase space reconstruction, correlation dimension, and Lyapunov exponent methods are employed for detecting chaos, whereas Hurst exponent and R\'enyi dimension analyses are performed to identify scaling characteristics. Brief descriptions of these methods are presented in the following.
\subsection{Phase space reconstruction}
Phase space is a useful tool for representing the evolution of a system in time. Phase space is essentially a graph or a coordinate diagram, whose coordinates represent the variables to completely describe the state of the system at any time \cite{Packard80}. The trajectories in the phase space describe the evolution of the system from some initial state, which is assumed to be known, representing the history of the system. The region of attraction of these trajectories in the phase space provides important qualitative information on the extent of complexity of the system. The idea behind such a reconstruction is that a (nonlinear) dynamic system is characterized by self-interaction, and a time series of a single variable can carry the information about the dynamics of the entire multivariable system (When multiple variables representing the system are available, it is obviously desirable to reconstruct the phase space using such multiple variables). Many methods are available for phase space reconstruction from a scalar time series. Among these, the method of delays \cite{Takens81} is the most widely used one. According to this method, for a scalar time series $\{x_{n}\}$, ($n=1,2,3,...,N$) multi-dimensional phase space can be reconstructed as
\begin{eqnarray}
\mathbf{s}_{n} = \(x_{n}, x_{n-\tau}, ..., x_{n-(m-1)\tau} \)
\end{eqnarray}
where $m$ is the dimension of the phase space, called embedding dimension, and $\tau$ is an appropriate (integer) delay time.
Many different guidelines have been offered in the literature for the selection of $m$ and $\tau$. For instance, according to the embedding theorem of Takens \cite{Takens81}, an $m = 2D_{2} + 1$-dimensional phase space is required to completely characterize a dynamic system with an attractor dimension $D_{2}$, whereas, in practice, $m > D_{2}$ would be sufficient \cite{Abarbanel90}. Similarly, for the selection of $\tau$, some studies suggest the autocorrelation function method \cite{Holzfuss}, while mutual information method \cite{Fraser86} and correlation integral method \cite{Liebert89} are also suggested. As of now, there is no general consensus on the selection of $m$ and $\tau$. In this study, the average mutual information (AMI) method is used to select $\tau$ and the false nearest neighbor (FNN) algorithm is employed to obtain $m$ following Ref. \cite{Cellucci03}, where the best values for the embedding dimension and the lag time have been identified within these approaches among the available ones.
The average mutual information (AMI) approach gets the optimal delay time $\tau$ as the first local minimum \cite{Fraser86} of the information measure
\begin{eqnarray}
I(\tau)=\sum_{ij}p_{ij}(\tau)\log \frac{p_{ij}(\tau)}{p_{i}(\tau)p_{j}(\tau)}
\end{eqnarray}
where $p_{i}$ and $p_{j}$ are the individual probabilities of $x_{n}$ and $x_{n+\tau}$ respectively, and $p_{ij}$ is the joint probability. The AMI quantifies the amount of information about $x_{n+\tau}$ if $x_{n}$ is known: when $x_{n+\tau}$ carries the maximum information about the knowledge of $x_{n}$, AMI is locally minimum. The optimal embedding dimension $m$ can be obtained from the false nearest neighbors (FNN) search in phase space \cite{Kennel92}: the number of false neighbors in the phase space generally changes between two successive embedding dimensions $m_{0}$ and $m_{0} + 1$; the optimal embedding is realized when this number is zero. However, real time series are noisy and, hence, the dimension $m_{0} + 1$ is an optimal embedding when the percentage of false neighbors respect to the embedding $m_{0}$ is less than a certain threshold, generally fixed to be 1\%.
\subsection{Correlation dimension method}
Grassberger and Procaccia \cite{GrassbProc83} introduced a fractal measure, namely the correlation sum, to quantify the amount of correlations in a time series. Their function is defined as the number of pairs of points closer than a given distance $\epsilon$, respect to some norm $|\cdot|$, in the embedding space. An improved definition of the correlation sum, excluding $\tilde{n}$ time-correlated neighbors in the phase space, is given by
\begin{equation}
C(\epsilon,m)=\frac{2}{(N-\tilde{n})(N-\tilde{n}-1)}\sum_{i+\tilde{n}<j}\Theta\(\epsilon-|\mathbf{s}_{i}-\mathbf{s}_{j}|\)
\end{equation}
where $\Theta$ is the Heaviside step function, $\tilde{n}$ is called Theiler window and the sum is extendend to all pairs of points in the embedding space. Correlation sum $C(\epsilon,m)$ behaves as $\epsilon^{m}$ (for any $\epsilon$) for stochastic systems and as $\epsilon^{d}$ (for values of $\epsilon$ in the scaling region) for deterministic ones and some types of coloured noise \cite{GrassbProc83, Provenzale89, Provenzale91}. When a scaling relationship between $C(\epsilon,m)$ and $\epsilon$ exists, the scaling exponent, namely, the \emph{correlation dimension} $D_{2}(\epsilon, m)$, is defined as
\begin{eqnarray}
D_{2}(\epsilon,m)=\frac{\partial \log C(\epsilon,m)}{\partial \log \epsilon}
\end{eqnarray}
For Takens embedding theorem, correlation dimension is an estimation of the degrees of freedom of the underlying process, i.e. it quantifies the number of equations needed to describe the phenomenon if it is deterministic. Numerically, $D_{2}(\epsilon,m)$ can be obtained by plotting $\ln C(\epsilon,m)$ versus $\ln\epsilon$ for different $m$ and $\epsilon$, and by taking the local slopes: for values of $m$ greater than the optimal embedding needed for phase space reconstruction, $D_{2}$ is expected to reach a plateau in the scaling region. The local slopes approach is often subjected to poor estimation of the correlation dimension: if a scaling region exists, the Takens-Theiler estimator, defined by
\begin{eqnarray}
\label{d2takens}D_{2}^{TT}(\epsilon)=\frac{C(\epsilon)}{\int_{0}^{\epsilon}\frac{C(\epsilon')}{\epsilon'}d\epsilon'}
\end{eqnarray}
can be used to improve the accuracy of estimation. Chaotic dynamical systems takes a fractional value for $D_{2}$ while deterministic, but not chaotic, dynamical systems takes an integer value. For a wide range of stochastic dynamics, $D_{2}$ goes to infinity: unfortunately the existence of a plateau and of a finite and fractional correlation dimension is not enough to distinguish chaotic time series from some stochastic ones, as coloured noise \cite{Provenzale89, Provenzale91}.
\subsection{Lyapunov exponent method}\label{Sec-Lyap}
Processes can be characterized by their sensitivity to initial conditions. If $\delta(t)$ is the distance between two nearby orbits in the phase space at some time $t$, the evolution of the \emph{sensitivity} to initial conditions $\xi(t)=\delta(t)/\delta(0)$ shows quite different behaviors for deterministic and stochastic dynamical systems \cite{Gao}:
\begin{eqnarray}
\label{lyapdef}\frac{d\xi_{\text{Det.}}(t)}{dt} &=& \lambda_{max} \xi_{\text{Det.}}(t)\\
\label{hurstdef}\frac{d\xi_{\text{Stoc.}}(t)}{dt} &=& t^{H}
\end{eqnarray}
where $\lambda_{max}$ is called \emph{largest Lyapunov exponent} and $H$ is called \emph{Hurst exponent}. Recent unifying approaches, strictly connected to nonextensive thermodynamics \cite{Tsallis97}, are still under investigations.
Largest Lyapunov exponent is a robust measure of the mean convergence (or divergence) rate of $\xi(t)$: for processes that exhibit chaotic behavior, nearby trajectories diverge along time and $\lambda_{max}>0$, while $\lambda_{max}\leq 0$ for deterministic, but not chaotic, dynamical systems. Thus, a positive largest Lyapunov exponent is a strong signature of chaos.
A positive $\lambda_{max}$ defines a prediction horizon $t^{*}(\alpha)=N^{*}\Delta t$, i.e. the maximum number of samples $N^{*}$ that can be predicted at a sampling time $\Delta t$ within an uncertainty $\alpha$. Real time series are affected by a measurement error $\varepsilon$: it follows that $\delta(0)=\varepsilon$. The worst uncertainty on the prediction of the series is $\alpha=\text{max}\{x_{n}\}-\text{min}\{x_{n}\}$, i.e. it equals the whole range of the series. A desirable uncertainty should be $\alpha=1.96\varepsilon$, corresponding to $95\%$ of confidence band. From eq. (\ref{lyapdef}) it follows
\begin{eqnarray}
t^{*}(\alpha) = \frac{1}{\lambda_{max}}\log\frac{\alpha}{\varepsilon}
\end{eqnarray}
and $t^{*}(1.96\varepsilon)=\lambda_{max}^{-1}\log 1.96$. The largest Lyapunov exponent can be estimated from an observed times series with different approaches \cite{Rosenstein,Sano85,Kantz94}: in the following we will adopt Rosenstein's one \cite{Rosenstein}. In the embedded space, the distance of a reference point $\mathbf{s}_{n_{0}}$ from all the other points $\mathbf{s}_{n'}$ in its neighborhood of size $\epsilon$, is calculated and averaged over the number of neighbors $|\mathcal{U}_{\mathbf{s}_{n_{0}}}|$. This procedure, iterated for each point along an orbit of $\mathcal{N}$ samples, defines the stretching factor $S(\epsilon,m,t)$ that depends on the mean convergence (or divergence) rate of nearby trajectories:
\begin{eqnarray}
\label{stretchfactor}S(\epsilon,m,t)=\frac{\Delta t}{t}\sum_{n_{0}=1}^{t/\Delta t} \log\[ \frac{1}{|\mathcal{U}_{\mathbf{s}_{n_{0}}}|} \sum_{n'} |\mathbf{s}_{n'}-\mathbf{s}_{n_{0}}| \]
\end{eqnarray}
where $t=\mathcal{N}\Delta t$. The largest Lyapunov exponent is the slope of the linear region obtained by plotting $S(\epsilon,m,t)$ versus $t$, by keeping fixed $\epsilon$ and $m$. The above procedure avoids the estimation of the tangent map, it is fast and easy to implement and it is suitable for small and noisy data sets \cite{Rosenstein}.
However, the estimation of Lyapunov exponents may produce spurious results if long time series of high quality are not available \cite{eckmann1992fundamental} or in presence of a strong stochastic contaminating signal \cite{tanaka1998analysis}.
The estimation of the prediction horizon, as previously introduced, can be verified by making use of a popular forecasting technique based on nonlinear prediction \cite{Farmer87,Sugihara90}. The procedure is as follows. In the reconstructed phase space, find all the embedding states $\mathbf{s}_{n_{0}}$ in the neighborhood $\mathcal{U}_{\epsilon}(\mathbf{s}_{N})$ of size $\epsilon$ of the current state $\mathbf{s}_{N}$. The future states $\mathbf{s}_{n_{0}+k}$, $k$ steps ahead, of all states $\mathbf{s}_{n_{0}}\in \mathcal{U}_{\epsilon}(\mathbf{s}_{N})$ are successively used for the prediction of the measurement at time $N+k$:
\begin{eqnarray}
\hat{s}_{N+k} = \frac{1}{|\mathcal{U}_{\epsilon}(\mathbf{s}_{N})|}\sum_{\mathbf{s}_{n_{0}}\in \mathcal{U}_{\epsilon}(\mathbf{s}_{N})}s_{n_{0}+k}
\end{eqnarray}
i.e. the forecasting is obtained by averaging over all closest embedding states \cite{Kantz}. In the presence of an underlying chaotic dynamics, the forecasting error is expected to exponentially increase with the forecasting time at a rate $\lambda_{max}$, corresponding to the largest Lyapunov exponent.
The forecast error, divided by the standard deviation of the time series, is maximum when the forecast time equals the prediction horizon.
\subsection{Hurst exponent method}
Hurst exponent, previously defined in eq. (\ref{hurstdef}), has been originally introduced for investigating the diffusion features of the Nile river \cite{Hurst51, Hurst56}, and it is widely used to detect the scaling regions and to characterize the persistence of a process. Let $\{y_{n}\}$ be the partial sum time series of the series $\{x_{n}\}$, defined as
\begin{eqnarray}
y_{n}=\sum_{i=1}^{n}\[x_{i}-\langle x\rangle\], \quad\quad \langle x \rangle = \frac{1}{N}\sum_{j=1}^{N}x_{j}\nonumber
\end{eqnarray}
If $\nu$ is an integer delay time, the Hurst exponent $H(q)$ is obtained from the structure function
\begin{eqnarray}
\label{structfunc}S_{H}(q,\nu) &=& \frac{1}{t/\Delta t-\nu}\sum_{n_{0}=1}^{t/\Delta t-\nu} |y_{n_{0}+\nu}-y_{n_{0}}|^{q}
\end{eqnarray}
when $S_{H}(q,\nu)\sim \nu^{qH(q)}$: the power-law is typical of a fractal process at the time scale $\nu\Delta t$. If $H(q)$ is not a constant function of $q$, the process is said to be \emph{multifractal} \cite{Gao}. Hurst exponent is a bounded measure ($0\leq H(q) \leq 1$) characterizing the persistence of a process:
\begin{itemize}
\item $H(q) \approx 0.5$ indicates a memoryless time series, with neither short-tem nor long-term correlation between states, typical of uncorrelated stochastic processes as white noise;
\item $0 \leq H(q)< 0.5$ indicates anti-persistence: increasing trends will be followed by decreasing ones, or viceversa, and this behavior tends to be dominant for $H\to 0$;
\item $0.5<H(q)\leq 1$ indicates persistence: there is only one persistent trend typical of processes where diffusion is faster than simple brownian motion.
\end{itemize}
For a review about Hurst analysis application to hydrological sciences we refer to \cite{Koutsoyiannis02}. However, the above method could not be robust if applied to nonstationary signals showing evident linear or seasonal trends. Instead, methods based on detrended fluctuation analysis (DFA) have been developed to remove linear (or higher order) trends, that may exists in nonstationary signals, before performing the scaling analysis. We refer to Ref. \cite{chen2002effect} for a study of the impact of nonstationary contaminating signals on the scaling features of the original data and to Ref. \cite{koscielny2006long} for a multifractal study of river flow data by means of DFA. Of course, in the case of stationary signals, Hurst analysis and DFA agree on the scaling parameters.
\subsection{Generalized $q-$th order entropy}
Given a scalar time series $\{x_{n}\}$ defined on some set $\mathcal{D}$, let us cover this set with a partition $\mathcal{P}_{\epsilon}$ of disjoint boxes of size $\epsilon$. Let $p_{\epsilon}(x)$ be the probability distribution function of the series: $p_{\epsilon}(x_{i})$ is the probability that the series takes the value $x_{i}$ for the partition $\mathcal{P}_{\epsilon}$.
A measure of the average information, needed to specify a point with accuracy $\epsilon$, is the Shannon entropy \cite{Shannon48}
\begin{eqnarray}
\mathcal{H}_{1}(\mathcal{P}_{\epsilon})=-\left\langle\log p_{\epsilon}(x)\right\rangle = -\sum_{i}p_{\epsilon}(x_{i})\log p_{\epsilon}(x_{i})\nonumber
\end{eqnarray}
When a scaling relationship exists between the amount of information and the accuracy $\epsilon$, the scaling exponent, called \emph{information dimension}, is defined as
\begin{eqnarray}
D_{1} = \lim_{\epsilon\to 0}\frac{\mathcal{H}_{1}(\mathcal{P}_{\epsilon})}{\log \frac{1}{\epsilon}}\nonumber
\end{eqnarray}
In a similar way, the concept of entropy and dimension can be respectively generalized to the $q-$order R\'enyi entropy \cite{Renyi61}
\begin{eqnarray}
\mathcal{H}_{q}(\mathcal{P}_{\epsilon})=\frac{1}{1-q}\log\sum_{i} \[p_{\epsilon}(x_{i})\]^{q}
\end{eqnarray}
and the $q-$th order R\'enyi dimension
\begin{eqnarray}
D_{q} = \lim_{\epsilon\to 0}\frac{\mathcal{H}_{q}(\mathcal{P}_{\epsilon})}{\log \frac{1}{\epsilon}}
\end{eqnarray}
where $q\in[-\infty,\infty]$ and $H_{1}$ and $D_{1}$ are obtained if $q\to 1$. The behavior of $D_{q}$ versus $q$ defines the \emph{fractal spectrum}: the underlying process of the time series $\{x_{n}\}$ is fractal when $D_{q}$ is a constant function of $q$, otherwise it is said to be multifractal. $D_{q}$ can be numerically estimated by plotting $\mathcal{H}_{q}(\mathcal{P}_{\epsilon})$ versus $\log\frac{1}{\epsilon}$ and by taking the slope of the linear region. This type of analysis requires long time series in order to avoid spurious results.
\section{Data analysis and results}
The Karoon river, with a watershed area of $58,180 \text{ km}^{2}$, is located in southwest of the I.R. of Iran, the Khuzestan province is chosen for this study. The river lies between the city of Ahwaz ($31^{\circ} 20^{\prime} \text{ N}, 48^{\circ} 41^{\prime}\text{ E}$) and the Bahmanshir river ($30^{\circ} 25^{\prime} \text{ N}, 48^{\circ} 12^{\prime} \text{ E}$), which is about 190 km long. The Karoon river is a meandering river which supplies water for the irrigation of sugarcane cultivation projects, as well as other agricultural lands.
River flow data, observed over a period of 6 years (from January 1999 to December 2004), are considered. Fig. \ref{ts} shows the variations of daily river flow time series for a sampling time of 1 day.
\begin{figure}[!htb]
\begin{center}
\includegraphics[angle=0, scale=0.3]{karoon.eps}
\caption{{Daily Karoon River flow from January 1999 to December 2004.}}
\label{ts}
\end{center}
\end{figure}
Optimal parameters for the phase space reconstruction, namely delay time $\tau=115$ days and embedding dimension $m=9$, are obtained from the first local minimum of AMI and FNN search, respectively. Correlation dimension $D_{2}(m,\epsilon)$ is estimated from the Takens-Theiler measure (\ref{d2takens}) up to an embedding dimension $m=18$: Fig. \ref{d2} shows a plateau for $D_{2} = 2.60 \pm 0.07$, a necessary, but no sufficient, condition for the evidence of low-dimensional chaotic dynamics, as previously discussed.
\begin{figure}[!htb]
\begin{center}
\includegraphics[angle=0, scale=0.3]{d2.eps}
\caption{{Takens-Theiler estimator for the correlation dimension $D_{2}(m,\epsilon)$ and plateau fit (inner panel) corresponding to $D_{2}=2.60\pm0.07$ for $m\geq 7$ (dashed line).}}
\label{d2}
\end{center}
\begin{center}
\includegraphics[angle=0, scale=0.3]{lyapunov.eps}
\caption{{Stretching factor and largest Lyapunov exponent, $\lambda_{max}=(0.014 \pm 0.001) \text{ day}^{-1}$, corresponding to the slope of the dashed line. \emph{Inner panel.} Normalized forecast error versus the forecast time: the maximum error is obtained for $t= 46$ days; the prediction horizon obtained from the Lyapunov analysis corresponds to $48.1\pm 3.4$ days.}}
\label{lyap}
\end{center}
\end{figure}
The stretching factor (\ref{stretchfactor}), needed for the estimation of the largest Lyapunov exponent, is shown in Fig. \ref{lyap}. We found a positive exponent, namely $\lambda_{max}=(0.014 \pm 0.001) \text{ day}^{-1}$, and a prediction horizon of $48.1\pm 3.4$ days for $\alpha=1.96\varepsilon$.
We verified the estimation of the prediction horizon by making use of the technique based on forecasting, previously described. In Fig. \ref{lyap} (inner panel) is shown the normalized forecast error versus the forecast time: the maximum error is obtained for $t= 46$ days, corresponding to a prediction horizon in excellent agreement with that one estimated from Lyapunov analysis. It is worth remarking that our estimation of both largest Lyapunov exponent and prediction horizon, characterize global features of the underlying dynamics, because they are evaluated by averages on hundreds of reference embedding states.
\begin{figure}[!htb]
\begin{center}
\includegraphics[angle=0, scale=0.3]{hurst.eps}
\includegraphics[angle=0, scale=0.3]{hurst2.eps}
\caption{{Structure function $S_{H}(q,t)$ for $0.5 \leq q \leq 4.5$ (upper panel) and Hurst exponent $H(q)$ (lower panel) at different time scales. In the lower panel, the solid line corresponds to the Hurst exponent expected for uncorrelated stochastic processes.}}
\label{hurst}
\end{center}
\end{figure}
The structure function (\ref{structfunc}) is estimated for several values of $q$ at some different time scales, as shown in Fig. \ref{hurst} (upper panel): scaling, and thus multifractal behavior, emerges at those scales, according to a recent analysis with a multifractal detrended fluctuation method on different river flow data \cite{Movahed08}. Hurst exponents, for each $q$, are estimated from the slopes of the straight lines in Fig. \ref{hurst} (upper panel): a significant slope change defines a time scale braking. In Fig. \ref{hurst} (lower panel) Hurst exponents $H(q)$ are shown versus $q$. We identified three significant time scales from slope variations:
\begin{enumerate}
\item From 1 to 28 days;
\item From 29 to 60 days;
\item From 61 to 114 days.
\end{enumerate}
At time scales 1) and 2), Hurst exponents depend on $q$, as for multifractal dynamical systems, $H(q)>0.5$ for all $q\in[0.5,4.5]$ and thus the time series shows a persistent behavior typical of processes exhibiting long-range dependence. At time scale 3) multifractality is stronger than previous time scales, and a transition emerges from persistent to anti-persistent behavior through a memoryless state around $q=3$, as shown in Fig. \ref{hurst} (lower panel).
Finally, the $q-$th order R\'enyi dimension $D_{q}$ is estimated for several values of $q\in[0.5,4.5]$, as shown in Fig. \ref{Renyi}: multifractal behavior emerges again, according to the Hurst exponent analysis, from an information theoretic point of view.
\begin{figure}[!t]
\begin{center}
\includegraphics[angle=0, scale=0.3]{renyi.eps}
\caption{{The $q-$th order R\'enyi dimension $D_{q}$ estimated from generalized R\'enyi entropy (inner panel) for $0.5 \leq q \leq 4.5$.}}
\label{Renyi}
\end{center}
\end{figure}
Recent studies \cite{Provenzale00, Toniolo02, Ferraris02} reveal that complex behaviors as multifractality, self-organized criticality and on-off intermittency can emerge from nonlinearly-filtered linear autoregressive processes (NFLAP). For a correct interpretations of our results, we follow the procedure suggested in Ref. \cite{Provenzale00, Toniolo02, Ferraris02}. First, we generate 1000 surrogates with the same length of the Karoon river flow time series; second, for each surrogate time series, we estimate the correlation dimension, the largest Lyapunov exponent, the R\'enyi dimensions and the Hurst exponents. Two types of surrogates, for testing two different null hypothesis, are obtained as follows:
\begin{itemize}
\item First type: a surrogate is generated by shuffling the data through a random permutation of $\{x_{n}\}$. The shuffled time series has the same probability distribution function of the original one, but time correlations are completely destroyed;
\item Second type: a surrogate is generated with the same probability distibution function and (almost) the same power spectrum of $\{x_{n}\}$. The amplitude-adjusted Fourier transform (AAFT) algorithm \cite{Theiler92,Schr-Schm96} is used to obtain this surrogate.
\end{itemize}
First type surrogates are used to test the data against the null hypothesis of an underlying uncorrelated stochastic process; second type surrogates are used to test the data against the null hypothesis of an underlying linear process distorted by a nonlinear filter.
We found no surrogates, of first or second type, showing, at the same time, a low fractional correlation dimension, a positive largest Lyapunov exponent and multifractal behavior from both R\'enyi and Hurst analyses. Thus, we can reject both the null hypotheses of an underlying uncorrelated stochastic process or an underlying linear process distorted by a nonlinear filter.
\section{Discussion}
In the last years a strong interest emerged for the underlying dynamics of river flow. Many studies reveal that both chaotic and stochastic behavior may emerge at different spatial and temporal scales. Indeed, multifractality appears to be an important feature of this process. However, each study focused or on chaotic features either on multifractal ones, but not both at the same time. In fact, deterministic and stochastic multiplicative cascades, mainly adopted as models to explain data, show similar multifractal signatures and it is not possible to deduce the real nature of the river flow without making use of both chaos and scaling analyses.
Within the present work, we investigated the salient characteristics of dynamics of the Karoon river (Iran), by examining the daily discharge time series over a period of six years (1999-2004). We followed a nonlinear approach to detect the chaotic and the scaling characteristics of the flow dynamics: the presence of chaos has been analyzed through the correlation dimension and largest Lyapunov exponent methods, while the scaling features have been explored through the Hurst and R\'enyi analyses.
Both fractional correlation dimension ($2.60 \pm 0.07$) and positive largest Lyapunov exponent ($0.014 \pm 0.001$) suggest the presence of low-dimensional chaos: flow dynamics are dominantly governed by three degrees of freedom and can be reliably predicted up to 48 days. The estimation of the prediction horizon is in excellent agreement with the value obtained from the nonlinear forecasting analysis: the forecast error increases with the forecast time and it reaches its maximum for a time scale of 46 days. According to recent studies, our results reveal the presence of scaling typical of (chaotic) deterministic dynamical systems, although the apparently irregular behavior of the data. The fractal structure of a strange attractor emerges in the phase space and, of consequence, the underlying dynamics of the Karoon river can be successfully modelled by few deterministic equations. However, in the absence of a realistic model of the river flow, future discharge can be only monthly predicted from past and current measurements. Because of our lack of information on rainfall and atmospherical data in the same region, we can not directly quantify their effect on the prediction horizon, although our results are in good agreement with recent studies employing different forecasting techniques \cite{porporato1997nonlinear, jayawardena2000noise, lisi2001chaotic, IslamMN, sivakumar2002river}.
Results from the Hurst exponent analysis avoid a memoryless phenomenon and reveal, at different time scales, persistent or anti-persistent behavior of the river flow. Indeed, nor a unique Hurst exponent neither a unique R\'enyi dimension can be attributed to the entire process, suggesting that river discharge is characterized by anomalous scaling typical of multifractal dynamics. In particular, Hurst analysis puts in evidence three time scaling regimes. The first and the second ones, corresponding to monthly and bimonthly scales, show a persistent behavior: long-range dependence dominates the underlying dynamics and diffusion is faster than a simple brownian motion. The third scale, ranging from 60 to about 115 days, is the most extended one: for $q<3$ the process is still persistent, whereas the transition from persistent to anti-persistent behavior is evident for $q>3$. Anti-persistence is symptomatic of a diffusion process slower than a standard brownian motion: in the case of Karoon river flow, it emerges in last scaling regime depending on the way we look at the data by means of $q$. Interestingly, 115 days corresponds to the same time that minimizes the average mutual information, quantifying the maximum temporal delay, between different measurements, before both can be considered no more correlated from an information theoretic point of view. Unfortunately, because of our lack of further data, we are not able to directly relate these results to rainfall or seasonal effects.
The dependence on $q$ of the R\'enyi generalized dimensions is another important signature of multifractal behavior, although it does not produce useful information on the real nature of the underlying dynamics. Our findings from scaling analyses agree, in general, with recent results on the investigation of the multifractal nature of the river flow.
Finally, we performed the same analyses on two types of surrogate time series, to test the null hypotheses that Karoon river flow is an uncorrelated stochastic process or a linear stochastic process distorted by a nonlinear filter. We found no surrogates showing, at the same time, similar chaotic and scaling characteristics of the flow dynamics and we thus rejected both the null hypotheses.
\newpage
\addcontentsline{toc}{section}{References}
\begin{small}
\bibliographystyle{mprsty}
|
\section{Introduction}
\begin{figure}[tb]
\setlength{\unitlength}{0.9mm}
\begin{picture}(450,20)(-15,20)
\put(25.2,1.6){\makebox(20,20)[t]{$\mathbb S(k)$}}
\put(42,11){\makebox(18,22)[t]{$E_1$}}
\put(33,17){\makebox(20,20)[t]{$E_2$}}
\put(9,3.5){\makebox(20,21)[t]{$E_i$}}
\put(34.5,-12){\makebox(20,20)[t]{$E_n$}}
\thicklines
\put(35.2,20){\circle{10}}
\put(38.6,23.6){\line(1,1){14}}
\put(30.5,20){\line(-1,0){19}}
\put(38.6,16.6){\line(1,-1){14}}
\put(36,24.6){\line(1,3){6}}
\put(20,3){\makebox(20,20)[t]{$.$}}
\put(20.9,5){\makebox(20,20)[t]{$.$}}
\put(23.8,6.6){\makebox(20,20)[t]{$.$}}
\put(20,-4){\makebox(20,20)[t]{$.$}}
\put(21.9,-6){\makebox(20,20)[t]{$.$}}
\put(24.8,-7.3){\makebox(20,20)[t]{$.$}}
\put(46,31){\vector(1,1){0}}
\put(46,9){\vector(1,-1){0}}
\put(39.62,35){\vector(1,3){0}}
\put(20,20){\vector(-1,0){0}}
\end{picture}
\vskip 2truecm
\caption{A star graph $\Gamma$ with scattering matrix $\mathbb S(k)$ at the vertex.}
\label{sgraph}
\end{figure}
The past two decades have shown a constantly growing
interest\cite{kf-92,nfll-99,SS,sdm-01,mw-02,y-02,lrs-02,cte-02,ppil-03,coa-03,
dgst-03,rs-04,kd-05,klvf-05,gs-05,
emabms-05,ff-05,drs-06,Bellazzini:2006jb,Bellazzini:2006kh,Bellazzini:2008mn,
hc-08,drs-08,dr-08,hkc-08,adrs-08,Bellazzini:2008cs,dr-09,Bellazzini:2008fu,Ines,Bellazzini:2009nk,rhca-10} in
the physics of quantum wire junctions.
The quantum nature of the transport properties of these devises is fairly well
described by the Tomonaga-Luttinger model\cite{h-81} on graphs
of the type shown in FIG. \ref{sgraph}.
The edges $\{E_i\, :\, i=1,...,n\}$ of such a {\it star} graph $\Gamma$ are
modeling the wires, whereas the vertex represents
the junction, which can be treated\cite{Bellazzini:2006jb} as a point-like defect (impurity) in the Luttinger liquid.
The defect is implemented by a nontrivial one-body scattering matrix $\mathbb S(k)$, $k$ being the momentum.
At criticality the boundary conditions at the vertex are scale invariant and the scattering matrix
is a constant $n\times n$ matrix $\mathbb S$. Various authors\cite{coa-03,Bellazzini:2006kh,drs-06,hc-08,rhca-10}
have investigated in this regime the physical properties of the system. In particular, they
derived the conductance tensor $\mathbb G$, getting the following simple expression
\begin{equation}
\mathbb G_{ij} = \mathbb G_{\rm line} (\delta_{ij} - \mathbb S_{ij}) \, ,
\label{C1}
\end{equation}
the coefficient $\mathbb G_{\rm line}$ being the conductance of a single wire without junctions, which depends
on the parameters (see eq. ({\ref{Gline}) below) of the Luttinger liquid.
In the present paper we pursue further the study of off-critical Luttinger junctions started in
Ref. \onlinecite{Bellazzini:2008fu}. Our main goal here is to
extend the analysis to the case when $\mathbb S(k)$ admits
bound states and to derive the off-critical generalization (see eq. (\ref{C2}) below)
of the formula (\ref{C1}). This is not a purely
theoretical problem, because realistic quantum wire junctions are usually noncritical. In
order to reproduce this situation as close as possible, in what follows we will keep the
theory critical in the {\it bulk} of $\Gamma$, allowing for breaking of scale
invariance only at the vertex. In this regime the scattering matrix is no longer
constant, but depends on the momentum $k$. The continuation of $\mathbb S(k)$ in the complex
$k$-plane is a meromorphic function with poles of the type
$k={\rm i} \eta$ with real $\eta \not= 0$. It turns out that the behavior
of the theory depends in a crucial way on the sign of $\eta$.
In absence of bound states (all $\eta <0$), the model has\cite{Bellazzini:2008fu}
a unitary time evolution respecting time-translation invariance.
Accordingly, the energy is conserved.
The situation radically changes when $\mathbb S(k)$ admits bound states (some $\eta >0$).
Each of them generates in the spectrum of the theory a kind of damped harmonic oscillator.
These oscillators lead to a breakdown of time-translation invariance. The energy of the system is no
longer conserved, which signals a nontrivial energy flow through the boundary
(the vertex of $\Gamma$). The essential point here is that the relative non-equilibrium
state is fixed by the fundamental physical principle of causality (local commutativity).
Physically, the off-critical boundary conditions with $\eta >0$
generate a specific boundary interaction with the environment. One possibility to implement
such interaction in realistic quantum wire junctions might be\cite{ppil-03,coa-03,gs-05}
an external magnetic field crossing the junction.
The paper is organized as follows. In the next section we briefly
review the Tomonaga-Luttinger (TL) model
on a star graph with off-critical boundary conditions at the vertex.
The associated scattering matrix $\mathbb S(k)$ and its analytic properties
are described in detail. The symmetry content of the model is also
analyzed here. In section III we examine the impact of the analytic structure of $\mathbb S(k)$
on the physical properties of the theory. The generalization of the conductance formula
(\ref{C1}) away from criticality is derived as well. The basic features of the
off-critical conductance are discussed in detail, confirming the different role played by
the bound and the antibound states of $\mathbb S(k)$. Section IV is devoted to our conclusions.
The Appendix \ref{appA} contains some technical details concerning the quantization on
$\Gamma$ with off-critical boundary conditions.
\section{Luttinger liquid with off-critical boundary conditions on $\Gamma$}
\subsection{Bulk theory and boundary conditions}
The dynamics of the Luttinger liquid in the bulk is defined by Lagrangian density
\begin{multline}
{\cal L} = {\rm i} \psi_1^*(\partial _t - v_F\partial _x)\psi_1 + {\rm i} \psi_2^*(\partial _t + v_F\partial _x)\psi_2\\
-g_+(\psi_1^* \psi_1+\psi_2^* \psi_2)^2 - g_-(\psi_1^* \psi_1-\psi_2^* \psi_2)^2 \, .\qquad
\label{lagr}
\end{multline}
Here $\{\psi_\alpha (t,x,i)\,:\, \alpha =1,2\}$ are complex fields, depending on the time $t$
and the position $(x,i)$, where $x > 0$ is the distance from the vertex
and $i=1,...,n$ labels the edge, as shown in FIG. \ref{sgraph}. Finally, $v_F$ is the Fermi velocity and
$g_\pm \in \mbox{${\mathbb R}$}$ are the coupling constants\cite{f1}. The equations of motion following from
(\ref{lagr}) imply the conservation law
\begin{equation}
\partial _t \rho (t,x,i) -v_F\partial _x j (t,x,i)= 0\, ,
\label{conservation}
\end{equation}
where
\begin{eqnarray}
\rho (t,x,i) &=& \left (\psi_1^*\psi_1 + \psi_2^*\psi_2 \right )(t,x,i)\, , \\
\label{density}
j (t,x,i) &=& \left (\psi_1^*\psi_1 - \psi_2^*\psi_2 \right )(t,x,i)\, ,
\label{current}
\end{eqnarray}
are the charge density and electric current respectively. Our main task below
is to derive the relative conductance. For this purpose we have to
complete first the description of the dynamics, specifying the interaction
at the vertex of $\Gamma$. In other words, we must fix the boundary conditions
on $\psi_\alpha$ at $x=0$. This is a very delicate point because it
strongly interferes with the solution of the model. It is useful to recall in this respect
that the model (\ref{lagr}) is exactly solvable on the line\cite{h-81} via
bosonization. In order to preserve this nice feature on
$\Gamma$, it is more convenient to formulate the boundary conditions
directly in bosonic terms. We will show now that this strategy, which
works\cite{Bellazzini:2006kh,Bellazzini:2008mn,Bellazzini:2008fu} nicely
at criticality, can be extended to off-critical boundary conditions as well.
The basic ingredient for solving the model (\ref{lagr}) via bosonization is the
scalar field $\varphi$ satisfying
\begin{equation}
\left ({\partial}_t^2 - {\partial}_x^2 \right )\varphi (t,x,i) = 0\, , \qquad x>0
\label{eqm1}
\end{equation}
and some boundary conditions for $x=0$. Following a standard QFT procedure,
the latter are fixed by requiring that the operator $K\equiv -{\partial}_x^2$ on $\Gamma$ is
{\it self-adjoint}. This problem has been
intensively investigated in the recent mathematical literature\cite{ks-00,qg,H1},
where the subject goes under the name of ``quantum graphs". From
these studies one infers that $K$ is self-adjoint on $\Gamma$
if and only if the field $\varphi$ satisfies the
boundary condition\cite{ks-00,H1}
\begin{equation}
\sum_{j=1}^n \left [\lambda (\mbox{${\mathbb I}$}-\mathbb U)_{ij}\, \varphi (t,0,j) -{\rm i} (\mbox{${\mathbb I}$}+\mathbb U)_{ij}
({\partial}_x\varphi ) (t,0,j)\right ] = 0\, ,
\label{bc}
\end{equation}
where $U$ is any unitary matrix and $\lambda > 0$ is a
parameter with dimension of mass, characterizing the breaking of scale invariance.
Eq. (\ref{bc}) generalizes to the graph $\Gamma$ the familiar
mixed (Robin) boundary condition on the half-line $\mbox{${\mathbb R}$}_+$.
The matrices $\mathbb U=\mbox{${\mathbb I}$}$ and $\mathbb U=-\mbox{${\mathbb I}$}$ define the Neumann and Dirichlet
boundary conditions respectively. The physical interpretation of
(\ref{bc}) in the context of bosonization was discussed in Ref. \onlinecite{Bellazzini:2008mn}.
At criticality (\ref{bc}) describes the splitting of the electric current (\ref{current})
at the junction.\cite{f2}
\subsection{Scattering matrix and solution of the model}
As already mentioned in the introduction, for the explicit construction
of $\varphi$ it is convenient to interpret \cite{Bellazzini:2006jb} the vertex of $\Gamma$ as a
point-like impurity (defect) \cite{Liguori:1996xr,Mintchev:2002zd,Mintchev:2003ue,Mintchev:2004jy},
characterized by a non-trivial scattering matrix $\mathbb S(k)$. This matrix
is associated \cite{ks-00,qg} to the operator $K$ on $\Gamma$
and is fully determined by the boundary conditions (\ref{bc}). The
explicit form is \cite{ks-00}
\begin{equation}
\mathbb S (k) = -\frac{[\lambda (\mbox{${\mathbb I}$} - \mathbb U) - k(\mbox{${\mathbb I}$}+\mathbb U )]}{[\lambda (\mbox{${\mathbb I}$} - \mathbb U) + k(\mbox{${\mathbb I}$}+\mathbb U )]}
\label{S1}
\end{equation}
and has a simple physical interpretation: the diagonal element $\mathbb S_{ii}(k)$
represents the reflection amplitude from the defect on the edge $E_i$, whereas
$\mathbb S_{ij}(k)$ with $i\not=j$ equals the transmission amplitude from $E_i$ to $E_j$.
The matrix (\ref{S1}) has a number of remarkable features. It is
unitary $\mathbb S(k)^*=\mathbb S(k)^{-1}$ by construction and satisfies
$\mathbb S(k)^*=\mathbb S(-k)$, known as Hermitian analyticity. Notice also that
$\mathbb S(\lambda ) = \mathbb U$, showing that the boundary condition (\ref{bc}) is fixed actually
by the value of scattering matrix at the scale $\lambda$.
Let us summarize now the analytic properties of $\mathbb S(k)$ needed in what follows.
We denote by ${\cal U}$ the unitary matrix diagonalizing $\mathbb U$ and parametrize
\begin{equation}
\mathbb U_d={\cal U}^{-1}\, \mathbb U\, {\cal U}
\label{d1}
\end{equation}
as follows
\begin{equation}
\mathbb U_d = {\rm diag} \left ({\rm e}^{2{\rm i} \alpha_1}, {\rm e}^{2{\rm i} \alpha_2}, ... , {\rm e}^{2{\rm i}
\alpha_n}\right )\, , \qquad \alpha_i \in \mbox{${\mathbb R}$}\, .
\label{d2}
\end{equation}
Using (\ref{S1}), one easily verifies that ${\cal U}$
diagonalizes also $\mathbb S(k)$ {\it for any} $k$ and that
\begin{multline}
\mathbb S_d(k) = {\cal U}^{-1} \mathbb S(k) {\cal U} = \\
{\rm diag} \left (\frac{k+{\rm i} \eta_1}{k-{\rm i} \eta_1}, \frac{k+{\rm i} \eta_2}{k-{\rm i} \eta_2}, ... , \frac{k+{\rm i} \eta_n}{k-{\rm i} \eta_n} \right ) \, ,
\label{d3}
\end{multline}
where
\begin{equation}
\eta_i = \lambda \tan (\alpha_i)\, ,
\qquad -\frac{\pi}{2} \leq \alpha_i \leq \frac{\pi}{2}\, .
\label{p1}
\end{equation}
We conclude therefore that $\mathbb S(k)$ is a meromorphic function with {\it simple} poles
located on the imaginary axis and different from 0. In what follows we denote by
${\cal P} = \{{\rm i} \eta \, :\, \eta\not= 0\}$ the set of {\it distinct} poles of $\mathbb S(k)$:
the subset ${\cal P}_+ = \{{\rm i} \eta \, :\, \eta>0\}$ in the upper half-plane
corresponds to {\it bound} states whereas ${\cal P}_- = \{{\rm i} \eta \, :\, \eta<0\}$
in the lower half-plane gives raise to {\it antibound} states.\cite{f3}
We will make use below also of the {\it residue} matrix defined by
\begin{equation}
R^{(\eta )}_{ij} =
\frac{1}{{\rm i} \eta }\, \lim_{k\to {\rm i} \eta } (k-{\rm i} \eta )\mathbb S_{ij}(k)\, , \qquad
{\rm i}\eta \in {\cal P} \, .
\label{residuem}
\end{equation}
We recall that in the above parametrization the angular variables $\alpha_i$
characterize the departure from criticality, the critical points corresponding\cite{Bellazzini:2008fu}
to the values $\alpha_i = 0, \pm \pi/2$. The classification and the stability
properties of these critical points have been extensively studied in the
literature\cite{nfll-99,lrs-02,coa-03,drs-06,Bellazzini:2006kh,Bellazzini:2008mn,Bellazzini:2008fu,hc-08,Bellazzini:2009nk}.
As already stated in the introduction, our goal here is to go beyond and explore the theory away from criticality.
The basic steps in the construction of the field $\varphi$, which satisfies
the wave equation (\ref{eqm1}) and the off-critical boundary condition (\ref{bc}),
are given in Appendix \ref{appA}. The subtle point
is the contribution of the bound states of $\mathbb S(k)$, which need a
separate treatment where local commutativity turns out to be essential.
The dual field $\phd$ is defined by
\begin{align}
\partial_{t}\phd(t,x,i)=&-\partial_{x}\varphi(t,x,i)\, ,\\
\partial_{x}\phd(t,x,i)=&-\partial_{t}\varphi(t,x,i)\, ,
\label{D1}
\end{align}
and the solution of the Luttinger model
on $\Gamma$ can be expressed in terms of the pair $\{\varphi,\, \phd\}$. For the details we
refer to Refs. \onlinecite{Bellazzini:2006kh,Bellazzini:2008mn,Bellazzini:2008fu},
recalling here only the solution
\begin{eqnarray}
\psi_1(t,x,i) &\sim&
:{\rm e}^{{\rm i} \sqrt {\pi} \left [\zeta_+ \varphi (vt, x,i) - \zeta_- \phd (vt, x,i)\right ]}:\,,
\label{psi1}\\
\psi_2(t,x,i) &\sim&
:{\rm e}^{{\rm i} \sqrt {\pi} \left [\zeta_+\varphi (vt, x, i) + \zeta_- \phd (vt+x)\right ]}:\,,
\label{psi2}
\end{eqnarray}
and the expression of the electric current (\ref{current})
\begin{equation}
j (t,x,i) = - \frac{v}{v_F \zeta_+ \sqrt {\pi }}\, \partial _x \varphi (vt, x, i) \, ,
\label{bcurrent}
\end{equation}
needed in the derivation of the conductance.
In (\ref{psi1}, \ref{psi2}) $: \cdots :$ denotes the normal product, whereas
\begin{eqnarray}
\zeta_\pm &=& \sqrt{|\kappa|} \left (
\frac{\pi \kappa v_F+2g_+}{\pi \kappa v_F+2g_-}\right )^{\pm \frac{1}{4}}\, ,
\label{z}\\
v&=&\frac{\sqrt{(\pi \kappa v_F+2g_-)(\pi \kappa v_F+2g_+)}}{\pi|\kappa|}\, ,
\label{v}
\end{eqnarray}
$\kappa$ being the {\it statistical parameter}.
The conventional fermionic Luttinger liquid
is obtained for $\kappa =1$. For $\kappa \not=1$ one has {\it anyonic}
Luttinger liquids. Notice also that the above solution of the Luttinger
model is meaningful for coupling constants satisfying $2g_\pm >-\pi \kappa v_F$.
\bigskip
\subsection{Symmetry content}
The bulk theory, defined by the Lagrangian (\ref{lagr}), is invariant under
time reversal and global $U(1)$ gauge transformations. Attempting to lift
these symmetries to theory on whole star graph $\Gamma$,
one gets some restrictions on the boundary conditions (\ref{bc}). Time reversal
symmetry implies\cite{Bellazzini:2009nk} that $\mathbb U$ must be symmetric,
\begin{equation}
\mathbb U^t = \mathbb U \, .
\label{T1}
\end{equation}
Concerning the conservation of the $U(1)$ (electric) charge
\begin{equation}
Q= \sum_{i=1}^n \int_0^\infty {\rm d} x\, j(t,x,i)
\label{echarge}
\end{equation}
on $\Gamma$, one knows already from classical electrodynamics that the Kirchhoff's rule
\begin{equation}
\sum_{i=1}^n j(t,0,i)= 0
\label{K1}
\end{equation}
must be satisfied at the vertex of $\Gamma$. Using (\ref{bcurrent}), one can
verify\cite{Bellazzini:2006kh,Bellazzini:2008mn,Bellazzini:2008fu} that (\ref{K1})
holds if and only if
\begin{equation}
\sum_{i=1}^{n}\mathbb S_{ij}(k)=1\, , \quad \forall j =1,...,n\, ,
\label{K2}
\end{equation}
or, equivalently
\begin{equation}
\sum_{i=1}^{n}\mathbb U_{ij} = 1\, , \quad \forall j =1,...,n\, .
\label{K3}
\end{equation}
Combining (\ref{residuem}) and (\ref{K2}), one gets
\begin{equation}
\sum_{i=1}^n R^{(\eta )}_{ij} = 0\, , \qquad
\forall \; {\rm i}\eta \in {\cal P} \, ,
\label{K4}
\end{equation}
which will be essential below for checking the Kirchhoff's rule for the conductance away from
criticality.
In what follows we assume that both (\ref{T1}) and (\ref{K3}) hold. The case with broken
time reversal has been analyzed recently in Ref. \onlinecite{Bellazzini:2009nk}.
\section{Basic features of the model away from criticality}
\subsection{Impact of the analytic structure of $\mathbb S(k)$}
The analytic properties of the scattering matrix $\mathbb S(k)$ deeply influence the
physics of the Luttinger liquid on $\Gamma$.
The simplest observable, one can investigate in order to illustrate this fact, is the
electric current $j$. More precisely, it is enough to study the relative two-point function,
which is fixed\cite{Bellazzini:2008cs} up to a real parameter $\tau$ by causality (local commutativity).
Postponing the discussion of the physical meaning of $\tau$ to
the end of this subsection, we consider first the current-current correlator following
from the definition (\ref{bcurrent}). In the Fock representation of the field $\varphi$, defined
by equations (\ref{dec1}), (\ref{sol1}) and (\ref{bfield}) in the Appendix \ref{appA}, one finds
\begin{widetext}
\begin{equation}
\langle j(t_1,x_1,i_1) j(t_2,x_2,i_2) \rangle =
\frac{v^2}{\left (2\pi \zeta_+v_f \right )^2}\left [ D_{i_1i_2}(vt_{12},x_1,x_2) +
A_{i_1i_2}(vt_{12},x_1,x_2) + B_{i_1i_2}(vt_1,vt_2,x_1,x_2;\tau )\right ] \, ,
\label{jj}
\end{equation}
where
\begin{equation}
D_{i_1i_2}(t,x_1,x_2) = -\delta_{i_1i_2} \left [d^2(t-x_{12}) + d^2(t-x_{12}) +
d^2(t+{\widetilde x}_{12}) +d^2(t-{\widetilde x}_{12})\right ]\, ,
\label{fd1}
\end{equation}
\begin{multline}
A_{i_1i_2}(t,x_1,x_2) = \sum_{{\rm i} \eta \in {\cal P}_-} R^{(\eta)}_{i_1i_2}\bigl \{d^2(t+{\widetilde x}_{12}) + d^2(t-{\widetilde x}_{12})
+\eta \left [d(t+{\widetilde x}_{12}) -d(t-{\widetilde x}_{12})\right ] \\ -
\eta^2\left [w_-(-\eta(t-{\widetilde x}_{12}) +w_+\left (\eta(t+{\widetilde x}_{12})\right ) \right ]\bigr \}\, ,
\label{fa1}
\end{multline}
\begin{multline}
B_{i_1i_2}(t_1,t_2,x_1,x_2;\tau) = \sum_{{\rm i} \eta \in {\cal P}_+} R^{(\eta)}_{i_1i_2}
\bigl \{d^2(t_{12}+{\widetilde x}_{12}) + d^2(t-{\widetilde x}_{12})
+\eta \left [d(t_{12}+{\widetilde x}_{12}) -d(t_{12}-{\widetilde x}_{12})\right ] \\ -
\eta^2\left [w_+(-\eta(t_{12}-{\widetilde x}_{12}) +w_-\left (\eta(t_{12}+{\widetilde x}_{12})\right ) \right ]
+2\eta^2{\rm e}^{-\eta {\widetilde x}_{12}}\left [\cosh(\eta ({\widetilde t}_{12}-2\tau)) -{\rm i} \sinh (\eta t_{12})\right ] \bigr \}\, .
\label{fb1}
\end{multline}
\end{widetext}
Here and in what follows $t_{12}= t_1-t_2$, $x_{12}= x_1-x_2$,
${\widetilde t}_{12}= t_1+t_2$, ${\widetilde x}_{12}= x_1+x_2$ and
\begin{equation}
d (\xi ) = \frac{1}{\xi - i\epsilon } \, , \qquad
w_{\pm}(\xi)={\rm e}^{-\xi}\, {\rm Ei} (\xi \pm i\epsilon ) \, ,
\label{dw}
\end{equation}
with $\epsilon > 0$ and $\rm Ei$ the exponential integral function.
By construction the functions $A_{i_1i_2}$ and $B_{i_1i_2}$ collect
the contributions of the antibound and the bound states respectively.
The main feature distinguishing these two functions is their time dependence.
Notice indeed that $A_{i_1i_2}$ depends exclusively on $t_{12}$, whereas $B_{i_1i_2}$
depends in addition on ${\widetilde t}_{12}$ and on the parameter $\tau$.
This fact suggest the consideration of two separate cases:
(i) When bound states are absent (${\cal P}_+ = \emptyset $), the behavior of the system
is quite orthodox; the function $B_{i_1i_2}$ vanishes identically
and time translations invariance is preserved. Accordingly, the energy
is conserved. Because of (\ref{T1}), the theory is also invariant under time reversal
\begin{equation}
t \rightarrow -t \, .
\label{str}
\end{equation}
(ii) When bound states are present (${\cal P}_+ \not= \emptyset $) the situation
changes drastically. Besides on $t_{12}$, the dynamics depends also on
${\widetilde t}_{12}$, implying that the theory is no longer invariant under time translations.
Therefore, the energy is not conserved. The theory depends on the parameter $\tau$
as well and is invariant under the time reversal transformation
\begin{equation}
t \rightarrow -t +2\tau \, ,
\label{mtr}
\end{equation}
which is actually a reflection with respect to $\tau$.
The properties collected in point (ii) deserve a more detailed discussion.
They imply that for boundary conditions admitting bound states
(i.e. some $\alpha_i \in (0\, ,\pi/2)$ - see eqs.(\ref{bc}, \ref{d1}-\ref{p1})),
the system is not isolated. In fact, there exists a nontrivial energy flow
crossing the boundary\cite{f4} at $x=0$, the direction
(outgoing or incoming) being controlled by the parameter $\tau$.
In order to illustrate this statement, let us consider the energy density $\theta $ of the TL
model on $\Gamma$. In bosonic coordinates $\theta $ takes the form\cite{DellAntonio:1971zt}
\begin{equation}
\theta (t,x,i) = \frac{g_-}{\pi \zeta_-^2} (\partial_x \varphi)^2(t,x,i) +
\frac{g_+}{\pi \zeta_+^2} (\partial_x \phd\, )^2(t,x,i) \, .
\label{ed}
\end{equation}
As well known, the field products at coinciding points in (\ref{ed}) contain divergences
which must be subtracted. A natural way to fix the subtraction is to take as a reference
point the vacuum energy $\langle \theta (t,x)\rangle_{\rm line}$ on the line.\cite{f5} Using
conventional point splitting regularization, one gets in this way\cite{Bellazzini:2008cs}
\begin{multline}
{\cal E} (t,x,i) = \langle \theta (t,x,i)\rangle - \langle \theta (t,x)\rangle_{\rm line} \sim \\
{\cal E}(x,i)+ \sum_{{\rm i} \eta \in {\cal P}_+} \eta^2 R^{(\eta)}_{ii} {\rm e}^{-2\eta x} \cosh [2\eta (t-\tau)] \, ,
\label{venergy}
\end{multline}
where ${\cal E}(x,i)$ is a time independent contribution,
derived in Ref. \onlinecite{Bellazzini:2006jb}} but
irrelevant for what follows. From (\ref{venergy}) we deduce
that $\tau$ is the instance in which the vacuum energy flow inverts its direction,
being outgoing for $t<\tau$ and incoming for $t>\tau$. This feature is related to the fact
that the theory is invariant under the operation (\ref{mtr}), which obviously
exchanges the two time intervals $(-\infty\, ,\, \tau)$ and $(\tau\, ,\, \infty )$.
These considerations fix the physical meaning of the parameter $\tau$,
which appears in the theory when ${\cal P}_+ \not= \emptyset $.
Summarizing, the behavior of off-critical Luttinger junctions is characterized by
the two different regimes described in points (i) and (ii) above. In the case
(i) one deals with an isolated system in equilibrium. Both energy and electric
charge are conserved. The presence of bound states in the regime (ii)
significantly modifies this behavior. The energy is no longer conserved
because the relative boundary conditions determine a specific boundary
interaction with the environment, which drives the system out of equilibrium.
A remarkable feature is that the corresponding non-equilibrium state is
fixed, up to the value of the parameter $\tau$, by the basic physical
requirement of causality.
We stress in conclusion that the above results are obtained in an abstract
setup, where we attempt to describe the physics of a quantum wire junction
by a Luttinger liquid with specific off-critical boundary conditions at the vertex
of the star graph approximating the junction. Further investigations are
needed for clarifying the plausibility of these assumptions and the applicability
of the results to real-life quantum wire junctions. The case (i) looks
more physical since the energy is conserved and bounded from below.
In our opinion however, also the regime (ii) can not be excluded a priori, if one
considers the possible interactions of the junction with the environment. In order to clarify
the situation, one should analyze in the above framework some physical
observables, which in principle can be checked experimentally.
As an example, we derive in the next subsection the electromagnetic
conductance in explicit form. Another attractive possibility to apply the
above results is the development of an effective description of the Luttinger
liquid on complex quantum wire networks with several junctions and loops,
crossed possibly by magnetic fluxes implementing the interaction with the environment.
The complete field theory analysis of such networks
is usually a complicated problem. One approximate way to face the
problem could be to use the star product approach \cite{KS,Schrader:2009su},
the ``gluing" technique \cite{Mintchev:2007qt, Ragoucy:2009hf,Caudrelier:2009ay}
or transfer matrix formalism\cite{Khachatryan:2009xg} for deriving the {\it effective}
scattering matrix $\mathbb S_{\rm eff}(k)$ relative to the {\it external} edges of the network.
$\mathbb S_{\rm eff}(k)$ admits in general both bound and antibound states
and can be used\cite{Ragoucy:2009hf} for constructing a simplified model with one effective
off-critical junction.
\bigskip
\subsection{Off-critical conductance}
In order to compute the conductance, we couple
the Luttinger liquid on $\Gamma$ to a uniform {\it classical} electric field\cite{f6}
$E(t,i)=\partial_t A_x(t,i)$, performing in (\ref{lagr}) the substitution
\begin{equation}
{\partial}_x \longmapsto {\partial}_x + {\rm i} A_x (t,i) \, .
\label{covder}
\end{equation}
This external field coupling deforms the bosonized version of the current according to\cite{Bellazzini:2006kh}
\begin{multline}
j(t,x,i) \longmapsto J(t,x,i) = \\
- \frac{v}{v_F \zeta_+ \sqrt {\pi }}\left [ \partial _x \varphi (vt, x, i) + \frac{1}{\zeta_+ \sqrt {\pi }} A_x(t,i)\right ]
\label{Jcurrent}
\end{multline}
The Hamiltonian encoding the interaction of $\varphi$ and $A_x$ is time dependent and reads\cite{Bellazzini:2006kh}
\begin{equation}
H_{\rm int}(t) =
\frac{1}{\zeta_+ \sqrt{\pi }}\sum_{i=1}^n \int_0^\infty {\rm d} x ({\partial}_x \varphi)(vt,x,i) A_x(t,i) \, .
\label{hint}
\end{equation}
We stress that now the system can exchange energy with the environment in
two different ways. The first one corresponds to the external force
produced by the coupling with the time dependent electric field $E(t,i)$.
The second one represents an intrinsic property in the regime (ii), being related to the presence
of bound states of $\mathbb S(k)$. The conductance, derived below, keeps track of both of them.
In order to derive the conductance, one should compute the expectation value of the current (\ref{Jcurrent})
in the external field $A_x$. This expectation value is given by the following
well known\cite{FW} series expansion
\begin{widetext}
\begin{multline}
\langle J(t,x,i)\rangle_{A_x} =
\langle J(t,x,i) \rangle -{\rm i}\int_{-\infty}^{t} {\rm d}\tau_1
\langle [J(t,x,i)\, ,\, H_{\rm int}(\tau_1)]\rangle + \cdots \\
+(-{\rm i})^n \int_{-\infty}^{t} {\rm d}\tau_1 \int_{-\infty}^{\tau_1} {\rm d}\tau_2 \cdots \int_{-\infty}^{\tau_{(n-1)}} {\rm d}\tau_n
[\cdots[[J(t,x,i)\, ,\, H_{\rm int}(\tau_1)]\, ,\, H_{\rm int}(\tau_2)]\, ,\cdots , H_{\rm int}(\tau_n)] + \cdots
\label{lrt1}
\end{multline}
in powers of $H_{\rm int}$. This expansion is particularly simple in our case.
In fact, from (\ref{Jcurrent},\ref{hint}) one obtains that
\begin{equation}
[[J(t,x,i)\, ,\, H_{\rm int}(\tau_1)]\, ,\, H_{\rm int}(\tau_2)]=0\, ,
\label{dcomm}
\end{equation}
cutting the infinite series (\ref{lrt1}) to the sum of the first two terms, namely
\begin{equation}
\langle J(t,x,i)\rangle_{A_x} =
\frac{v}{v_F \zeta_+^2 \pi }\left [A_x(t,i)+
{\rm i} \sum_{j=1}^{n}\int_{-\infty}^{t} {\rm d}\tau
\int_0^\infty {\rm d} y A_y(\tau,j)
\langle [\partial_y\varphi(v\tau,y,j)\, ,\, \partial_x\varphi(vt,x,i)]\rangle \right ]\, .
\label{lrt2}
\end{equation}
The expectation value $\langle J(t,x,i)\rangle_{A_x}$ is therefore linear in $A_x$, implying that the
linear response approximation\cite{FW} to $\langle J(t,x,i)\rangle_{A_x}$ is actually exact in this case.
Let us suppose now that the external field is switched on at
$t=t_0$, i.e. $A_x(t,i) = 0$ for $t<t_0$.
Using the representation (\ref{comm}) for the commutator in the left hand side of (\ref{lrt2}),
one finds for $t >t_0$
\begin{equation}
\langle J(t,0,i)\rangle_{A_x} =
G_{\rm line}\sum_{j=1}^n\int_{-\infty}^\infty \frac{{\rm d} \omega}{2\pi}
{\hat A}_x(\omega, j){\rm e}^{-{\rm i} \omega t}
\left [\delta_{ji} - S_{ji}\left (\frac{\omega}{v} \right ) -
\sum_{{\rm i} \eta \in {\cal P}} R_{ji}^{(\eta)} \frac{v \eta}{v \eta + {\rm i} \omega}
{\rm e}^{(t-t_0)(v \eta+{\rm i} \omega)} \right ]\, ,
\label{lrt3}
\end{equation}
where
\begin{equation}
G_{\rm line} = \frac{v}{v_F \zeta_+^2 \pi }\, , \qquad
{\hat A}_x(\omega, i) = \int_{-\infty}^\infty {\rm d} \tau\, {\rm e}^{{\rm i} \omega \tau} A_x(\tau,i)\, .
\label{Gline}
\end{equation}
The conductance can be extracted directly from (\ref{lrt3}). The result is
\begin{equation}
G_{ij}(\omega, t-t_0) =
G_{\rm line} \left [\delta_{ij} - S_{ij}\left (\frac{\omega}{v} \right ) -
\sum_{{\rm i} \eta \in {\cal P}} R_{ij}^{(\eta)} \frac{v \eta}{v \eta + {\rm i} \omega}
{\rm e}^{(t-t_0)(v \eta+{\rm i} \omega)} \right ]\, , \quad t >t_0\, ,
\label{C2}
\end{equation}
\end{widetext}
which represents the off-critical generalization of the formula (\ref{C1}) we are looking for.
Notice that away from criticality $G_{ij}(\omega, t-t_0)$ is in general complex, showing
that off-critical junctions may have a nontrivial impedance. Since eq. (\ref{lrt2}) involves
the commutator of the field $\varphi$, the conductance does not depend on the parameter
$\tau$, but depends on the time $t-t_0$ elapsed after switching on the external field.
Before discussing the main features of (\ref{C2}), we would like to mention two
useful checks. First of all we observe that at criticality, where scale invariance
implies that $\mathbb S$ is constant ($k$-independent),
the sum over the poles vanishes and (\ref{C2}) precisely reproduces the expression
(\ref{C1}) from the introduction. A second highly nontrivial check is the Kirchhoff
rule
\begin{equation}
\sum_{i=1}^n G_{ij}(\omega, t-t_0) = 0 \, ,
\label{K5}
\end{equation}
which is a consequence of the conservation of the electric charge, namely
of equations (\ref{K2}, \ref{K4}).
The sum in the right hand side of (\ref{C2}) runs over both negative and
positive poles ${\cal P} = {\cal P}_-\cup {\cal P}_+$.
The antibound states ${\cal P}_-$ produce damped oscillations in $t-t_0$.
If bound states are absent (${\cal P}_+ = \emptyset$),
\begin{multline}
\lim_{t\to \infty} G_{ij}(\omega, t-t_0) =
\lim_{t_o\to -\infty} G_{ij}(\omega, t-t_0) = \\
G_{\rm line} \left [\delta_{ij} - S_{ij}\left (\frac{\omega}{v} \right )\right ] \, ,
\label{limits}
\end{multline}
which gives the conductance one will observe in the regime (i) long time after
switching on the external field.
{}Finally, the bound states ${\cal P}_+$ give origin to oscillations
whose amplitude is growing exponentially with $t-t_0$.
The oscillations in (\ref{C2}) provide therefore a nice experimental
signature for testing the analytic structure of the scattering matrix $\mathbb S(k)$.
\section{Outlook and conclusions}
Off-critical Luttinger junctions are characterized by a scattering matrix $\mathbb S(k)$ which admits in general
bound and antibound states. The presence/absence of bound states determines
two different regimes of the theory.
A scattering matrix without bound states gives raise to an isolated equilibrium
system. One bound state is enough to change radically the situation.
We have show in fact that each such a state generates an oscillator degree of freedom,
whose contribution is fixed by local commutativity up to a common free parameter $\tau$.
These additional degrees of freedom break the invariance under time translations
and drive the system out of equilibrium. Accordingly,
the vacuum energy is time dependent: it decays exponentially in the interval
$(-\infty, \tau)$ and grows at the same rate in $(\tau, \infty )$. These two intervals are
related by time reversal and are characterized by a nontrivial outgoing and incoming vacuum energy flows.
Recalling that the traditional way \cite{CL} for driving a system out of equilibrium is to couple it
with a ``bath" of external oscillators, we have shown above that under certain conditions such oscillators
can be automatically generated by boundary effects. Junctions with bound states provide
therefore an intrinsic mechanism for constructing non-equilibrium quantum systems,
whose behavior is governed by purely boundary phenomena.
Since this mechanism is based on the general physical requirement of
causality, it extends\cite{BMS} also to systems in space dimensions greater then one.
The off-critical conductance formula (\ref{C2}) has various interesting properties.
A first remarkable feature of (\ref{C2}) is the different
impact of antibound and bound states, which hopefully can be tested experimentally.
The result (\ref{C2}) turns out to be useful also at the level of effective theory for quantum wire
networks with several junctions, where it applies for the effective scattering matrix
$\mathbb S_{\rm eff}(k)$ associated to the external edges of the network.
\acknowledgments
We thank Pasquale Calabrese, Benoit Dou\c{c}ot, In\`es Safi and Robert Schrader for
enlightening discussions and correspondence.
The research of B. B. has been supported in part by the NSF grant PHY-0757868.
|
\section{Introduction}
Although lattice QCD has been used successfully for simulations at zero and finite temperatures
and at zero density,
Monte Carlo simulations at non-zero densities suffer from
a technical problem: the lattice QCD action becomes complex, which
prevents its customary probabilistic interpretation.
In principle one could perform simulations at zero density, and use the
reweighting technique to obtain information at finite densities.
An early attempt known as the Glasgow method~\cite{Glasgow} did not work
due to the overlap problem: the configurations at zero density were too
``far'' from the target configurations at non-zero densities.
Considerable progress has been accomplished by generalizing the Glasgow method
to two-parameter reweighting~\cite{Fodor}.
Nevertheless, the range of reliability of this technique is difficult to
assess, and its failure can go undetected.
Therefore, another, more conservative way to deal with finite baryon densities
may be useful. It consists of calculating Taylor coefficients of observables
with respect to the chemical potential $\mu$ about $\mu=0$.
Those Taylor coefficients can be expressed as expectation values of
complicated observables, which can be measured at zero density. Thus,
there is no difficulty to perform Monte Carlo simulations in this method.
A first, pioneering attempt to obtain quark susceptibilities~\cite{Gottlieb}
has been followed by numerous works, obtaining in particular
the response of screening masses to chemical potential~\cite{QCDTARO,QCDTARO2,QCDTARO3}.
The Taylor expansion method has also been used for
studies of the equation of state, of the phase transition and of higher order susceptibilities~\cite{Allton,Allton2,Gupta,Gupta2}.
However, the complexity of the observable representing the Taylor coefficient, and the computer
effort to measure it, increase rapidly with the order of the Taylor expansion.
This motivates us to follow a different strategy.
Since no difficulty appears for simulations at imaginary chemical potential $\mu=i \mu_I$,
one can obtain information at finite baryon densities by analytic continuation of
observables measured at finite $\mu_I$.
Actually, this imaginary chemical potential strategy has been applied
with success to the determination of the phase transition~\cite{Forcrand}.
In this study, we perform simulations at finite $\mu_I$ and
measure derivatives of the pressure {\em as a function of $\mu_I$}.
These derivatives contain information about the Taylor coefficients
of the $\mu=0$ expansion, which can be extracted by fitting.
Finally, we try to reconstruct the equation of state at finite baryon and isospin densities.
The strategy of our method and preliminary results were presented rather long
ago in \cite{GENOA}.
Here we report further progress on this project.
A related approach, where the quark density is measured at imaginary quark and real isospin chemical potentials and then fitted by a polynomial ansatz, has recently been presented in
\cite{DELIA}.
\section{Equation of State at Finite Chemical Potential}
The lattice QCD partition function with $N_f$ flavors of staggered fermions can be written as
\begin{equation}
Z=\int \Pi_i^{N_f} \det M(U,m_i,\mu_i)^{1/4}\exp(-S_g[U])dU,
\end{equation}
where $S_g[U]$ is the gauge action and $M(U,m_i,\mu_i)$ stands for
the staggered Dirac operator with quark mass $m_i$ and chemical potential $\mu_i$\footnote{We set aside potential problems with ``rooting'' the determinant,
particularly at non-zero chemical potential.}.
In this study we consider $N_f=2$ degenerate fermion species
and use the standard Wilson gauge action.
The pressure or the equation of state with chemical potential $\mu_u$ and $\mu_d$ is given by
\begin{equation}
p(\mu_u,\mu_d) =-\frac{F}{V}=\frac{T}{V}\ln Z(\mu_u,\mu_d),
\end{equation}
and can be expanded in a Taylor series about $\mu_u=\mu_d=0$ as
\begin{equation}
\frac{\Delta p}{T^4} \equiv \frac{p(\mu_u,\mu_d)- p(0,0)}{T^4} =
\sum_{n,m=1}\frac1{n!m!}f_{nm}\left(\frac{\mu_u}{T}\right)^n \left(\frac{\mu_d}{T}\right)^m,
\end{equation}
where $f_{nm}$ are the Taylor expansion coefficients.
They vanish when $(n+m)$ is odd due to CP symmetry.
Furthermore, for equal quark masses there is another symmetry $f_{nm}=f_{mn}$.
The $f_{nm}$'s are related to derivatives $\chi_{ij}$ of the pressure measured
at {\em non-zero} chemical potential by
\begin{equation}
T^{i+j-4} \chi_{ij}=\frac{\partial^{i+j} (p(\mu_u,\mu_d)/T^4)}
{\partial(\mu_u/T)^i \partial(\mu_d/T)^j}
=\sum_{n=i,m=j}\frac1{(n-i)!(m-j)!}f_{nm}\left(\frac{\mu_u}{T}\right)^{n-i} \left(\frac{\mu_d}{T}\right)^{m-j}.
\label{Deri}
\end{equation}
While at zero density $\chi_{ij}=f_{ij} T^{4-i-j}$,
at non-zero densities $\chi_{ij}$ includes higher order $f_{nm}$ terms,
and does not vanish for odd $(i+j)$.
This suggests to
use all available $\chi_{ij}$'s at non-zero densities, in order to estimate the $f_{nm}$'s.
Here, we try to estimate ${f_{nm}}$ by fitting all $\chi_{ij}$ simultaneously to the polynomial expansions eq.(\ref{Deri}).
Of course, $\chi_{ij}$ at non-zero baryon density is not directly obtainable from simulations on the lattice because of the sign problem.
However, $\chi_{ij}$ can be obtained through simulations at imaginary quark chemical potential or at real isospin density.
Here, we calculate $\chi_{ij}$ at imaginary chemical potentials.
Therefore, we set $\mu=i\mu_I$.
Each $\chi_{ij}$ depends on higher order Taylor coefficients following eq.(\ref{Deri}).
Therefore, with sufficiently accurate data on $\chi_{ij}$ one can also obtain higher order Taylor coefficients $f_{nm}$, $n>i, m>j$.
The measurements of the derivatives involve computing traces of inverse Dirac matrix products.
These traces were estimated using the noise method with 40 $Z_2$ random vectors.
In this study we measure $\chi_{ij}$ up to $i+j=4$. Thus we have 8 different $\chi_{ij}$'s.
We fit all the data to the corresponding 8 polynomial expansions eq.(\ref{Deri}) truncated to
a given order $(n+m)$, and try to obtain
the Taylor coefficients $f_{nm}$.
As we will see, in the confined phase a Taylor expansion is not the most compact description
of the pressure.
Instead, for $T\leq T_c$ we use the Hadron Resonance Gas (HRG) model.
In the HRG model the pressure is given as\footnote{This expression is taken from (4.3) in \cite{Allton2}.}
\begin{eqnarray}
\label{HRG}
\frac{\Delta p(\mu_u,\mu_d)}{T^4} & = &G[\cosh(\frac{2\mu_{Is}}{T})-1] +R[\cosh(\frac{3\mu_q}{T})\cosh(\frac{\mu_{Is}}{T})-1] \\
&+&W[\cosh(\frac{3\mu_q}{T})\left(\cosh(\frac{\mu_{Is}}{T})+\cosh(\frac{3\mu_{Is}}{T})\right)-2], \nonumber
\end{eqnarray}
where $G$,$R$ and $W$ are constants related to the hadron spectrum,
and
quark and isospin chemical potentials $\mu_q$ and $\mu_{Is}$ are defined as
$\mu_q=(\mu_u+\mu_d)/2$ and $\mu_{Is}=(\mu_u-\mu_d)/2$ respectively.
The derivatives of the pressure with respect to $\mu_u$ and $\mu_d$,
instead of having the polynomial form eq.(\ref{Deri}), are now obtained by differentiating eq.(\ref{HRG}).
The coefficients $G$,$R$ and $W$ are then extracted by fitting imaginary-$\mu$
data.
In terms of $G$,$R$ and $W$, the first Taylor coefficients are given by
\begin{equation}
f_{20}=\left(G+\frac52 R +7W\right),
\end{equation}
\begin{equation}
f_{11}=-\left(G+2(R+W)\right),
\end{equation}
\begin{equation}
f_{22}=G+4(R+W),
\end{equation}
\begin{equation}
f_{31}=G+\frac72 R+ 49W,
\end{equation}
\begin{equation}
f_{40}=-G+5(R+W).
\end{equation}
\begin{table}[h]
\centering
\caption{$\chi^2/dof$ for polynomial ansatz of degree 4, 6 and 8 (maximum value of $(n+m)$ in eq.(\protect\ref{Deri})).}
\label{tab:1}
\begin{tabular}{cllllllllllllllll}
\hline
$T/T_c$ & 0.99 & 1.00 & 1.03 & 1.04 & 1.065 & 1.085 & 1.1 & 1.2 & 1.3 & 1.4 & 1.5 & 2.0 \\
\hline
4th & 85.1 & 134.9& 3.15 & 3.14 & 3.50 & 7.24 & 3.20 & 10.9 & 11.3 & 6.67 & 9.95 & 9.15 \\
6th & 19.1 & 42.1 & 1.60 & 2.19 & 0.82 & 5.50 & 5.53 & 1.52 & 0.89 & 2.46 & 1.09 & 2.10 \\
8th & 4.53 & 5.29 & 1.64 & 1.77 & 0.81 & 1.01 & 2.15 & 1.72 & 0.91 & 2.16 & 1.22 & 1.28 \\
\hline
\end{tabular}
\end{table}
\begin{table}[h]
\centering
\caption{$\chi^2/dof$ for HRG ansatz.}
\label{tab:2}
\begin{tabular}{clllllllllllllll}
\hline
$T/T_c$ & 0.83 & 0.9 & 0.95 & 0.98 & 0.99 & 1.00 \\
\hline
HRG & 1.29 & 1.00 & 2.10 & 15.8 & 10.5 & 29.4 \\
\hline
\end{tabular}
\end{table}
\begin{figure}
\vspace{5mm}
\centering
\includegraphics[height=5.5cm]{chisq_poly.eps}
\caption{
$\chi^2/dof$ of various polynomial ans\"atze as a function of $T/T_c$.
The fitting range of $a\mu_I$ is 0.0-0.24.}
\label{fig:chisq2}
\end{figure}
\begin{figure}
\vspace{5mm}
\centering
\subfigure[]{
\includegraphics[height=5.5cm]{c11all2-2.eps}
}~
\subfigure[]{
\includegraphics[height=5.5cm]{c20all2-2.eps}
} \\
\vspace{5mm}
\subfigure[]{
\includegraphics[height=5.5cm]{c22all2.eps}
}~
\subfigure[]{
\includegraphics[height=5.5cm]{c40all2.eps}
} \\
\vspace{5mm}
\subfigure[]{
\includegraphics[height=5.5cm]{c31all2.eps}
}
\caption{
Taylor coefficients: (a)$f_{11}$, (b)$f_{20}$, (c)$f_{22}$, (d)$f_{40}$ and (e)$f_{31}$.}
\label{fig:Taylor}
\end{figure}
\begin{figure}
\vspace{-5mm}
\centering
\subfigure[]{
\includegraphics[height=5.5cm]{b553pressure.eps}
}~
\subfigure[]{
\includegraphics[height=5.5cm]{b490pressure.eps}
}
\caption{
Equation of state (Pressure) as a function of $a\mu_q$ and $a\mu_{Is}$ at (a) $\beta=5.53 (T/T_c\sim 1.1)$ and at (b)$\beta=4.90 (T/T_c\sim 0.83)$.}
\label{fig:Pb553}
\end{figure}
\section{Simulations at Imaginary Chemical Potential}
We have performed simulations on $8^3 \times 4$ lattices
at a quark mass $m_q=0.05$ and imaginary chemical potentials $a\mu_I=0.0,\dots,0.24$.
We have chosen 16 values of $\beta$ ranging from 4.90 to 6.85, which correspond to
$T/T_c=0.83 \sim 2.0$.
Most of the simulations were performed using
the R-algorithm with a step size $\Delta t=0.02$. We also used the Rational HMC algorithm~\cite{RHMC} to
check the systematic stepsize errors caused by the R-algorithm, and found
no significant difference for this lattice size and quark mass.
At each simulation point
we have accumulated 12000 to 20000 measurements.
The measurements were taken every 5 trajectories
to balance the computational effort of the R-algorithm simultion and measurements.
\subsection{Fitting to $\chi_{ij}$}
We determine $f_{nm}$ by fitting all the derivatives
simultaneously to the corresponding ansatz of $\chi_{ij}$.
We used the polynomial ansatz eq.(\ref{Deri}) for the data at $T/T_c\geq0.99$,
and the HRG ansatz eq.(\ref{HRG}) at $T/T_c\leq1.0$.
Tables 1 and 2 show the $\chi^2/dof$ for the polynomial and HRG fits, respectively.
The fitting range of $a\mu_I$ is $0.0-0.24$, which covers most of the range up to
the Roberge-Weiss transition at $\mu_I = \pi T/3$.
Fig.\ref{fig:chisq2} compares the $\chi^2/dof$ among various polynomial and HRG fits.
One can see that the 4th order polynomial ($(n+m)\leq 4$ in the expansion eq.(\ref{Deri}))
is not good over the whole temperature range, and
that the 6th order one becomes poor in the vicinity of $T_c$.
Similarly, one can also see that the quality of the fit based on the HRG ansatz becomes poor
for $T/T_c\ge 0.95$.
While the failure of the HRG ansatz near $T_c$ has been noticed before~\cite{Cheng,DELIA},
it is remarkable that we can see clear indications of 6th order, and even 8th order
Taylor coefficients with our modest study. The measurement of 8th order Taylor coefficients
represents the current state of the art~\cite{Gupta2}.
Fig.\ref{fig:Taylor} shows the Taylor coefficients $f_{11}$, $f_{20}$, $f_{22}$, $f_{40}$ and $f_{31}$ as a function of temperature.
Those results are obtained by fitting a 6th order polynomial in a range of $a\mu_I=0.0-0.24$.
For $T/T_c\ge 1.0$, they agree well with those obtained
from the direct measurement of derivatives at $\mu=0$, i.e. $\chi_{ij}|_{\mu=0}$,
but are more accurate.
We do not show the 6th order Taylor coefficients $f_{60}, f_{51}, f_{42}$ and $f_{33}$: even though their collective effect is statistically significant, they cannot be individually
determined with any statistical accuracy. We only observe that $f_{60}$ is dominant at this
order.
Similary, the Taylor coefficients obtained from the HRG ansatz for $T/T_c\le 0.95$
also agree well with direct measurements of $\chi_{ij}|_{\mu=0}$, with higher accuracy.
However, for $T/T_c > 0.95$ the results from the HRG ansatz fits deviate from $\chi_{ij}|_{\mu=0}$.
This observation is consistent with the measured $\chi^2/dof$, which increase considerably for $T/T_c > 0.95$.
\subsection{Equation of State at Finite Densities}
Once we obtain the Taylor coefficients of the pressure or the parameters of the HRG model,
we can reconstruct the equation of state.
Here, we present two cases at $\beta=5.53(T/T_c\sim 1.1)$ and $\beta=4.90(T/T_c\sim 0.83)$
which are reconstructed with the Taylor series and the HRG ansatz, respectively.
Fig.~\ref{fig:Pb553}(a) shows the equation of state
at $\beta=5.53 (T/T_c\sim 1.1)$ as a function of $a\mu_q$ and $a\mu_{Is}$.
Similarly Fig.~\ref{fig:Pb553}(b) shows the equation of state
at $\beta=4.90 (T/T_c\sim 0.83)$.
One can also reconstruct other interesting quantities.
Fig.~\ref{fig:Numb553} shows the quark number density $N_q$ and the isospin number density $N_{Is}$
at $\beta=5.53$ as a function of $a\mu_q$ and $a\mu_{Is}$.
Similarly, Fig.~\ref{fig:Numb490} shows $N_q/T^3$ and $N_{Is}/T^3$ at $\beta=4.90$.
Here,
$N_q$ and $N_{Is}$ are defined as
$\displaystyle N_q = \frac{\partial p}{\partial \mu_q}$ and
$\displaystyle N_{Is}= \frac{\partial p}{\partial \mu_{Is}}$, respectively.
\begin{figure}
\vspace{-5mm}
\centering
\subfigure[]{
\includegraphics[height=5.5cm]{b553num.eps}
}~
\subfigure[]{
\includegraphics[height=5.5cm]{b553iso.eps}
}
\caption{
Number density at $\beta=5.53 (T/T_c\sim 1.1)$ as a function of $a\mu_q$ and $a\mu_{Is}$: (a)$N_q/T^3$ and (b) $N_{Is}/T^
3$.}
\label{fig:Numb553}
\end{figure}
\begin{figure}
\vspace{-5mm}
\centering
\subfigure[]{
\includegraphics[height=5.5cm]{b490num.eps}
}~
\subfigure[]{
\includegraphics[height=5.5cm]{b490Iso.eps}
}
\caption{
Number density at $\beta=4.90 (T/T_c\sim 0.83)$ as a function of $a\mu_q$ and $a\mu_{Is}$: (a)$N_q/T^3$ and (b) $N_{Is}/T
^3$.}
\label{fig:Numb490}
\end{figure}
\section{Conclusions}
We have performed simulations at imaginary chemical potentials and
measured the derivatives of the pressure with respect to $\mu$, at zero and non-zero imaginary $\mu$.
By fitting all the derivatives to a polynomial ansatz or an HRG ansatz,
we obtained the Taylor coefficients of the $\mu/T$ expansion of the pressure about $\mu=0$.
The Taylor coefficients obtained by a polynomial fit for $T/T_c \ge 1.0$
agree well with the direct measurement of derivatives at $\mu=0$, $\chi_{ij}|_{\mu=0}$,
but are more accurate.
Remarkably, we find it impossible to obtain a good fit, at any temperature, without including
6th order derivatives. For $T_c \leq T \leq 1.04 T_c$, 8th order derivatives are necessary.
Thus, our approach may provide a cheaper alternative to the direct measurement of high-order
derivatives at $\mu=0$.
Similarly, below $T_c$ we observed that the Taylor coefficients obtained by the HRG ansatz
deviate from $\chi_{ij}|_{\mu=0}$ for $T/T_c \ge 0.95$, and the HRG ansatz itself gives a
poor description of the imaginary-$\mu$ data.
The same observation has been made in \cite{DELIA}.
Finally, using the obtained Taylor coefficients we reconstructed the equation of state
and the number densities as a function of $\mu_q$ and $\mu_{Is}$ up to 4th order.
\section*{Acknowledgements}
The numerical calculations were carried out on SX8 at YITP in Kyoto University.
|
\section{Introduction}
\label{secdef}
An area that has been gaining increasing interest over the last years is that of out-of-equilibrium quantum physics. An example of particular simplicity is that of quantum quenches in which some of the parameters of the hamiltonian of an isolated quantum system are changed instantaneously. Then one practically has to study the time evolution of a trial wavefunction, which is typically the ground state of the hamiltonian before the quench, under the influence of the hamiltonian after the quench. Although one expects a periodic collapse and revival of the initial state, in practice this period diverges rapidly with the system size and for large systems local observables may exhibit stationary behaviour at long times, eventhough the global wavefunction itself may never become such. This has been shown to be the case in many different settings \cite{rig-06,rig-07,caz-06,cc-06,cc-07,l-07,gdlp-07,kol-08,bs-08,qr-08,ke-08,fcmse-08,cfmse-08,bpgda-09,ro-09,fm-09}.
An obvious interesting question is whether this stationary behaviour is thermal as one may reasonably expect. It turns out that in many integrable systems the stationary behaviour is described by a statistical distribution which is similar to but not exactly thermal \cite{rig-06,rig-07,caz-06,cc-07,ke-08,fm-09}.
More precisely it is a generalized Gibbs ensemble, subject to the constraints imposed by the integrals of motion. It was then conjectured that non-integrability is responsible for the exact thermalization of a system \cite{rig-08}. To what extend this is true is however still under investigation, since theoretical arguments that support this conjecture are based on semiclassical conjectures \cite{sr-94}, while numerical studies\cite{kol-07,mwnm-07,ro-09,rig-08,rig-09,bkl-09} lead to rather controversial results: some of them \cite{kol-07,mwnm-07,ro-09} reveal non-thermal behaviour even for non-integrable systems, while others \cite{rig-08,rig-09} are in good agreement with the thermal predictions and attribute the previous disagreement to finite size effects. There are some analytical studies in lattice models too \cite{qr-08,ek-08}: the first \cite{qr-08} refers to the Bose-Hubbard model but after the quench the system evolves under the free hamiltonian of the superfluid regime. In the second \cite{ek-08} dynamical mean-field theory (DMFT) applied to the non-integrable Falicov-Kimball model shows non-thermal features.
On the other hand, an interaction quench in the Fermi-Hubbard model is possible to lead to thermalization, as shown using two different analytical \cite{mk-08} and numerical DMFT \cite{ekw-09} approximations.
In the present work we will study quantum quenches employing a field theoretic approach, which is supposed to capture their essential general characteristics. We consider systems described by a relativistic dispersion relation with some energy gap (or mass) and a maximum group velocity of excitations. Then for free systems, a quantum quench of the energy gap leads to stationary behaviour and a momentum dependent effective temperature can be defined \cite{cc-06,cc-07}. This is true for quite general conditions: the energy gap after the quench must be nonzero in $1d$ and $2d$ while in $3d$ the result holds even if it is zero. Furthermore a $1d$ gapless system can only be interacting and it turns out that it exhibits similar behaviour too \cite{cc-06,cc-07}. It should be emphasized that the notion of effective thermalization used throughout the present and some related earlier work\cite{cc-07} refers to the thermal-like stationary behaviour that fits to the generalized Gibbs ensemble description rather than the standard thermal theory. This is manifest in the fact that each momentum mode corresponds to a different effective temperature, since in the absence of interactions each mode evolves independently from the others. This suggests that in interacting systems the energy exchange due to collisions between different momentum modes would result in a mixing of their effective temperatures. However if the interaction is such that the system is still integrable then there will be some other decomposition into independent modes (quasiparticles) and we expect that the system still exhibits stationary behaviour with a different effective temperature for each of these modes. Therefore it is only when the interaction makes the system non-integrable that thermalization to a unique common temperature is still a possibility.
The simplest interacting field theory is the $\phi^4$ model. We consider a simultaneous quench of the mass from $m_0$ to $m$ and of the coupling constant from $\lambda_0$ to $\lambda$. In order to study the evolution of the system we need to use an approximation scheme and the simplest one is the Hartree-Fock or self-consistent approximation. This can be applied in a number of different but equivalent ways. In perturbation theory it consists in ignoring all skeleton diagrams from which the diagrammatic expansions of correlation functions are constructed, except for the simplest one, i.e. the loop diagram. This turns out to be the same as approximating the system's state by gaussian wavefunctions or substituting the quartic interaction term in the hamiltonian by a quadratic one with a self-consistent coefficient. Notice however that in this simple approximation, collisions between particles of different momenta are neglected and this makes our approach incapable of answering the previous question about the relation between non-integrability and exact thermalization. Indeed, although the $\phi^4$ model is non-integrable, the Hartree-Fock approximation becomes exact only in the large-$N$ limit of the linear $\sigma$-model, i.e. the generalization of the $\phi^4$ model to an $N$-component field, which becomes integrable in this limit. Thus our approach provides the integrable counterpart of a non-integrable model that best approximates it. It is however the necessary first step towards understanding the effect of quantum quenches in interacting systems and should be expected to reveal some of their general features.
There is a significant number of publications that use the same method to study other closely related out-of-equilibrium problems, partially due to applications to cosmology. Cooper and Mottola \cite{como-87} make a detailed presentation of the method for the evolution of a general trial wavefunction and Boyanovsky et al. \cite{bo-93,bls-93,bdv-93} study the special case of a quench from the disordered to the ordered phase at large temperature and in $3d$. Also Wetterich et al. have studied the time evolution of out-of-equilibrium initial ensembles using a different method based on the numerical computation of the time-dependent effective action\cite{bw-98,bw-99}. Recently a remarkable numerical study based on the same method and including next-to-leading order effects in the large-$N$ expansion, has shown that an initial pure state evolves so that the reduced density matrix indeed thermalizes at large times\cite{gs-09}.
Using our approximation we find that the two point correlation function long after the quench is of the same form as the free correlation function but with a different mass that has to be determined self-consistently. This means that nothing really changes in terms of the relaxation of the system: once again it becomes stationary and a momentum dependent effective temperature can be defined, the only difference being that $m$ will be replaced by an effective mass $m^*$ which depends also on the coupling constant $\lambda$. The self-consistency equation for $m^*$ has always a real solution larger or equal to $m$. In the critical case $m=0$, we find that in $1d$ $m^*$ is also zero, but in $2d$ it becomes finite. This leads to the important conclusion that in $2d$, after a quench to zero mass which according to the above discussion would \emph{not} lead to relaxation if the system were free, now due to the presence of the interaction, it acquires a non-zero effective mass which allows it to relax. Furthermore, by studying the time evolution of the effective mass we find that if $m_0>m$ and $\lambda$ is sufficiently large then right after the quench the system is effectively set into an unstable state although it soon recovers its stability.
In the first part of this paper we focus on free systems which have been partially discussed earlier \cite{cc-06,cc-07}. Here we present an elegant simplified derivation of the quench propagator, develop an exact imaginary time formulation based on an earlier invented mapping to a slab geometry and define an average measure of the effective temperature first introduced in recent work \cite{scc-09}. These constitute a useful toolkit for many applications and extensions. For completeness we briefly report earlier results regarding $1d$ integrable systems with critical evolution. In the second part we study the composite quench in the $\phi^4$ model in the self-consistent approximation. This part is split into two sections: in the first we follow a heuristic approach based on perturbation theory and find an ansatz for the correlation function and in the second we start with the equations of motion and investigate the time evolution to verify the results obtained from our ansatz.
\section{Simple harmonic oscillator}
The simplest problem of a quantum quench one can start with is that of a simple harmonic oscillator whose frequency is quenched from $\omega_0$ to $\omega$. The hamiltonian before the quench is
\eq{H_0=\frac{1}{2} \pi^2 + \frac{1}{2} \omega^2_0 \phi^2 }
while after the quench it is
\eq{H=\frac{1}{2} \pi^2 + \frac{1}{2} \omega^2 \phi^2 }
The initial state is the ground state $|\Psi_0\rangle$ of $H_0$.
From a physical point of view, what happens is that $|\Psi_0\rangle$, as a trial state different from the ground state $|0\rangle$ of $H$, contains, compared to that, an energy excess which is distributed to the excitation levels of $H$. After the quench the evolution of the wavefunction in the Schr\"{o}dinger picture is given by
\eqq[evol]{|\Psi(t)\rangle = e^{-iHt}|\Psi_0\rangle = \sum{e^{-i(n+1/2)\omega t} |n\rangle\langle n|\Psi_0\rangle}}
where $|n\rangle$ is an arbitrary eigenstate of $H$.
It is trivial to observe that the evolution is periodic since after a period $T=2\pi/\omega$ the system returns back to the initial state, up to an irrelevant minus sign. This is a special case of quantum recurrence \cite{quant_rec}. In fact the wavefunction will exhibit periodicity or quasi-periodicity (i.e. it will return arbitrarily close to the initial state after sufficiently large time) in any system with \emph{discrete} energy eigenvalues. Systems with finite degrees of freedom always have such discrete spectra, while in the thermodynamic limit the spectrum becomes in general continuous and quantum recurrence may be lost. In practice even for finite but large systems, the corresponding period is usually so large that this periodicity is irrelevant.
\subsection{Propagator}
\label{secFFTprop}
We are also interested in the correlation function of the field operator $\phi$ at different times, i.e. the propagator $\langle \Psi_0 | \mathcal{T}\{\phi(t_1) \phi(t_2)\}| \Psi_0\rangle \equiv C_q(t_1,t_2)$ where $\mathcal{T}$ denotes time ordering. The time evolution of $x$ in the Heisenberg picture is given by the equations of motion
\eq{\ddot{\phi}+\omega^2 \phi = 0}
which can be solved easily
\eqq{\phi(t)=\phi(0)\cos \omega t + \pi(0)\frac{\sin \omega t}{\omega} }
We therefore have
\begin{multline}
\langle \Psi_0 | \phi(t_1) \phi(t_2) | \Psi_0\rangle = \langle \Psi_0|\phi^2(0)|\Psi_0\rangle \cos \omega t_1 \cos \omega t_2 + \\
+ \langle \Psi_0|\pi^2(0)|\Psi_0\rangle \frac{\sin \omega t_1 \sin \omega t_2}{\omega^2} + \\
+ \langle \Psi_0| \phi(0) \pi(0) + \pi(0) \phi(0) |\Psi_0\rangle \frac{\sin \omega (t_1+t_2)}{2 \omega} - \\
- i \frac{\sin \omega (t_1-t_2)}{2 \omega}\label{gen_prop}
\end{multline}
where the canonical commutation relation $[\phi(0),\pi(0)]=i$ has been taken into account in order to simplify the last term. All terms are symmetric under the interchange $t_1 \leftrightarrow t_2$ apart from the last one which is antisymmetric. Thus time ordering amounts to substituting $(t_1-t_2)$ in the last term by its absolute value.
It is now clear that the problem reduces to the calculation of the initial expectation values of $\phi^2(0)$, $\pi^2(0)$ and $\phi(0)\pi(0)+\pi(0)\phi(0)$. From the initial condition that the system lies in the ground state of $H_0$ we easily find
\begin{subequations}
\label{inexpval}
\begin{align}
& \langle \Psi_0| \phi^2 (0)|\Psi_0\rangle = \frac{1}{2 \omega_0} \\
& \langle \Psi_0| \pi^2 (0)|\Psi_0\rangle = \frac{\omega_0}{2} \\
& \langle \Psi_0| \phi(0) \pi(0) + \pi(0) \phi(0) |\Psi_0\rangle =0
\end{align}
\end{subequations}
and substituting into (\ref{gen_prop}) we obtain
\eem[q-prop]{C_q(t_1,t_2) = & \frac{(\omega -\omega_0)^2}{4 \omega^2 \omega_0} \cos \omega (t_1-t_2)+ \nonumber \\
& + \frac{\omega^2 -\omega_0^2}{4 \omega^2 \omega_0} \cos \omega (t_1+t_2)+ \frac{1}{2 \omega} e^{-i \omega|t_1-t_2|}}
Notice that we have separated the Feynman propagator $e^{-i \omega |t_1-t_2|}/2 \omega$ which, as expected, is the only term that survives if $\omega=\omega_0$ i.e. if there is no quench at all. Also notice that the only term that breaks time invariance is the second one.
\section{Linearly coupled oscillators (free fields)}
Let us now move on to study a system of linearly coupled harmonic oscillators or equivalently a free field theory. In general such a system is described by a quadratic hamiltonian of the form
\eq{H=\frac{1}{2} \sum_r{ \pi^2(r) + \frac{1}{2} \sum_{r,r'} K(r-r')( \phi(r) - \phi{(r')})^2}}
which can be easily diagonalised in momentum space where it takes the form
\eqq[Hk]{H=\sum_k{\frac{1}{2} \pi_k \pi_{-k} + \frac{1}{2} \omega^2_k \phi_k \phi_{-k} }}
We will assume a relativistic dispersion relation
\eqq[disp1]{\omega^2_k = c^2 k^2 + m^2 c^4}
with energy gap (or mass, in the language of quantum field theory) $m$ and speed of sound $c$. This can also describe successfully non-relativistic systems with the same energy gap $m$ and maximum velocity of excitations $c$.
The quantum quench that we will consider consists in an instantaneous change of the mass from $m_0$ to $m$. For brevity we can set $c=1$. An investigation of a quench of the speed of sound $c$ is done elsewhere\cite{scc-09}. As earlier, we assume that before the quench at $t=0$ the system lies in the ground state of the initial hamiltonian $|\Psi_0\rangle$. In addition the system is kept isolated from the environment before and after the quench.
In order to study the time evolution, it is sufficient to find the two-point correlation function i.e. the propagator
\eq{ \langle \Psi_0 |\mathcal T \{\phi(r_1,t_1)\phi(r_2,t_2)\}|\Psi_0\rangle \equiv C_q(t_1,t_2,r_1-r_2) }
since in a free theory all physical observables can be obtained from this. From (\ref{Hk}) we see that the system is decomposed into a set of independent momentum modes each of which evolves as a simple harmonic oscillator. Thus the propagator is simply the Fourier transform with respect to $k$ of the expression (\ref{q-prop}) with $\omega_{0k} = \sqrt{k^2+m_0^2}$ and $\omega_k = \sqrt{k^2+m^2}$
\eqq[q-prop2]{C_q(t_1,t_2, r) = \int{\frac{d^d k}{(2 \pi)^d} \; e^{i \boldsymbol{k} \cdot \boldsymbol{r}} C_q(t_1,t_2;k)}}
\subsection{Properties of the propagator}
\label{secpropprop}
Let us now study the physical properties of the equal time propagator in real space. For simplicity we will mainly use its asymptotic form for $m_0 \gg m$ and $t,r \gg m_0^{-1}$. This will be called the \emph{deep quench limit} and should obviously exhibit all characteristic features of a quantum quench since it is one of the two most extreme possibilities for the relation between the two masses. In this limit the propagator simplifies to
\eqq[integral1]{C_{dq}(r,t) = \int{\frac{d^d k}{(2 \pi)^d} \; e^{i \boldsymbol{k \cdot r}} \frac{m_0}{4 \omega_k^2} (1-\cos 2 \omega_k t) }}
The massless ($m=0$) and massive ($m\neq 0$) cases are different and should be investigated separately.
\subsubsection{Massless case}
In this case $\omega_k = |k|$ and after some algebra using Fourier transforms of common functions, we obtain the following exact results:
\begin{itemize}
\item $d=1$
\eqq[1dprop]{C_{dq}^{(1d)}(r,t) =
\begin{cases}
0 \quad &\text{ if } r>2t, \\
m_0 (2 t -r) /8 \quad &\text{ if } r<2t.
\end{cases}}
\item $d=2$
\eq{C_{dq}^{(2d)}(r,t) =
\begin{cases}
0 \quad &\text{ if } r>2t, \\
\frac{m_0 }{8 \pi} \log [({2 t+\sqrt{4 t^2-r^2}})/{r}] \quad &\text{ if } r<2t.
\end{cases}}
\item $d=3$
\eq{C_{dq}^{(3d)}(r,t) =
\begin{cases}
0 \quad &\text{ if } r>2t, \\
m_0 /16 \pi r \quad &\text{ if } r<2t.
\end{cases}}
\end{itemize}
In all dimensions we distinguish between two spacetime regions in which the behaviour of the propagator is qualitatively different: for $r>2t$ it is always zero, unlike for $r<2t$. This means that the correlations between two points at distance $r$ remain unchanged until $t=r/2$. Also notice that in $3d$ the propagator is time independent for $r<2t$.
\subsubsection{Massive case}
By evaluating the integral (\ref{integral1}) we notice that, as before, we have to distinguish between two spacetime regions in which the behaviour of the propagator is qualitatively different. If $r>2t$ then we can close the integration contour in the upper half of the complex $k$-plane and since there is no pole the integral is zero. If on the other hand $r<2t$ then for the time independent part of the integrand we close the integration contour in the upper half plane but for the time dependent part we have to rotate it by $90^\circ$ instead. Each part has single poles at $k=\pm i m$ and the outcome is nontrivial. Exact results cannot be found and we have to employ asymptotic methods for large $r$ and $t$. In particular using the stationary phase method we find that for fixed $r$ and large $t$ the time dependent part of the integral tends to zero like $t^{-d/2} \cos{2 m t}$. Also the rest decreases for large $r$ like $e^{-mr}/r^{(d-1)/2}$.
We thus conclude that the propagator changes sharply as we cross the lines $t=r/2$. Before this time there are no correlations between two distant points, while afterwards the two points become correlated. This feature, which is a direct consequence of the causality principle, is called the \emph{horizon effect}. Fig.~\ref{fig:horizon} illustrates the main features of the massive propagator in $1d$.
\begin{figure}[htbp]
\centering
\includegraphics[width=1.00\columnwidth]{horizon.eps}
\caption{\emph{Top}: Spacetime plot of the deep quench propagator $C_{dq}(r,t)$ in $1d$ and for $m=1$, as obtained by numerical integration of (\ref{integral1}). The horizon effect is clearly demonstrated. Outside the horizon the value is exactly zero. \emph{Bottom}: Time dependence of $C_{dq}(r,t)$ (blue line) at fixed distance $r=r_0=2$, denoted by the vertical red line in the above figure. The dashed lines give the large time asymptotic expressions. Notice the decaying oscillations $\sim t^{-1/2} \cos{2 m t}$ (purple line) around the stationary value $\sim e^{-mr}$ (red line). }
\label{fig:horizon}
\end{figure}
Another particularly important conclusion is that if $m \neq 0$ then for fixed distance the propagator becomes \emph{stationary} for large times. The same is true for $m=0$ in $3d$, but not in $1d$ or $2d$. In addition this result is robust and does not rely on the deep quench approximation. Indeed if we use the full expression of $C_q(t;k)$ (\ref{q-prop}) for $m=0$ and $3d$ we find that the time dependence decays exponentially. The $1d$ case is more complex and requires special treatment. We will talk about this in section \ref{massless}.
\subsection{Comparison with the slab propagator}
\label{secslab}
We will now study a completely different problem which however turns out to be an imaginary time formulation of a quantum quench. We consider a euclidean free field theory defined on a $(d+1)$-dimensional slab of thickness $L$ with Dirichlet boundary conditions, that is the two-point correlation function or Green's function vanishes when one of the points are on the boundaries of the slab $\tau=-L/2$ and $\tau=+L/2$, where $\tau$ is the transverse coordinate.
The Green's function $G_{sl}({r_1},\tau_1,{r_2},\tau_2)$ for this problem can be found using the method of images as follows: to reproduce the boundary conditions we put an infinite set of alternating positive and negative `charges' at the reflections of the `source' on the boundaries (Fig.~\ref{fig:slab}.a).
\begin{figure}[htbp]
\centering
\includegraphics[width=.70\columnwidth]{slab.eps}
\caption{Images required for the slab with Dirichlet (a) or periodic (b) boundary conditions.}
\label{fig:slab}
\end{figure}
Then $G_{sl}({r}_1,\tau_1,{r}_2,\tau_2)$ is the superposition of (euclidean) Feynman propagators between $({r}_2,\tau_2)$ and each of the images of the source $({r}_1,\tau_1)$. Since the problem is translationally invariant in the $d$ longitudinal directions, in the mixed $({k},\tau)$ representation we find that $G_{sl}(\tau_1,\tau_2;k)$ is
\ee{
& \frac{1}{2 \omega_k} \left( {\sum\limits_{n=0}^\infty{e^{-\omega_k(|\tau_1-\tau_2|+2nL)}} + \sum\limits_{n=1}^\infty{e^{-\omega_k(-|\tau_1-\tau_2|+2nL)}} }\right.- \nonumber \\
& \left.{- \sum\limits_{n=0}^\infty{e^{-\omega_k\left (\tau_1+\tau_2+(2n+1)L\right )}} - \sum\limits_{n=1}^\infty{e^{-\omega_k\left (-\tau_1-\tau_2+(2n-1)L\right )}}} \right)
}
This is a geometric series and the result is
\eem[slab-prop]{
& \frac{{e^{-\omega_k |\tau_1-\tau_2|} + e^{+\omega_k (|\tau_1-\tau_2|-2L)} - 2 e^{-\omega_k L}\cosh \omega_k (\tau_1+\tau_2)}}{2 \omega_k (1-e^{-2 \omega_k L})} \nonumber \\
& = \frac{\cosh \omega_k (\tau_1-\tau_2) }{\omega_k (e^{2\omega_k L}-1)} - \frac{e^{\omega_k L} \cosh \omega_k (\tau_1+\tau_2)}{\omega_k (e^{2\omega_k L}-1)} + \nonumber \\
& + \frac{1}{2\omega_k} e^{-\omega_k|\tau_1-\tau_2|}
}
By analytically continuing to real times $\tau \to i t$ we find
\begin{multline}
\label{slab-prop2}
G_{sl}(t_1,t_2;k) = \\
= \frac{\cos \omega_k (t_1-t_2) }{\omega_k (e^{2\omega_k L}-1)} - \frac{e^{\omega_k L} \cos \omega_k (t_1+t_2)}{\omega_k (e^{2\omega_k L}-1)} + \frac{1}{2\omega_k} e^{-i\omega_k|t_1-t_2|}
\end{multline}
If we now compare the slab propagator in real time (\ref{slab-prop2}) with the quench propagator (\ref{q-prop}) we notice that these are exactly equal if and only if
\begin{subequations}
\label{conditions}
\begin{align}
\frac{(\omega_{0k}-\omega_k)^2}{4 \omega_k \omega_{0k}} & = \frac{ 1}{e^{2 \omega_k L}-1} \\
\frac{\omega_{0k}^2-\omega_k^2}{4 \omega_k \omega_{0k}} & = \frac{ e^{\omega_k L}}{e^{2 \omega_k L}-1}
\end{align}
\end{subequations}
Remarkably, the above two conditions are consistent and the solution is
\ee{\tanh{(\omega_k L/2)} =
\begin{cases}
{\omega_k}/{\omega_{0k}} \qquad \text{ if } \omega_k<\omega_{0k}, \\
{\omega_{0k}}/{\omega_{k}} \qquad \text{ if } \omega_k>\omega_{0k}.
\end{cases}}
Notice that if we solve with respect to $L$, the answer is a function of $k$.
Thus the problem of a quantum quench can be equivalently formulated as a euclidean theory on a slab with momentum dependent thickness. The initial conditions in real time are translated into boundary conditions on the slab. In the deep quench limit $m_0\to\infty$ the condition becomes $L \sim 2/m_0 $, independent of $k$ and therefore the analogy between the quantum quench and the slab is asymptotically exact. The reason is that Dirichlet boundary conditions correspond to vanishing initial value of the quench propagator, which is indeed the case for $m_0\to \infty$, since $C_q(0,0;k)=1/2 \omega_{0k} \to 0$.
It should be mentioned that our choice of Dirichlet boundary conditions has nothing special: in fact it is only important in the deep quench limit. One can verify that the quench propagator can be similarly identified with the Green's function corresponding to the following general boundary conditions (known as Robin or `impedance' boundary conditions due to their application to electromagnetics)
\eq{a {\hat{G}_{sl}(\tau_1,\tau_2;k)} + b \frac{\partial {\hat{G}_{sl}(\tau_1,\tau_2;k)}}{\partial n} = 0 }
where
\ee{& a = \omega_k \sinh(\omega_k/\omega_{0k}) - \omega_{0k} \cosh(\omega_k/\omega_{0k}) \nonumber \\
& b = \cosh(\omega_k/\omega_{0k}) - \omega_{0k}/\omega_k \sinh(\omega_k/\omega_{0k})}
$\partial /\partial n$ denotes the normal derivative at the boundary $\tau = \pm L/2$ and in this case $L$ is chosen to be $L=2/\omega_{0k}$. Note that for $\omega_{0k} \gg \omega_k$ the latter condition reduces to Dirichlet type. To intuitively understand the meaning of these boundary conditions, we can use an analogy from electromagnetics. There, Dirichlet boundary conditions correspond to complete reflection by a perfect conductor, while Robin boundary conditions correspond to partial reflection and refraction by an imperfect conductor with a large refractive index.
The correspondence between a quantum quench and the slab construction turns out to be valid, at least in the deep quench limit, even in interacting theories where an exact solution may not be possible\cite{cc-07}.
\subsection{Comparison with the thermal propagator}
\label{sec:compth}
Let us now compare the above two propagators with the thermal or \emph{Matsubara} propagator, which describes a system at thermal equilibrium at finite (inverse) temperature $\beta$. As is well-known, in imaginary time this corresponds to the Green's function in the geometry of a $(d+1)$-dimensional cylinder of circumference $\beta$, i.e. a slab of equal thickness with periodic instead of Dirichlet boundary conditions. Among other ways, this can also be derived using the method of images. To reproduce the periodic boundary conditions we now need to put only the positive images (Fig.~\ref{fig:slab}.b) and the result is
\eqq[ther-prop]{ G_{th}(\tau_1,\tau_2;k)= \frac{1}{2 \omega_k} \left({e^{-\omega_k |\tau_1-\tau_2|} + 2 \frac{\cosh \omega_k (\tau_1-\tau_2)}{e^{\beta \omega_k}-1}}\right)}
or in real time, after the analytical continuation $\tau\to i t$
\eqq[ther-prop2]{ G_{th}(t_1,t_2;k)= \frac{1}{2 \omega_k} \left({e^{-i\omega_k |t_1-t_2|} + 2 \frac{\cos \omega_k (t_1-t_2)}{e^{\beta \omega_k}-1}}\right)}
We observe that if we could ignore the $(t_1+t_2)$-dependent part then the slab propagator $G_{sl}(t_1,t_2;k)$ and the quench propagator $C_q(t_1,t_2;k)$ would be the same as the thermal propagator $G_{th}(t_1,t_2;k)$ with $L=\beta/2$. This can actually be correct for the real space form of the quench propagator at large times, as we have already seen in section \ref{secpropprop}. Indeed this is the case if $m\neq 0$ or $m=0$ and $d=3$.
As a conclusion, at large times the system tends to a state with thermal-like correlation functions, which is what we named \emph{effective thermalization}. The effective temperature is given, according to all the above, by the condition
\ee{\tanh{(\beta_{\text{eff}} \omega_k /4)} =
\begin{cases}
{\omega_k}/{\omega_{0k}} \qquad \text{ if } \omega_k<\omega_{0k}, \\
{\omega_{0k}}/{\omega_{k}} \qquad \text{ if } \omega_k>\omega_{0k}.
\end{cases}}
Notice that the effective temperature is momentum dependent, which could be expected since, as we already mentioned, in a free system each momentum mode evolves independently from the others and there is no reason why they should all thermalize to the same temperature. Yet in the deep quench limit the effective temperature becomes momentum independent $\beta_{\text{eff}}\sim 4/m_0$.
It should be emphasized that the state itself is neither thermal nor stationary: the density operator still exhibits oscillating behaviour for example. However, since in a free system all local observables can be derived from the two-point correlation function which does become stationary, the same happens to all such observables as well. It is crucial that the system is in the \emph{thermodynamic limit} and the observables under consideration are \emph{local} since then an integration over an infinite set of momenta is required and it is exactly this interference of all independent momentum modes that leads to thermalization. Such observables include those defined on any finite subsystem $A$ of the whole system, like the reduced density operator of $A$ \cite{p-03}. In this sense the complement of $A$ acts as a thermal bath with which $A$ comes into thermal equilibrium. This explains why the effective thermalization that we consider does not contradict with the fact that in a free or more generally integrable system, there is an infinite set of conserved quantities that prevent the system from thermalizing as a whole. The subsystem $A$ is not closed and there are no such restrictions to prevent its thermalization.
\subsection{Estimation of the effective temperature from the field fluctuations}
\label{avetemp}
As we saw, the effective temperature in our free model is different for each momentum mode. Since the low momentum modes are those that determine the large distance behaviour, for most purposes $\beta_{\text{eff}}(k=0)$ is sufficient in order to macroscopically describe the system. We can define \cite{scc-09}
however an estimate of the effective temperature that averages over all momentum modes in a natural way, by comparing the field fluctuations long after the quench $\langle \phi^2(x=0,{t\to\infty}) \rangle$ with those of a system at thermal equilibrium. We can call this \emph{average effective temperature} and denote it as $\bar{\beta}$. Then $\bar{\beta}$ must satisfy
\eqq[avtemp]{\int{d^dk \; {C}^*_{q}(k;m,m_0)} = \int{d^dk \; G_{th}(k;m,\bar\beta)}}
where ${C}^*_q$ stands for the stationary part of the quench propagator. More explicitly
\eq{\int_0^\infty{k^{d-1} dk \frac{(\omega_{0k}-\omega_k)^2}{4 \omega_{0k} \omega_k^2}} =
\int_0^\infty{k^{d-1} dk \frac{1}{\omega_k (e^{\bar\beta \omega_k}-1)}} }
from which we can find $\bar{\beta}$ as a function of $m$ and $m_0$.
The latter can be written in dimensionless form as
\eq{m_0^{d-1} f_d(m/m_0) = m^{d-1} g_d(\bar\beta m) }
where
\eqq[int1]{f_d(s) = \int\limits_0^\infty{k^{d-1} dk \; \frac{(\sqrt{k^2+1}-\sqrt{k^2+s^2})^2}{4 \sqrt{k^2+1} ({k^2+s^2})}}}
and
\eqq[int2]{g_d(s) = \int\limits_0^\infty{k^{d-1} dk \; \frac{1}{\sqrt{k^2+1} (e^{s \sqrt{k^2+1}} - 1)}}}
In units of $m_0$, setting $x=m/m_0$ and $y=\bar\beta m_0$ we have
\eq{{x^{d-1} = \frac{f_d(x)}{g_d(xy)}}}
Tables \ref{tab1} and \ref{tab2} show the asymptotic behaviour of the integrals in several limits for the relation between the parameters and when possible their exact form.
\begin{table*}[htbp]
\caption{Asymptotic behaviour of the quench integral $f_d(s)$~(\ref{int1})}
\label{tab1}
\begin{center}
\begin{tabular}{|c|l|l|l|l|}
\hline
$d$ & exact & $s\approx 0$ & $s\approx 1$ & $s\to \infty$ \\
\hline
1 & $ [ 2 \log s +({\sqrt{1-s^2}}/{s}) \, \arccos s ]/4 $ & $ ({\pi}/{2 s} + 2 \log s )/4$ & $(s-1)^2/6$ & $(\log s)/4$ \\
2 & $ [ 2 (s-1)- \sqrt{s^2-1} \, \arccos (1/s) ]/4 $ & $ -(\log s)/4 $ & $(s-1)^2/12$ & $(1+{\pi}/{4})s/2$ \\
3 & $ [ {(1-s^2)}/{2}-s^2 \log s- s \sqrt{1-s^2} \, \arccos s ]/4 $ & $(1-\pi s)/8$ & $(s-1)^2/12$ & $(\log 2 - {1}/{2}) s^2/4$ \\
\hline
\end{tabular}
\end{center}
\end{table*}
\begin{table*}[htbp]
\caption{Asymptotic behaviour of the thermal integral $g_d(s)$~(\ref{int2})}
\label{tab2}
\begin{center}
\begin{tabular}{|c|c|l|l|}
\hline
$d$ & exact & $s\approx 0$ & $s\to \infty$\\
\hline
1 & -- & $ ({\pi}/{2 s}) + (\log s)/2 $ & $e^{-s}\sqrt{\pi/2s} $ \\
2 & $-\log(1-e^{-s}){/s}$ & $-{(\log s)/s} $ & $ {e^{-s}/s}$ \\
3 & -- & ${({\pi^2}/{6 s^2})[1-({3s}/{\pi})]}$ & ${e^{-s}\sqrt{\pi/2} s^{-3/2}}$\\
\hline
\end{tabular}
\end{center}
\end{table*}
Figure \ref{fig:1} shows a plot of $\bar\beta$ as a function of $m$ in units of $m_0$ as obtained numerically from the above equation. Note that for $m = m_0$ i.e. no quench at all, the effective temperature $1/\bar\beta$ is zero as it should be. Apparently in the deep quench limit the small wavelength behaviour dominates and according to an earlier comment in section \ref{sec:compth}
we expect to find $\bar\beta \sim m_0^{-1}$ for any dimension. For small values of $m/m_0$ the asymptotic expressions of $f_d$ and $g_d$ allow us to calculate analytically the first order corrections of $\bar\beta$ as a function of $m/m_0$. In this way we find
\begin{itemize}
\item $d=1$
\eq{\bar\beta = \frac{4}{m_0}+\frac{32 \log 2 m}{\pi m_0^2} + ...}
\item $d=2$
\eq{\bar\beta = \frac{4}{m_0}\left (1 + \frac{3 \log 2 -2}{\log(m/m_0)} + ... \right )}
\item $d=3$
\ee{\bar\beta & = \frac{1}{m_0}\left({2 \pi}/{\sqrt{3}} - \pi (2-\pi/\sqrt{3}){m}/{m_0}+...\right) \nonumber \\
& \approx m_0^{-1}(3.6276-0.584967 m/m_0 + ...)}
\end{itemize}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=1.00\columnwidth]{T_eff.eps}
\caption{Effective temperature as a function of the final mass $\bar\beta m_0 = F_d (m/m_0)$ in units of the initial mass $m_0=1$. Inset: Asymptotic behaviour for small $m$. Notice the logarithmic corrections in $2d$.}
\label{fig:1}
\end{center}
\end{figure}
In $1d$ and $2d$ the $k=0$ momentum mode dominates so that the first order term is $\bar\beta = 4/m_0$. In $2d$ however the logarithmic corrections could render comparison with data difficult. In $3d$ the contribution of nonzero but small $k$ modes causes a small shift of the numerical factor from 4 to ${2 \pi}/{\sqrt{3}} \approx 3.6276$.
\section{Massless $1d$ theories}
\label{massless}
In section \ref{secpropprop} we saw that for $d=1$ and $m=0$ the propagator does not become stationary. However the situation is different when we consider a physical 1$d$ quantum system, for the following reason. A massless free theory is not physically meaningful. The infrared divergences impose the introduction of interaction counterterms of all orders in perturbation theory over the introduced coupling constant. The field renormalization finally results in the physical field defined as the exponential of the original gaussian field $\phi$ (the vertex operator). Thus in a physically meaningful 1$d$ system, interaction terms must always be present and the correlation function is given by the expectation values of vertex operators $\langle e^{i q\phi(x)} e^{-i q\phi(x')}\rangle$ for an appropriate value of the constant $q$. This can be evaluated readily using the well-known property of gaussian integrals
\eq{ \langle e^{iq\phi(x)} e^{-iq\phi(x')} \rangle = e^{-q^2\langle(\phi(x)-\phi(x'))^{2}\rangle/2} = e^{-q^2(C(0)-C(x-x'))} }
where $C(x-x')=\langle\phi(x)\phi(x')\rangle$ is the free propagator we have already found. From (\ref{1dprop}) we obtain
\eqq[cft1]{\langle e^{iq\phi(0,t)} e^{-iq\phi(r,t)} \rangle =
\begin{cases}
e^{-q^2 m_0 t/4} \qquad \text{ if } r>2 t, \\
e^{-q^2 m_0 r/8} \qquad \text{ if } r<2 t.
\end{cases}}
Thus the linearly increasing time dependence of $C(r,t)$ leads to an exponentially decaying correlation function outside of the horizon and a static form inside the horizon. Therefore thermalization also occurs in 1$d$ systems. This has been shown to be the case for any massive to massless quench on a 1$d$ bosonic system, using the mapping to the slab and the powerful methods of conformal field theory \cite{cc-05,cc-06,cc-07}.
\section{Anharmonic coupled oscillators (interacting field theory) in self-consistent approximation.} \label{anh}
Let us now consider a system of anharmonic coupled oscillators. The simplest form of a hamiltonian describing such a system is
\eqq[Hint]{H=\sum_r{\tfrac{1}{2} \pi^2 + \tfrac{1}{2} (\nabla \phi)^2 + \tfrac{1}{2} m^2 \phi^2 + \tfrac{1}{4!} \lambda \phi^4}}
In the continuum limit this corresponds to the simplest form of an interacting quantum field theory, the $\phi^4$ model. At $t=0$ we instantaneously change the mass from $m_0$ to $m$ and at the same time the coupling constant from $\lambda_0$ to $\lambda$. As before, we assume that initially the system lies in the ground state of the hamiltonian before the quench.
Such a model is non-integrable and can be solved only approximately. In this paper we will focus solely on the Hartree-Fock or self-consistent approximation. Roughly speaking in this approach we assume that the quartic interactions can be approximated by a `mean field' quadratic term with a parameter that should be calculated self-consistently. More specifically the $\phi^4$ interaction term of the hamiltonian can be substituted as follows {\cite{ch-75}}
\eqq[subHF]{\phi^4 \to {-} 3 \langle \phi^2\rangle^2 + 6 \langle \phi^2\rangle \phi^2 }
where we have taken into account that $\langle \phi \rangle = 0$ and the numerical factors are derived by Wick's theorem as the number of combinations of operator contractions.
Such a substitution is justified in the large-$N$ limit of the linear $\sigma$-model, which is a variant of the $\phi^4$ model where the field $\phi$ has $N$-components
\eqq[Hsigma]{H=\sum_{i=1}^N{\sum_r{\tfrac{1}{2} \pi_i^2 + \tfrac{1}{2} (\nabla \phi_i)^2 + \tfrac{1}{2} m^2 \phi_i^2 + \tfrac{1}{4!} \lambda (\phi^2_i)^2}}}
In the limit $N\to\infty$ with $\lambda N$ kept fixed, the Hartree-Fock approximation becomes exact.
Just by staring at (\ref{subHF}) we notice that the second term corresponds to a mass term but with a `mass' that has to be determined from the two point correlation function. The first term is just a number and does not affect the equations of motion but, as shown in appendix \ref{app:cons}, ensures the conservation of the total energy. Therefore we can define an effective mass $m_{\text{eff}}$ according to
\eqq[self1]{m_{\text{eff}}^{2} = m^2 + \frac{\lambda}{2} \sum_k{\langle\phi_k^2\rangle}}
and since the right hand side also depends on $m_{\text{eff}}$, this is in fact a self-consistency equation for $m_{\text{eff}}$. Note that the effective mass should correspond to the pole of the correlation function on the imaginary axis in the complex $k$-plane, which is what is physically measurable as the mass of the particles of the system. As a result of this approximation, our initial non-integrable problem has been effectively reduced to an integrable and in fact free one, subject to the self-consistency equation. In our out-of-equilibrium case we should keep in mind that the effective mass will be time dependent.
We will use two slightly different methods in applying this approach. The first one is a perturbative method. After introducing the Schwinger-Keldysh method which is suitable for out of equilibrium problems, we soon realize that the usual perturbative expansion does not converge and a resummation of Feynman diagrams using the Dyson equation is needed. This leads us to a simple ansatz for the asymptotic form of the two point correlation function at large times. The second method emphasizes on the time evolution of the system and is based on a direct integration of the equations of motion in their, simplified by the self-consistent approximation, version. In order to solve these equations we employ an approximate analytical and an exact numerical method. The results of both calculations are in agreement with each other and additionally they verify our earlier ansatz.
Before we start, it is worth to remind ourselves of the large-$N$ results for the ground state of our system, since in order to proceed to the out-of-equilibrium problem we will need to know more about the initial properties of the system. This will also introduce us to a discussion of the renormalization procedure and its application to the present problem.
\subsection{Divergences and renormalization}\label{sec:div}
The initial two-point correlation function of our system in the large-$N$ limit is simply that of a free system with a mass equal to its effective value
\eqq[self1b]{{{m_{\text{eff}}^{2}}_0} = m_0^2 + \frac{\lambda_0}{2} \sum_k{\langle\phi_k^2\rangle}}
The sum in the right hand side of (\ref{self1b}) represents the fluctuations of the field. In the continuum limit this corresponds to the integral
\eqq[ffluct]{\int{\frac{d^d k}{(2\pi)^d} \; \frac{1}{2 \sqrt{k^2+{m_{\text{eff}}^2}_0}} }}
which exhibits ultraviolet (UV) divergences in all dimensions. In $1d$ and $2d$ these can be absorbed completely by a mass renormalization, while in $3d$ an additional coupling constant renormalization is required \cite{cjp74}.
The mass renormalization amounts to allowing the bare mass $m_0$ to be divergent so as to compensate the divergent integral. The (finite) renormalized mass is defined by $m_{0R}^2 = m_0^2+\delta m_0^2$ where the mass counterterm $\delta m_0^2$ is
\eqq[dm]{\delta m_0^2 = \frac{\lambda_0}{2} \int{\frac{d^d k}{(2\pi)^d} \; \frac{1}{2 \sqrt{k^2+m_{0R}^2}} }}
The effective mass in terms of $m_{0R}$ is then
\eem[self1c]{ & {m_{\text{eff}}^{2}}_0 = m_{0R}^2 + \nonumber \\
& + \frac{\lambda_0}{2} \int{\frac{d^d k}{(2 \pi)^d} \left ( \frac{1}{2 \sqrt{k^2 + {m_{\text{eff}}^2}_0}} - \frac{1}{2 \sqrt{k^2 + m_{0R}^2}} \right ) } }
which is finite in $1d$ and $2d$.
In $3d$ there is still a logarithmic UV divergence in (\ref{self1c}) which can be absorbed by a coupling constant renormalization. A suitable renormalization counterterm can be determined by studying the 4-point correlation function and turns out to be of the form
\eqq[dl]{\delta \lambda_0 = \int{\frac{d^3 k}{(2 \pi)^3} \; \frac{1}{8 (k^2+m_{0R}^2)^{3/2}}}}
The resulting renormalized coupling constant $\lambda_{0R}$ satisfies
\eqq[rencc]{\lambda_0 = \frac{\lambda_{0R}}{1-\lambda_{0R} \, \delta \lambda_0}}
and replacing in (\ref{self1c}) we obtain
\eem[self1d]{{m_{\text{eff}}^2}_0 & = m_{0R}^2 + \frac{\lambda_{0R}}{2} \int{\frac{d^3 k}{(2 \pi)^3} \; \left ( \frac{1}{2 \sqrt{k^2+{m_{\text{eff}}^2}_0}} - \right.}\nonumber \\ & {\left.-\frac{1}{2 \sqrt{k^2+m_{0R}^2}} + \frac{{m_{\text{eff}}^2}_0 - m_{0R}^2}{4 (k^2+m_{0R}^2)^{3/2}} \right )}}
which is indeed finite. Note that in all dimensions the solution to the above equations is \eq{{m_{\text{eff}}}_0 = m_{0R}}
i.e. the renormalized mass is identical to the effective mass. In what follows we should keep in mind the well-known renormalization group result that in $3d$ the critical point of this model corresponds to zero coupling constant i.e. in the continuum limit the macroscopic behaviour of the theory is effectively free.
Therefore interactions are not physically meaningful in the continuum limit. In physical systems however, the existence of a finite lattice spacing that induces a natural UV cutoff renders all momentum integrals finite and there is not such a restriction.
After the quench, the change in the mass and the coupling constant results in a change of the corresponding counterterms as well. This is required, otherwise new divergences in the equation for the effective mass are inevitably born. On the other hand, this should not be regarded as a failure of renormalization theory as the latter does not have to apply to expectation values taken in states which are not obtained by renormalised operators acting on the vacuum and our initial state is not such. In absence of a definite rule for the selection of the renormalization counterterms like the one for the ground state or thermal expectation values, several choices can be applied\cite{como-87}.
In the present work we will be using the ground state counterterms of the theory after the quench.
Since the field fluctuations right after the quench are exactly the same as before it, the equation for the effective mass right after the quench is
\eqq[inmef0]{m_{\text{eff}}^2(t\to 0^+) = m^2 +\frac{\lambda}{2} \int{\frac{d^d k}{(2 \pi)^d} \frac{1}{2 \sqrt{k^2 + m_{0R}^2}} } }
or introducing the mass renormalization
\eem[inmef1]{ & m_{\text{eff}}^2(0^+) = m_R^2 + \nonumber \\
& + \frac{\lambda}{2} \int{\frac{d^d k}{(2 \pi)^d} \left ( \frac{1}{2 \sqrt{k^2 + m_{0R}^2}} - \frac{1}{2 \sqrt{k^2 + m_{R}^2}} \right ) } }
where $m_R$ is the renormalized mass after the quench and
\eq{\delta m^2 = \frac{\lambda}{2} \int{\frac{d^d k}{(2 \pi)^d} \frac{1}{2 \sqrt{k^2 + m_R^2}}}}
is the corresponding mass counterterm. As before the last expression is convergent in $1d$ and $2d$, but not in $3d$. If we use a coupling constant renormalization counterterm
\eqq[dl2]{\delta \lambda = \int{\frac{d^3 k}{(2 \pi)^3} \; \frac{1}{8 (k^2+m_{R}^2)^{3/2}}}}
we find
\eem[mef3d]{m_{\text{eff}}^2(& 0^+) = m_{R}^2 + \frac{\lambda_{R}}{2} \int{\frac{d^3 k}{(2 \pi)^3} \; \left ( \frac{1}{2 \sqrt{k^2+m_{0R}^2}} - \right.}\nonumber \\
& {\left.-\frac{1}{2 \sqrt{k^2+m_{R}^2}} + \frac{m_{\text{eff}}^2(0^+) - m_{R}^2}{4 (k^2+m_{R}^2)^{3/2}} \right )}}
which is only convergent in the trivial case ${m_{\text{eff}}(0^+) = m_{0R}}$ where there is no jump in the effective mass i.e. no quench at all.
Recalling our previous remark, we realise that this problem is due to the fact that the presence of interactions does not make sense in the continuum limit. In lattice systems however there is not such a problem and the practical meaning of the above is simply that $m_{\text{eff}}(0^+)$ is very large. In the following we will therefore keep a large UV cutoff $\Lambda$ in all expressions for the $3d$ case and investigate the dependence of our results on this.
An interesting first observation is that as defined by (\ref{inmef1}) the initial mass-square $m_{\text{eff}}^2(0^+)$ can be negative. Indeed for $m_0<m$, $m_{\text{eff}}^2(0^+)$ is always positive, but for $m_0>m$ the mass shift induced by the interactions is negative and if $\lambda$ is large enough then $m_{\text{eff}}^2(0^+)<0$. In particular if $m=0$ the latter is always true. From a physical point of view this negativity means that the quench can effectively drag the system into an unstable initial state like that of a double well (or generally `mexican hat') potential. We will come back to this aspect of the problem later.
The integration in (\ref{inmef1}) can be done analytically. Expressing the integral in dimensionless form we have
\eq{m_{\text{eff}}^2(0^+) = m_{R}^2 + \frac{\lambda_{R}}{2} \frac{\Omega_d}{(2\pi)^d} m_0^{d-1} h_d(m_R/m_0) }
where $\Omega_d$ is the total solid angle in $d$ dimensions ($\Omega_1 = 2, \Omega_2 = 2\pi, \Omega_3 = 4\pi$) and the function $h_d(s)$ is defined as
\eqq[hdef]{h_d(s) = \int\limits_0^{\Lambda\to\infty}{k^{d-1} dk \; \left ( \frac{1}{2 \sqrt{k^2+1}} - \frac{1}{2 \sqrt{k^2+s^2}}\right )}}
and can be easily shown to be
\eqq[hdef2]{ h_d(s) =
\begin{cases}
(\log s)/2 \quad &\text{ if } d=1, \\
(s-1)/2 \quad &\text{ if } d=2, \\
(s^2-1) (\log \Lambda )/4 \quad &\text{ if } d=3.
\end{cases} }
\subsection{Perturbative approach}
In order to study the evolution of the effective mass for $t>0$ we have to calculate the two-point correlation function and the usual way to do this is to use perturbation theory. The two-point correlation function is
\eq{ \tilde{C}(r,t_1,t_2) \equiv \left\langle {\Psi _0 } \right|\phi (0,t_1 )\phi (r,t_2 )\left| {\Psi _0 } \right\rangle }
where, as before, $\left| {\Psi _0 } \right\rangle$ is the ground state of the initial hamiltonian. For simplicity let us first assume that $\lambda_0 =0$, i.e. that there is no interaction before the quench so that $\left| {\Psi _0 } \right\rangle$ is the ground state of a free hamiltonian.
At this point we encounter an important difference with the usual QFT methods:
at zero temperature the starting point of such a calculation is usually the following formula
\eqq[qftcf]{
\frac{{\left\langle 0 \right|\mathcal{T}\{ \phi_i (0,t_1 )\phi_i (r,t_2 )\exp{\textstyle\left ( - i\int_{ - \infty }^{ + \infty } {dt\, H_{int}(t)} \right )}\} \left| 0 \right\rangle }}{{\left\langle 0 \right|\mathcal{T}\{ \exp{\textstyle\left ( - i\int_{ - \infty }^{ + \infty } {dt\, H_{int}(t)} \right )}\} \left| 0 \right\rangle }}}
But in our case this expression is inappropriate since it relies on the condition that $\left| 0 \right\rangle $ is the ground state of the free part of the hamiltonian and the interactions are swithed on and off adiabatically. In a quantum quench this is not valid because $\left| {\Psi _0 } \right\rangle $ is the ground state of a different hamiltonian and the changes are done instantaneously. Thus we have to trace back to the origin of (\ref{qftcf}) which follows from the interaction picture formalism
\eqq[Keldysh]{\left\langle {\Psi _0 } \right|\mathcal{T}\{ {\phi_i (0,t_1 ) \phi_i (r,t_2 )\exp{\textstyle\left ( - i \int_{\mathcal{K}} {dt \,H_{int}(t)}\right )}} \}\left| {\Psi _0 } \right\rangle }
where $t$ is integrated over a contour $\mathcal{K}$ that starts from some initial time $t_i$, passes through $t_1$ and $t_2$ where the interaction picture field operators $\phi_i$ are placed, extends to some final time $t_f$ and then goes back to $t_i$ so that times on the second half of the contour are considered to be later than those on the first half (Fig.~\ref{fig:Keldysh}).
\begin{figure}[htbp]
\centering
\includegraphics[width=1.00\columnwidth]{Keldysh.eps}
\caption{The Schwinger-Keldysh contour for a quantum quench.}
\label{fig:Keldysh}
\end{figure}
This is the well-known Schwinger-Keldysh method for non-equilibrium quantum systems \cite{schw-61,keld-64,keld-65,chsu-85,calhu-88,kam-04,kam-09} and is applicable to any choice of initial state.
If $\left| {\Psi _0 } \right\rangle = \left| 0 \right\rangle $ and the interaction is switched on and off adiabatically then we can extend $t_i \to -\infty$ and $t_f \to +\infty$. In this case, from the adiabatic theorem, the action of $\exp{\textstyle ( + i\int_{ - \infty }^{ + \infty } {dt\, H_{int}(t)} )}$ (i.e. the evolution operator along the second half of the contour) on $\left| {\Psi _0 } \right\rangle$ yields just a multiplicative constant and (\ref{Keldysh}) reduces to (\ref{qftcf}). In the present problem we need to use the original expression (\ref{Keldysh}) instead. The initial time can be set to be $t_i=0$, when the interaction is switched on. However the same choice can be used even in the general case when $\lambda_0 \neq 0$, i.e. when the interaction is present before the quench, since as explained above, in our approximation the initial state is still that of a free theory but with the mass replaced by its effective value.
It is worth to remark that an alternative way of deriving the Keldysh contour is by using the slab construction mentioned earlier. In this approach one would have to integrate in imaginary time from $-L/2$ to $+L/2$, i.e. from one to the other boundary of the slab, then analytically continue the arguments of the operators from imaginary to real times as in $\tau \to i t$ and finally take the limit $L \to 0$ thus recovering (\ref{Keldysh}).
We can now expand (\ref{Keldysh}) in powers of $\lambda$. According to the above, the zeroth order perturbative term $\left\langle {\Psi _0 } \right|\phi_i (0,t_1 )\phi_i (r,t_2 )\left| {\Psi _0 } \right\rangle $ is exactly the quench propagator (\ref{q-prop2}) with the masses $m_0$ and $m$ replaced by their renormalized values.
The first order correction $C^{(1)}(t_1,t_2;k)$ corresponds to the single loop Feynman diagram (Fig.~\ref{fig:loop}).
\begin{figure}[htbp]
\centering
\includegraphics[width=0.50\columnwidth]{one-loop.eps}
\caption{First order Feynman diagram.}
\label{fig:loop}
\end{figure}
After applying Wick's theorem we find that $C^{(1)}(t_1,t_2;k)$ reads
\eqq[cf0]{{\frac{\lambda}{2}}\int_{\mathcal{K}}{dt' \; C(t_1,t';k)C(t_2,t';k)\int{\frac{d^d k'}{(2\pi)^d} \,C(t',t';k') }}}
The explicit form of the loop momentum integral $\int{d^d k \; C(t,t;k)}$ is
\eqq[lmi]{ \int{d^d k \; \left(\frac{{(\omega_{0k} - \omega_k)^2 }}{{4\omega_k^2 \omega_{0k} }} - \frac{{m_0 ^2 - m^2}}{{4\omega_k^2 \omega_{0k} }}\cos 2\omega_k t + \frac{1}{2 \omega_{k}} \right)}}
but we also have to take into account the mass renormalization, which amounts to subtracting the UV divergent Feynman part and substituting the bare mass $m$ by the renormalized $m_R$
\eq{\int{d^d k \; C(t,t;k,m)} = \int{d^d k \; \left (C(t,t;k,m_R) - \frac{1}{2 \omega_{k,m_R}}\right ) }}
For brevity we redefine $m,m_0$ to be the renormalized masses $m_R,m_{0R}$ in all subsequent equations. Then (\ref{lmi}) can be written explicitly as
\eqq[expr1]{\int{d^d k \; \left(\frac{{(\omega_{0k} - \omega_{k})^2 }}{{4\omega_{k}^2 \omega_{0k} }} - \frac{{m_0 ^2 - m^2}}{{4\omega_{k}^2 \omega_{0k} }}\cos 2\omega_{k} t\right)}}
From the terms that remain in (\ref{expr1}), the first one which is the time independent part is always convergent since it decays like $k^{-5}$ for large $k$. On the other hand, the second term which is the time dependent part decays like $k^{-3} \cos 2\omega_k t$, which means that it converges in $1d$ and $2d$, while in $3d$ it is divergent only at $t=0$.
We also note that (\ref{expr1}) does not suffer from infrared (IR) divergences in the massless case $m=0$ except in $1d$.
Having analyzed the convergence of the loop integral, let us now calculate it. If we assume that $m\neq 0$, then the time-independent part has been calculated exactly already in section \ref{avetemp}: it is equal (up to a numerical factor involving the total solid angle in $d$ dimensions) to $m_0^{d-1} f_d(m/m_0)$ where $f_d(s)$ is given in Table~\ref{tab1}. On the other hand we recall that the time-dependent part has been shown to decrease with time. More specifically using the stationary phase method we find that for large times it decays like
\eq{\frac{(m^2-m_0^2)m^{d-2}}{m_0}\frac{\cos(2 m t + \varphi)}{(m t)^{d/2}}}
However for small times, this same time-dependent part can be important (or even divergent as we saw that happens in $3d$).
Thus we are naturally led to the question whether it is safe or not to completely ignore the time-dependent part of the loop integral in calculating $C^{(1)}$ for large times. If this is correct, then the effect of the loop diagram for large times is simply a mass shift equal to the time-independent part (recall that a mass renormalization counterterm induces a similar shift in the mass, but an infinite one). Higher orders in perturbation theory correspond to more loops and therefore one needs to employ a resummation of all orders in order to compute the actual mass shift. Such a resummation can lead to a non-perturbative dependence of the mass shift and the correlation function on the coupling constant. Indeed if we calculate $C^{(1)}$ from (\ref{cf0}) assuming that the loop integral is constant, then we find that the first order correction increases linearly with time, i.e. it will eventually become larger than the zeroth order term and therefore the perturbative series does not converge. The required resummation can be done using the Dyson equation as described in the next section.
\subsubsection{Resummation using the Dyson equation}
As well-known the Dyson equation is an integral equation satisfied by the two-point correlation function of an interacting theory that expresses the fact that the latter can be constructed from the propagator in a recursive fashion, using a number of `skeleton' diagrams as building blocks. In our problem and in the mixed representation the Dyson equation can be written in the form
\eem[Dyson0]{& \tilde{C}(t_1,t_2;k)= \nonumber \\
& C(t_1,t_2;k)+\int_{\mathcal{K}}{dt' \int_{\mathcal{K}}{dt'' \; C(t_1,t';k) \Sigma(t',t'';k) \tilde{C}(t'',t_2;k)}}}
where we denote the full correlation function by $\tilde{C}$ and $\Sigma$ is the self-energy insertion, i.e. a two-leg insertion also constructed recursively by the skeleton diagrams.
In the large-$N$ limit the loop diagram is the only skeleton diagram and therefore the self-energy is simply a loop of the full correlation function. The Dyson equation then takes the simplified form (Fig.~\ref{fig:Dyson})
\eem[Dyson]{& \tilde{C}(t_1,t_2;k)= \nonumber \\
& C(t_1,t_2;k)+\int_{\mathcal{K}}{dt'\; C(t_1,t';k) \Sigma(t') \tilde{C}(t',t_2;k)}}
where
\eqq[Dyson2]{\Sigma(t') = \frac{\lambda}{2} \int{\frac{d^d k'}{(2\pi)^d} \left (\tilde{C}(t',t';k') - \frac{1}{2 \omega_k}\right )}}
taking into account the mass renormalization. Notice that comparing with (\ref{self1}) we realize that $\Sigma(t)$ is nothing but the shift in the mass-square
\eqq[Dyson3]{\Sigma(t) = m_{\text{eff}}^2(t) - m^2}
\begin{figure}[htbp]
\centering
\includegraphics[width=1.0\columnwidth]{Dyson.eps}
\caption{Diagrammatic representation of the Dyson equation in the large-$N$ limit.}
\label{fig:Dyson}
\end{figure}
As we see the Dyson equation contains $\tilde{C}$ explicitly in the right hand side but also implicitly in the definition of $\Sigma$. Thus it is difficult to solve in general. In many cases it is useful as a check of validity for an ansatz: we assume a particular form for $\tilde{C}$, substitute in the Dyson equation and check the consistency or determine any free parameters. This is how we are going to use it in our problem.
Let us therefore construct an ansatz based on the hypothesis that the time-dependence of the loop be negligible. Then the same can be assumed for the self-energy since this is nothing but a dressed loop, i.e. the sum of all `cactus-diagrams'. This would mean that $\Sigma(t)$ can be replaced by its large time stationary value $\Sigma^* = \lim_{t\to +\infty} \Sigma(t)$ or, according to (\ref{Dyson3}), that the effective mass itself can be considered as time-independent and equal to its large time stationary value $m^*= \lim_{t\to +\infty} m_{\text{eff}}(t)$. In other words we suppose that the effective mass simply jumps at the time of the quench from $m_{0}$ to $m^*$ in which case the correlation function should simply be equal to the quench propagator for a quench from $m_{0}$ to $m^*$. Note that our assumption is twofold: first we assume that $m_{\text{eff}}$ tends to a stationary value and second that this happens fast enough to approximate its evolution by a jump.
According to the above, our ansatz is that the two point correlation function $\tilde{C}(t_1,t_2;k)$ is approximately the same as the propagator itself but with $m$ replaced by an asymptotic effective mass $m^*$
\eqq[ans]{\tilde{C}(t_1,t_2;k;m_0,m) \sim C(t_1,t_2;k;m_0,m^*)}
We expect this relation to be asymptotically exact for large times, when any memory of the initial evolution of the effective mass will have been lost.
Now that we have an ansatz for the correlation function we can use the Dyson equation to check its validity and determine the value of the free parameter $m^*$. By substituting (\ref{ans}) into (\ref{Dyson}) and (\ref{Dyson2}) and replacing $\Sigma(t)$ by $\Sigma^*$, we find that the Dyson equation is satisfied exactly when $m^{*}$ satisfies the self-consistency equation
\eqq[self-cons0]{m^{*2}-m^2 = \Sigma^* = \frac{\lambda}{2} \int{\frac{d^d k}{(2\pi)^d} \left ({C}^*(k;m_0,m^*) - \frac{1}{2 \omega_k}\right )} }
where ${C}^*(k;m_0,m)$ is the stationary part of the propagator.
One can also check whether the remainder of the large time asymptotic form of the Dyson equation
\eq{\int_{\mathcal{K}}{dt'\, C(t_1,t';k;m_0,m) (\Sigma(t')-\Sigma^*) C(t',t_2;k;m_0,m^*)} }
with
\eq{\Sigma(t') = \frac{\lambda}{2} \int{\frac{d^d k}{(2\pi)^d} \left (C(t',t';k;m_0,m^*) - \frac{1}{2 \omega_k}\right )}}
tends to zero as supposed to. This is however a cumbersome calculation and will not be presented. We will later show an alternative way to study the time evolution and verify our ansatz, but for the moment let us focus on the self-consistency equation (\ref{self-cons0}) and investigate its solutions.
\subsubsection{Self-consistent calculation of the mass shift}
Written explicitly the self-consistency equation (\ref{self-cons0}) is
\eqq[self-con]{{m^*}^2 = m^2 + \frac{\lambda}{2} \int{\frac{d^d k}{(2 \pi)^d} \; \left (\frac{(\omega_{0k} - \omega^*_{k})^2}{4 \omega_{0k} {\omega_k^*}^2} + \frac{\omega_{k} - \omega^*_{k}}{2 \omega_k \omega_{k}^{*}} \right )}}
where $\omega_k^* = \sqrt{k^2 + {m^*}^2}$.
Once again some comments about the 3$d$ case are due as (\ref{self-con}) contains a logarithmically divergent integral. Therefore a UV cutoff $\Lambda$ is assumed and the solutions $m^*$ will depend upon it. As can be verified however, the small $\lambda$ behaviour of $m^*$ is not affected by $\Lambda$. By the way the $\lambda$-counterterm (\ref{dl2}) would successfully remove the current divergence yielding the finite equation
\eem[self-con-3d]{{m^*}^2 = m^2 + & \frac{\lambda_R}{2} \int{\frac{d^3 k}{(2 \pi)^3} \; \left( \frac{(\omega_{0k} - \omega^*_{k})^2}{4 \omega_{0k} {\omega_k^*}^2} + \right. } \nonumber \\
& \left. + \frac{\omega_{k} - \omega^*_{k}}{2 \omega_k \omega_{k}^{*}} + \frac{{m^*}^2 - m^2}{4 \omega_k^3} \right) }
but according to the discussion in section \ref{sec:div}, this is supposed to be correct only for $\lambda \to 0$ and therefore provides no more information than (\ref{self-con}) with a cutoff.
Going back to the general case, if we make the momentum integrals dimensionless then the self-consistency equation can be written as
\eq{{m^*}^2 = m^2 + \frac{\lambda}{2} \frac{\Omega_d}{(2\pi)^d} \left [ m_0^{d-1} f_d\left (\frac{m^*}{m_0}\right ) + {m^*}^{d-1} h_d\left (\frac{m}{m^*}\right ) \right ]}
where $f_d(s)$ and $h_d(s)$ are the previously defined functions (\ref{int1}) and (\ref{hdef}).
The above equations can be solved numerically or even analytically in several asymptotic limits like for $\lambda\to 0$ or $m \to 0$.
Fig.~\ref{fig:2},\ref{fig:3} and \ref{fig:4} show plots of the solutions $m^*$ as a function of $\lambda$ for several values of $m$ in $1d$, $2d$ and $3d$, while Fig.~\ref{fig:3bs} shows $m^*$ as a function of $m$ for $\lambda\to\infty$ in $1d$ and $2d$. A first important remark is that for $m\neq 0$ and small $\lambda$ the first order correction $m^*-m$ is linear in $\lambda$, while for $m = 0$ this is not true. Instead $m^*$ depends on $\lambda$ in a non-perturbative way in this case. The first order corrections in $\lambda$ for $m=0$ are summarized below:
\begin{itemize}
\item for $d=1$
\eq{m^* = 0 \quad \text{ for all } \quad \lambda}
In fact it is more correct to talk about the limit $m \to 0$, since $m$ can never reach zero in $1d$. In this limit, $m^*$ follows $m$ to zero like
\eq{m^* \sim \frac{m_0\pi/2}{2 \log (m_0/m) + 1 - 16 \pi^2 m^2/\lambda m_0^2}}
\item for $d=2$
\eq{m^* = \frac{1}{4}\sqrt{ \frac{\lambda m_0}{2\pi} \log(m_0/\lambda)}}
\item for $d=3$
\eq{m^* = \frac{m_0}{4 \pi \sqrt{2}} \lambda^{1/2} }
independent of the cutoff.
\end{itemize}
\begin{figure*}[htbp]
\begin{center}
\includegraphics[width=1.8\columnwidth]{m_eff_1d.eps}
\caption{Solutions of the self-consistency equation (\ref{self-con}) in $1d$. The plots show the effective mass $m^*$ as a function of the coupling constant $\lambda$ for several values of $m$ in units of $m_0=1$. Notice that as $m\to 0$ the effective mass tends logarithmically to zero for all $\lambda$.}
\label{fig:2}
\end{center}
\end{figure*}
\begin{figure*}[htbp]
\begin{center}
\includegraphics[width=1.8\columnwidth]{m_eff_2d.eps}
\caption{The same plot in $2d$. Notice that, in contrast to the $1d$ case, as $m\to 0$ the effective mass tends to a non-zero value for all $\lambda>0$.}
\label{fig:3}
\end{center}
\end{figure*}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=.8\columnwidth]{m_eff_3d.eps}
\caption{The same plot in $3d$. The curves show a weak (but increasing for increasing $\lambda$) dependence on the cutoff $\Lambda$. The dashed lines correspond to $\Lambda=10^4$ while the full ones to $\Lambda=10^7$.}
\label{fig:4}
\end{center}
\end{figure}
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=.9\columnwidth]{lambda_inf.eps}
\caption{Effective mass as a function of $m$ for $\lambda \to \infty$ in $1d$ (blue line) and $2d$ (red line) in units of $m_0=1$. The dashed straight lines are for reference. Notice that as $m\to 0$, $m^*\to 0$ logarithmically in $1d$, while in $2d$ $m^* \to 0.24954$. }
\label{fig:3bs}
\end{center}
\end{figure}
On the other hand for large $\lambda$ and $m$, $m^*$ increases like $m^*\sim m^2/2 m_0$ in $1d$, $m^*\sim 4 m/\pi$ in $2d$ while in $3d$ the large $\lambda$ result is cutoff dependent. In addition, in $2d$ and for $m=0$ and $\lambda \to \infty$ we find $m^* \to 0.24954... \, m_0$.
Of particular interest are the $2d$ results for $m=0$. The fact that $m^* \neq 0$ means that from the critical evolution in the presence of interactions, there always emerges a finite effective mass which lets the system become stationary, in contrast to the free case.
\subsection{Time evolution}
We saw in section \ref{sec:div} that the initial value of the effective mass-square $m_{\text{eff}}^2(0^+)$ can be negative, while our ansatz suggests that its asymptotic final value is always positive. It is therefore worthwhile to investigate the time evolution of the effective mass in more detail. Although this can be done in the context of perturbation theory as in the previous section, an alternative and rather simpler way is by integrating the equations of motion for the field operator $\phi$. Since the exact equations are nonlinear, even if we were able to solve them the solution would depend on the initial operators $\phi(0),\dot \phi(0)$ in a nonlinear way, thus preventing a direct application of the initial conditions (\ref{inexpval}) as done in section \ref{secFFTprop}. Fortunately in the Hartree-Fock approximation this obstacle can be circumvented since the $\phi^4$ interaction term of the hamiltonian is substituted by a quadratic `mean field' term according to (\ref{subHF}). As explained in section \ref{anh}, this substitution reduces the interacting into a free problem with a time-dependent effective mass given by
\eqq[self1a]{m_{\text{eff}}^{2}(t) = m^2 + \frac{\lambda}{2} \sum_k{\left (\langle\phi_k^2(t)\rangle-\frac{1}{2 \omega_k}\right )}}
thus yielding a linear equation of motion.
Even after this simplification however the problem is not trivial. In the following two sections we will first apply an approximate method that leads to an analytical solution for small values of the coupling constant and later derive exact equations for the evolution of the correlation function which we will integrate numerically.
\subsubsection{Quasi-adiabatic self-consistent approximation}
A common approximation that could provide a completely analytical treatment is the adiabatic approximation which is based on the assumption that $m_{\text{eff}}(t)$ varies slowly in comparison with the fast oscillations that characterize the solution \cite{calhu-88}. This is not a reasonable assumption though, since it is the solution itself that determines the time dependence of $m_{\text{eff}}(t)$. However as we show below, one can establish an alternative argument leading to the same approximate solution. The latter becomes equivalent to our earlier ansatz (\ref{ans}) for small $\lambda$ and provides a first idea of the qualitative behaviour of the solution.
Since our problem is now free, it can once again be decomposed into a set of independent harmonic oscillators. Of course the time-dependence of the frequency of each oscillator involves a summation over the whole set of them, but for the moment it suffices to consider a single quantum harmonic oscillator with an arbitrary time-dependent frequency $\omega(t)$. The hamiltonian is
\eqq[timedepH]{H=\frac{1}{2}\pi^2 + \frac{1}{2}\omega^2(t) \phi^2}
The equation of motion for the field operator evolving under $\omega(t)$ is
\eqq[eqm]{\ddot{\phi} + \omega^2(t) \phi = 0}
If the frequency varies with time very slowly (adiabatically) then $\dot{\omega}/\omega^{2} \ll 1$ and as well-known the solution is given by
\eem[eq2]{\phi(t)= & \phi(0) \sqrt{\frac{\omega(0)}{{\omega(t)}}} \cos{\left ( \int_0^t{\omega(t') dt'}\right )} + \nonumber \\
& + \pi(0) \frac{1}{\sqrt{ \omega(t) \omega(0)}} \sin{\left ( \int_0^t{\omega(t') dt'}\right )} }
A detailed derivation of the above equation in the quantum case can be found in appendix \ref{app:adiab}.
Although, as we said, the adiabaticity condition does not apply to our problem because $\omega(t)$ may exhibit oscillations with the same frequency as the solution, the condition $\dot{\omega}/\omega^{2} \ll 1$ is also valid when the amplitude of the frequency oscillations is sufficiently small in comparison with the average value. This happens when the coupling constant $\lambda$ is sufficiently small so that from (\ref{self1a}) $m_{\text{eff}}(t) \approx m$.
In this quasi-adiabatic approximation we can still use the last expression (\ref{eq2}) as the solution to our problem.
Having found the time evolution of $\phi$ we can use the initial conditions to derive the correlation function $\langle \phi^2(t) \rangle$ which is all we need in order to find $m_{\text{eff}}(t)$. Recall that from (\ref{inexpval}) we have $ \langle \phi(0)\pi(0) + \pi(0)\phi(0) \rangle = 0$ and $ \langle \phi^2(0) \rangle = 1/2\omega_0$, $ \langle \pi^2(0) \rangle = \omega_0/2$. By a direct calculation
\eqq[a1]{\langle \phi^2(t) \rangle = {\frac{\omega^2(0) + \omega^2_0}{{4 \omega_0 \omega(t) \omega(0)}}} + \frac{\omega^2(0) - \omega^2_0}{{4 \omega_0 \omega(t) \omega(0)}} \cos\textstyle{\left (2\int_0^t{\omega_k(t') dt'}\right )} }
Now going back to the interacting field theory model, we coclude that the equal time correlation function for each momentum mode $\langle\phi_k^2(t)\rangle$ is given by (\ref{a1}) with $\omega_k(t)$ corresponding to the time-dependent effective mass (\ref{self1a}) i.e. $\omega_k^2(t)=k^2+m_{\text{eff}}^{2}(t)$. Therefore the self-consistency equation for $m_{\text{eff}}(t)$ is
\eem[mt]{m_{\text{eff}}^2 & (t) = m^2 + \frac{\lambda}{2} \sum_k \left[ {\frac{\omega_k^2(0) + \omega^2_{0k}}{{4 \omega_{0k} \omega_k(t) \omega_k(0)}} + } \right. \nonumber \\
& \left. + \frac{\omega_k^2(0) - \omega^2_{0k}}{{4 \omega_{0k} \omega_k(t) \omega_k(0)}} \cos\textstyle{\left (2\int_0^t{\omega_k(t') dt'}\right )} \displaystyle - \frac{1}{2 \omega_{k}}\right]}
This equation enables us to extract physical information about the evolution of the system through its only parameter $m_{\text{eff}}(t)$. A first observation is that $m_{\text{eff}}(t)$ depends on an average value over all previous times. The initial value of the effective mass $m_{\text{eff}}^2(0^+)$ seems to be crucial for the time evolution. If $m_{\text{eff}}^2(0^+)>0$ and $\lambda\to 0$ then $m_{\text{eff}}(t)$ exhibits weak oscillations and the adiabaticity condition is satisfied for all times. At large times the argument of the $\cos$ increases like $2\bar{\omega}_k t$ where $\bar{\omega}_k$ is the time average of ${\omega_k}(t)$. Therefore we can apply the stationary phase method to show that the oscillations decay in time and $m_{\text{eff}}(t)$ indeed tends to a stationary value given by
\eqq[eq4]{{m^*_{qa}}^2 = m^2 + \frac{\lambda}{2} \sum_k \left ({\frac{\omega^2_k(0) + \omega^2_{0k}}{{4 \omega_{0k} \omega_k^* \omega_k(0)}}} -\frac{1}{2 \omega_k}\right )}
If however $m_{\text{eff}}^2(0^+)<0$, the small $k$ modes exhibit, at least at short times, exponential instead of oscillating evolution and the adiabaticity condition is no longer satisfied. The latter is also true in the marginal case $m_{\text{eff}}^2(0^+)=0$.
Although (\ref{eq4}) is not the same as the corresponding equation of our ansatz (\ref{self-con}), they are in perfect agreement for $\lambda \to 0$ where the quasi-adiabatic approximation is correct. In the next section we will see that it is possible to construct a system of differential equations that describe the time evolution of $m_{\text{eff}}(t)$ exactly, thus allowing us to investigate the large $\lambda$ regime.
\subsubsection{Exact time evolution equations and numerical solution}
Let us go back to the problem of a quantum harmonic oscillator with a time dependent frequency, described by the hamiltonian (\ref{timedepH}) and the equations of motions (\ref{eqm}) and start from scratch. Inspired by the adiabatic solution (\ref{eq2}), we assume a solution of the form\cite{como-87}
\eqq[sol]{\phi(t) \sim \frac{1}{\sqrt{2 \Omega(t)}} \exp{\textstyle\left (-i\int_0^t{\Omega(t') dt'}\right )}}
where $\Omega(t)$ is a suitable function that we wish to determine. Substituting into (\ref{eqm}) we find that (\ref{sol}) is the exact solution if $\Omega(t)$ satisfies the equation
\eqq[condnum]{\frac{\ddot \Omega}{2 \Omega} - \frac{3}{4}\left (\frac{\dot \Omega}{\Omega}\right )^2 + \Omega^2 = \omega^2(t)}
By comparison with the constant frequency case we can find that the appropriate initial conditions for $\Omega(t)$ are
\eq{\Omega(0)=\omega(0), \qquad \dot \Omega(0) =0}
Notice that if the derivatives of $\omega$ are much smaller than $\omega$ itself, we reproduce the quasi-adiabatic limit where $\Omega(t) = \omega(t)$ to first order.
Taking into account the general initial conditions for $\phi(0),\pi(0)$ we have
\eem[sol2]{\phi(t) = & \phi(0) \sqrt{\frac{\Omega (0)}{\Omega (t)}} \cos{\textstyle\left (\int_0^t{\Omega (t') dt'}\right )} + \nonumber \\
+ & \pi(0) \frac{1}{\sqrt{\Omega (t)\Omega (0)}} \sin{\textstyle\left (\int_0^t{\Omega (t') dt'}\right )}}
from which, using once again the initial conditions (\ref{inexpval}), we find that the equal time correlation function is
\eem[corf2]{ \langle \phi^2(t)\rangle & = \frac{1}{2 \Omega(t)} \left[ 1 + \frac{(\omega(0)-\omega_{0})^2}{2 \omega(0) \omega_{0}} + \right. \nonumber \\
& \left. + \frac{\omega^2(0)-\omega^2_{0}}{2 \omega(0) \omega_{0}} \cos{\textstyle\left (2\int_0^t{\Omega(t') dt'}\right )} \right]}
In fact the only difference with (\ref{a1}) is that $\omega(t)$ has been substituted with $\Omega(t)$. The overall result is that instead of (\ref{eqm}) one has to solve another differential equation (\ref{condnum}). The advantage is that the former is an operator equation while the latter is an ordinary equation and it is easier to deal with real or complex-valued functions than operators, especially since we will have to solve it numerically.
In our interacting problem, the equal time correlation function for each momentum mode $\langle\phi_k^2(t)\rangle$ will be given as before by (\ref{corf2}) where $\Omega_k(t)$ is also a function of $k$. Note that $\Omega_k(t)$ itself does not have to be of the form $(k^2 + M^2(t))^{1/2}$ but for large $k$ it is asymptotically equal to $\omega_k(t)$, which ensures that nothing has changed as long as the convergence of the integral in (\ref{self1a}) is concerned.
The system of equations (\ref{corf2}), (\ref{condnum}) and (\ref{self1a}) completely determine the time evolution of the system. Although very difficult to deal with analytically, it can be easily integrated numerically by discretizing the $(k,t)$ space and iteratively applying the following loop:
\begin{enumerate}
\item calculate $\Omega_k(t)$ for each $k$ from (\ref{condnum}),
\item calculate $\langle\phi_k^2(t)\rangle$ for each $k$ from (\ref{corf2}),
\item calculate $m_{\text{eff}}^2(t)$ from the self-consistency equation (\ref{self1a}),
\item move one step forward in time $t \to t+dt$.
\end{enumerate}
Fig.~\ref{fig:time} shows typical plots of the time evolution of the effective mass. For $m_{\text{eff}}^2(0^+)>0$ we see that the latter exhibits decaying oscillations around an asymptotic stationary value. We observe that this is the case not only for small values of $\lambda$ as we proved using the quasi-adiabatic approximation, but also for large ones. Moreover we find that even when $m_{\text{eff}}^2(0^+)<0$ in which case the quasi-adiabatic approximation fails, $m_{\text{eff}}^2(t)$ increases quickly and soon becomes positive to follow an oscillating evolution similar to the previously described one. The reason is that the exponential growth of the momentum modes with $k^2<-m_{\text{eff}}^2(t)$ leads to a fast increase of $m_{\text{eff}}^2(t)$ that brings it to positive values, ceasing the exponential growth and leaving only oscillating modes\cite{bo-93,bls-93,bdv-93}.
The asymptotic value $m^*$ as numerically estimated from the above method is systematically compared with that derived by our ansatz in the next section. It is remarkable that they are in perfect agreement for all choices of values for the parameters we studied.
\begin{figure}[htbp]
\centering
\includegraphics[width=.80\columnwidth]{time.eps}
\includegraphics[width=.80\columnwidth]{time_neg.eps}
\caption{Typical plots of the time evolution of the effective mass as obtained numerically both in $1d$. (a) The first plot corresponds to parameter values $(m_0,m,\lambda) = (1,2,10)$ that yield a positive value for $m_{\text{eff}}^2(0^+)$. The effective mass exhibits oscillations of decaying amplitude $\sim t^{-1/2}$ about an asymptotic value that is accurately predicted by our ansatz $m^*$. (b) The second plot corresponds to $(m_0,m,\lambda) = (1,0.5,10)$ that yield a negative value for $m_{\text{eff}}^2(0^+)$. The initial exponential growth brings $m_{\text{eff}}^2$ to positive values and as before $m_{\text{eff}}$ tends to the value $m^*$ found with our ansatz. The 2$d$ and 3$d$ cases are similar. }
\label{fig:time}
\end{figure}
\subsubsection{Comparison of the quasi-adiabatic and numerical results with our ansatz}
Let us recall our earlier ansatz for the correlation function $\tilde C(k,t)$ stating that the latter is the same, at large times, as that for a free theory with $m$ replaced by the final effective value $m^* = m_{\text{eff}}(t\to\infty)$ which we find self-consistently, i.e.
\eem{C_{\text{ans}}(k,t) & \sim \frac{1}{2 {\omega_k^*}} \left[ 1 + \frac{( {\omega_k^*}-\omega_{0k})^2}{2 {\omega_k^*} \omega_{0k}} + \right. \nonumber \\
& \left. + \frac{ {\omega_k^{*2}}-\omega^2_{0k}}{2 {\omega_k^*} \omega_{0k}} \cos{\left (2 {\omega_k^* t} \right )} \right]}
On the other hand the quasi-adiabatic approximation gives
\eem{C_{\text{qa}}(k,t) & = \frac{1}{2 {\omega_k(t)}} \left[ 1 + \frac{({\omega_k(0)}-\omega_{0k})^2}{2 {\omega_k(0)} \omega_{0k}} + \right. \nonumber \\
& \left. + \frac{{\omega_k^2(0)}-\omega^2_{0k}}{2 {\omega_k(0)} \omega_{0k}} \cos{\textstyle\left (2{\int_0^t{ {\omega_k(t')} dt'}}\right )} \right]}
while the exact evolution in the Hartree-Fock approximation of the problem, presented in the last section, is
\eem{C_{\text{ex}}(k,t) & = \frac{1}{2 {\Omega_k(t)}} \left[ 1 + \frac{( {\omega_k(0)}-\omega_{0k})^2}{2 {\omega_k(0)} \omega_{0k}} + \right. \nonumber \\
& \left. + \frac{ {\omega_k^2(0)}-\omega^2_{0k}}{2 {\omega_k(0)} \omega_{0k}} \cos{\textstyle\left (2 {\int_0^t{\Omega_k(t') dt'}}\right )} \right]}
The last two expressions differ only in that $\Omega_k(t)$ is replaced by $\omega_k(t)$ in $C_{\text{qa}}(k,t)$. An important difference between both last two expressions and $C_{\text{ans}}$ is that in the latter $\omega_k^*$ replaces $\omega_k(0)$. Furthermore although the argument of the $\cos$ in $C_{\text{qa}}$ should tend to $2 {\omega_k^* t}$ as in $C_{\text{ans}}$ for large $t$, this is not necessary for $C_{\text{ex}}$.
Thus $C_{\text{ans}}$ is not apparently consistent with either $C_{\text{ex}}$ or $C_{\text{qa}}$, except for $\lambda\to 0$ where all of them are in agreement.
\begin{figure}[htbp]
\centering
\includegraphics[width=.90\columnwidth]{comp_2}
\caption{Comparison of numerical data (crosses) with our ansatz (lines) for $m=0$. The plots are $\Sigma_\infty=m_{\text{eff}}^2(\infty)-m^2$ as a function of $\lambda$ in units $m_0=1$. The red line corresponds to $2d$ and the blue one to $3d$ with $\Lambda=100$.}
\label{fig:comp0}
\
\
\includegraphics[width=.90\columnwidth]{comp_1}
\caption{Comparison of numerical data (crosses) with our ansatz (solid lines) and quasi-adiabatic predictions (dashed lines) for several values of $m$ (again in units $m_0=1$). The red lines correspond to $2d$ and $m=2$, the green ones to $2d$ and $m=5$ and the blue ones to $1d$ and $m=2$. It is clear that the numerics agree with our ansatz rather than the quasi-adiabatic approximation which is only good for small values of $\lambda$.}
\label{fig:comp}
\end{figure}
Lacking an analytical argument to verify our ansatz, we rely on the numerical evaluation of the exact expression and determination of the corresponding asymptotic value of $m_{\text{eff}}$. Fig.~\ref{fig:comp0}, \ref{fig:comp} show plots of the shift $\Sigma_\infty=m_{\text{eff}}^2(\infty)-m^2$ as a function of $\lambda$ for various choices of the parameter values and dimensionality, always in units $m_0=1$. The plots are based on the predictions of our ansatz, of the quasi-adiabatic approximation and estimates drawn from numerical integration of the exact equations. By comparison we observe that the numerical data agree with our ansatz very well even for large values of $\lambda$. In the contrary they do not agree with the quasi-adiabatic results, apart from first order in $\lambda$. We conclude that, although our ansatz is not manifestly consistent in form with the exact solution, it however reproduces the exact results very successfully.
\section{Conclusions}
We studied the problem of a quantum quench in which we simultaneously change the mass and the coupling constant of an interacting system. We restrict ourselves to the time dependent Hartree-Fock approximation and make the plausible hypothesis that for large times the two-point correlation function is the same as the propagator but with a mass shift. We verify the self-consistency of our ansatz and derive the asymptotic effective mass as a function of $m, m_0$ and $\lambda$ which is shown to be correct by numerics. We point out that if $m_{\text{eff}}(t)$ approaches its final value $m_{\text{eff}}(\infty)$ sufficiently quickly then in the Hartree-Fock approximation the composite quench of the mass and the coupling constant is essentially nothing but a simple quench of the mass from $m_0$ directly to $m_{\text{eff}}(\infty)$. In this case our ansatz would be justified and its generic success is probably an indication that such a fast `relaxation' process is indeed what happens.
Our findings show that effective thermalization, one of the highlights of quantum quenches in free and $1d$ conformal systems, is also possible in interacting systems such as the present model. Furthermore it is enhanced in some sense by the presence of interactions, since it occurs under more general conditions than in free systems (that is even in $2d$ massless systems). This is because of the shift in the effective mass of the system induced by the interactions. As this is their only effect in our approximation, the effective temperature is still given by the same relation as in a free model but with $m$ replaced by $m^*$, thus depending on the coupling constant. In particular, the effective temperature is still momentum dependent as in the free case, but this should not be surprising:
as explained in the introduction and the main text, in diagrammatic perturbation theory the Hartree-Fock approximation amounts to keeping only `cactus-diagrams', i.e. Feynman diagrams that can be constructed solely by loops, and ignores the effect of collisions between quasiparticles with different momenta that can induce a mixing of the different modes. The next order correction would be to take into account the `sunset' diagram shown in Fig.~\ref{fig:sunset}.
\begin{figure}[htbp]
\centering
\includegraphics[width=0.5\columnwidth]{sunset.eps}
\caption{The `sunset' diagram.}
\label{fig:sunset}
\end{figure}
We finally mention that, except for the stationary behaviour, also the other qualitative features of the two-point correlation function that we observed in section \ref{secpropprop} for the mass quench in the free case, are general and present also in interacting models and for quenches of the interaction strength. This comment refers not only to the horizon effect for which it is obvious, but also to the characteristic oscillations\cite{gdlp-07}, either decaying or not, and is valid at least for integrable models (or even for sufficiently small deviations from integrability) since, as the present work suggests, the first effect to the quench is only a shift of the quasiparticle masses (equivalently of the poles of the scattering matrix).
\begin{acknowledgments}
This work was supported by the EPSRC grant EP/D050952/1 and the grants INSTANS (from ESF) and 2007JHLPEZ (from MIUR). S.~Sotiriadis also acknowledges financial support from St John's College, Oxford, and the
A.~G.~Leventis Foundation.
\end{acknowledgments}
|
\section{Introduction}
\label{S:Introduction}
\subsection{Overview}
\label{SS:Overview}
One of the major breakthroughs in the theory of quantum groups in the last decade was the main quantization result from \cite{ESS}, the general explicit quantization of all classical dynamical $r$-matrices which fit Schiffmann's classification \cite{Schi}. This complemented the categorical quantization results of Etingof-Kazhdan \cite{EK} and provided a fully constructive method to quantize a given $r$-matrix.
In this note we initiate an analogous program of constructing explicit quantizations
in the context of Lie superalgebras. In particular we explicitly quantize zero-weight super dynamical $r$-matrices with zero coupling constant. We next discuss the classification problem for super dynamical $R$-matrices and provide some partial answers. Then we use our results to weigh in on the question of what the correct graded analogue should be for the Hecke condition. Thus the note overall contributes to the theory of super quantum groups, which is still widely incomplete.
\subsection{Results}
\label{SS:Results}
Our main quantization result is the following theorem, proved in Section \ref{S:QuantRes}:
\begin{theorem}
\label{T:QuantRat}
Let $\mathfrak{h}$ be a finite dimensional commutative Lie superalgebra over $\mathbb{C}$ and let $V$ be a finite dimensional semisimple $\mathfrak{h}$-module whose weights make up a basis for $\mathfrak{h}^*$. Then every super dynamical $r$-matrix ${r : \mathfrak{h}^* \rightarrow \operatorname{End}(V \otimes V)}$ with zero weight and zero coupling constant, holomorphic on an open polydisc $U \subset \mathfrak{h}^*$, can be quantized to a super dynamical $R$-matrix $R$ on $U$.
\end{theorem}
\noindent
(See \S\S\ref{SS:Notation}-\ref{SS:SDYBENoParam} and \S\S\ref{SS:QDYBEBasics} for the relevant definitions).
Zero-weight super dynamical $r$-matrices with no spectral parameters were classified by the author in \cite{Kar3} in a manner which generalized the analogous non-graded results of \cite{EV}.
In the proof of the above theorem, we make extensive use of this result, as well as results from \cite{EV2}.
The quantum theme of this note is developed mostly in Section \ref{S:QuantPict}. There we briefly study the classification problem for super dynamical $R$-matrices. and then focus on the super Hecke condition. The (non-graded) Hecke condition, introduced in \cite{EV2} as a desirable property of dynamical $R$-matrices, is a quantum analogue of the generalized unitarity condition. We proposed a super version of it in our \cite{Kar4}. In Section \ref{S:Conclusion}, we use our work here and some other considerations to weigh in on this issue of the correct super version.
The extension to the graded world of the general constructive quantization \cite{ESS} of all classical dynamical $r$-matrices which fit Schiffmann's classification \cite{Schi} is still an open problem, and work on it is still ongoing \cite{GK}. Part of the difficulty comes from the fact that there is not yet a complete classification result analogous to \cite{Schi}; see \cite{Kar1, Kar2, Kar3, Kar4} for partial results and counterexamples in this direction.
\subsection{The organization of this note}
\label{SS:Organization}
This note is organized as follows: In Section \ref{S:Definitions} we provide the basic definitions and summarize the result from \cite{Kar3} that we will need.
In Section \ref{S:QuantRes}, we prove Theorem \ref{T:QuantRat}.
In Section \ref{S:QuantPict}, we sketch the development of a super analogue for the classification of super dynamical $R$-matrices given in \cite{EV2}. Section \ref{S:Conclusion} concludes the note with a discussion of the implications of our work to the problem of determining the correct way to superize the Hecke condition.
\section{Definitions and relevant earlier results}
\label{S:Definitions}
\subsection{Basic notation and terminology}
\label{SS:Notation}
Let $\mathfrak{g}$ be a simple Lie superalgebra with non-degenerate Killing
form ${(\cdot \;,\cdot)}$. Let ${\mathfrak{h} \subset \mathfrak{g}}$ be a Cartan
subsuperalgebra, and let $\Delta \subset \mathfrak{h}^*$ be the set of
roots associated to $\mathfrak{h}$. Fix a set of simple roots $\Gamma$ or
equivalently a Borel $\b$. We will say that a set $X \subset \Delta$ of roots of $\mathfrak{g}$ is {\em closed} if it satisfies the following:
\begin{enumerate}
\item If $\alpha, \beta \in X$ and $\alpha + \beta$ is a root,
then $\alpha + \beta \in X$, and
\item If $\alpha \in X$, then $-\alpha \in X.$
\end{enumerate}
For any positive root $\alpha$ fix $e_{\alpha} \in \mathfrak{g}_{\alpha}$
and pick ${e_{-\alpha} \in \mathfrak{g}_{-\alpha}}$ dual to $e_{\alpha}$
i.e.
\[ (e_{\alpha}, e_{-\alpha}) = 1 \textmd{ for all } \alpha \in
\Delta^+.\]
\noindent
Note that we can do this uniquely up to scalars because all the $\mathfrak{g}_{\alpha}$ are one-dimensional,
(which follows from the nondegeneracy of the Killing form).
Define:
\begin{equation}
\label{defofA}
A_{\alpha} = \left\{ \begin{array}{cl}
(-1)^{|\alpha|} & \textmd{if } \alpha \textmd{ is positive} \\
1 & \textmd{if } \alpha \textmd{ is negative}
\end{array} \right.
\end{equation}
\noindent
It is easy to see that $A_{-\alpha} = (-1)^{|\alpha|}A_{\alpha}$.
We can use $A_{\alpha}$ for instance to write the duals of our
basis vectors in terms of one another:
\[ e_{\alpha}^* = A_{-\alpha} e_{-\alpha}\]
\noindent
or equivalently:
\[ (e_{\alpha},e_{-\alpha}) = A_{-\alpha}. \]
Finally let $\Omega$ be the quadratic Casimir element, i.e.\ the element of ${\mathfrak{g} \otimes \mathfrak{g}}$ corresponding to the Killing form.
The {\em super twist map} ${T_s : V \otimes V \rightarrow V
\otimes V}$ is defined on the homogeneous elements of a given super
vector space $V$ as
$${T_s (a\otimes b) = (-1)^{|a||b|}b\otimes a}.$$
Similarly the {\em super symmetrizing map} ${\operatorname{Alt}_s : V \otimes V \otimes V \rightarrow V \otimes V \otimes V}$ is defined on homogeneous elements by:
\[ \operatorname{Alt}_s(a\otimes b \otimes c) = a \otimes b \otimes c +
(-1)^{|a|(|b|+|c|)} b \otimes c \otimes a +
(-1)^{|c|(|a|+|b|)} c \otimes a \otimes b. \]
\subsection{The classical dynamical Yang-Baxter equation}
\label{SS:SDYBENoParam}
The \emph{classical dynamical Yang-Baxter equation} for a meromorphic function ${r:\mathfrak{h}^*
\rightarrow \mathfrak{g} \otimes \mathfrak{g}}$ is the equation:
\begin{equation}
\label{E:CDYBE}
\operatorname{Alt}_s(dr) +
[r^{12},r^{13}] + [r^{12},r^{23}] + [r^{13},r^{23}] = 0.
\end{equation}
Here, for a fixed (even) basis $\{x_i\}$ for $\mathfrak{h}$, the differential of $r$ is defined as:
\[ \begin{matrix}
dr &:& \mathfrak{h}^* &\longrightarrow& \mathfrak{g} \otimes \mathfrak{g} \otimes \mathfrak{g} \\
&& \lambda &\longmapsto& \sum_i x_i \otimes \frac{\partial
r}{\partial x_i} (\lambda)
\end{matrix} \]
\noindent
Thus, we can see that for $r = \sum
{r}_{(1)} \otimes {r}_{(2)}$, $\operatorname{Alt}_s(dr)$ may be rewritten as:
\[ \sum_i x_i^{(1)} \left(\frac{\partial r }{\partial x_i}
\right)^{(23)} +
\sum_i x_i^{(2)} \left(\frac{\partial r }{\partial x_i}
\right)^{(31)}
+ \sum_i (-1)^{|{r}_{(1)}| |{r}_{(2)}|} x_i^{(3)} \left(
\frac{\partial r}{\partial x_i}\right)^{(12)}.\]
We will say that a meromorphic function ${r : \mathfrak{h}^* \rightarrow \mathfrak{g}
\otimes \mathfrak{g}}$ is a \emph{super dynamical $r$-matrix with coupling
constant} $\epsilon$ if it is a solution to Equation
\eqref{E:CDYBE} and satisfies the {\em generalized unitarity
condition}:
\begin{equation}
\label{dynamicalunitarity}
r(\lambda) + T_s(r)(\lambda) = \epsilon \Omega.
\end{equation}
A super dynamical $r$-matrix $r$ satisfies the \emph{zero weight
condition} if:
\[ [h\otimes 1 + 1 \otimes h, r(\lambda)] = 0 \textmd{ for all }
h \in \mathfrak{h}, \lambda \in \mathfrak{h}^*.\]
\subsection{Classification of super dynamical $r$-matrices of zero weight}
\label{SS:Classify}
In \cite{Kar3} we proved:
\begin{theorem}
\label{0couple0weighttheorem}
Let $\mathfrak{g}$ be a simple Lie superalgebra with non-degenerate
Killing form ${(\cdot \; , \cdot )}$, ${\mathfrak{h} \subset \mathfrak{g}}$ a Cartan
subsuperalgebra, and $\Delta \subset \mathfrak{h}^*$ the set of
roots associated to $\mathfrak{h}$.
\begin{enumerate}
\item Let $X$ be a closed subset of the set of roots $\Delta$ of $\mathfrak{g}$. Let $\nu \in \mathfrak{h}^*$, and let $D = \sum_{i<j} D_{ij} dx_i \wedge
dx_j$ be a closed meromorphic $2$-form on $\mathfrak{h}^*$. If we set
$D_{ij} = -D_{ji}$ for $i \ge j$, then the meromorphic
function:
\begin{equation}
\label{rfor00thm}
r(\lambda) =
\sum_{i,j=1}^N D_{ij}(\lambda) x_i \otimes x_j + \sum_{\alpha
\in X} \frac{A_{\alpha}}{(\alpha,
\lambda-\nu)} e_{\alpha} \otimes e_{-\alpha}
\end{equation}
\noindent
is a super dynamical $r$-matrix with zero weight and zero
coupling constant.
\item Any super dynamical $r$-matrix with zero weight and zero
coupling constant is of this form.
\end{enumerate}
\end{theorem}
We further proved that
there are exactly two types of zero-weight solutions to Equation \eqref{E:CDYBE} satisfying the generalized unitarity condition: the rational ones (solutions of the form given by Equation \eqref{rfor00thm}), with zero coupling constant, and the trigonometric ones,
with a nonzero coupling constant. In fact we explicitly described the general form of the latter, but we will not need that result here.
\section{Quantization of zero weight $r$-matrices}
\label{S:QuantRes}
In this section we prove Theorem \ref{T:QuantRat}.
\subsection{The quantum dynamical Yang-Baxter equation}
\label{SS:QDYBEBasics}
Let $\mathfrak{h}$ be a finite dimensional commutative Lie superalgebra over $\mathbb{C}$, $V$ a finite dimensional super vector space over $\mathbb{C}$ with a diagonal(izable) $\mathfrak{h}$ action, and let $V = \oplus_{\omega \in \mathfrak{h}^*} V[{\omega}]$ be $V$'s $\mathfrak{h}$-weight decomposition. In other words, for every $v \in V[{\omega}]$ and $x \in \mathfrak{h}$, we have $x \cdot v = \omega(x)v$.
In this context, the {\em quantum dynamical Yang-Baxter equation with step $\gamma$} for a function ${R : \mathfrak{h}^* \rightarrow \operatorname{End}(V \otimes V)}$ is the equation:
\begin{equation}
\label{E:QDYBE}
R^{12}(\lambda-\gamma h^{(3)})R^{13}(\lambda)R^{23}(\lambda-\gamma h^{(1)})= R^{23}(\lambda)R^{13}(\lambda-\gamma h^{(2)})R^{12}(\lambda).
\end{equation}
Here the operator $R^{ij}$ is interpreted to be acting nontrivially on the $i$th and the $j$th components of a given $3$-tensor, and the notation $h^{(k)}$ is to be replaced by the weight of the $k$th component of the same. For instance ${R^{12)}(\lambda-\gamma h^{(3)})(v_1 \otimes v_2 \otimes v_3)} = {\left (R(\lambda-\gamma \omega_3)(v_1 \otimes v_2) \right ) \otimes v_3}$ whenever $v_3 \in V[{\omega_3}]$.
We will say that an invertible function ${R : \mathfrak{h}^* \rightarrow \operatorname{End}(V \otimes V)}$ is a {\em super dynamical $R$-matrix} if it is a solution to Equation \eqref{E:QDYBE} and satisfies the {\em zero weight condition}:
\[ [h\otimes 1 + 1 \otimes h, R(\lambda)] = 0 \textmd{ for all }
h \in \mathfrak{h}, \lambda \in \mathfrak{h}^*.\]
\subsection{The quantization problem}
\label{SS:QuantProb}
Let ${R_{\gamma} : \mathfrak{h}^* \rightarrow \operatorname{End}(V \otimes V)}$ be a smooth family of solutions to Equation \eqref{E:QDYBE} such that:
\[ R_{\gamma}(\lambda) = 1 -\gamma r(\lambda) + O(\gamma^2).\]
Then the function $r(\lambda)$ satisfies Equation \eqref{E:CDYBE} and is called the {\em semi-classical limit} of $R_{\gamma}(\lambda)$. In the same setup $R_{\gamma}(\lambda)$ is called a {\em quantization} of $r(\lambda)$.
Alternatively we can begin with a super dynamical $r$-matrix $r : \mathfrak{h}^* \rightarrow \operatorname{End}(V \otimes V)$ defined on an open subset $U$ of $\mathfrak{h}^*$. We then call $r$ {\em quantizable} if there is a power series in $\gamma$ of the form:
\[ R_{\gamma}(\lambda) = 1 -\gamma r(\lambda) + \sum_{n=2}^{\infty} \gamma^nr_n(\lambda)\]
satisfying Equation \eqref{E:QDYBE}.
The {\em quantization problem} for us must now be obvious: Given a super dynamical $r$-matrix construct a power series $R_{\gamma}(\lambda)$ of the form above (or prove the impossibility of such a construction).
\subsection{Multiplicative forms}
\label{SS:Forms}
In the following we will make use of
multiplicative $k$-forms
a la \cite[\S\S 1.4]{EV2}. We now briefly recall some of the relevant constructions to keep our paper self-contained.
Let $V = V_{\overline{0}} \oplus V_{\overline{1}}$ be a super vector space with a homogeneous linear coordinate system $\lambda_1, \cdots, \lambda_N$. We define a {\em multiplicative $k$-form} on $V$ to be a collection:
\[ \phi = \{\phi_{i_1,\dots,i_k}(\lambda_1, \cdots, \lambda_N) \}\]
of meromorphic functions, where the ordered $k$-tuples $(i_1,\dots,i_k)$ run through all $k$-element subsets of $\{1,\cdots, N\}$, and we require that:
\[ \phi_{\tau(I)}\phi_I = 1 \]
whenever $I = (i_1,\dots,i_k)$ is some ordered $k$-tuple and $\tau(I)$ is a transposition $(i_si_{s+1})$ switching the consecutive indices $i_s,i_{s+1}$ for some $1 \le s < k$. Let $\Omega^k(V) = \Omega^k$ be the set of all multiplicative $k$-forms on $V$. There is a natural abelian group structure on $\Omega^k$.
Now we fix a complex number $\gamma$. We define, for each $i = 1, \cdots, N$, an operator $\delta_i$ on the space of all meromorphic functions on the $N$ variables $\lambda_1, \cdots, \lambda_N$:
\begin{equation*}
\delta_i : f(\lambda_1, \cdots, \lambda_N) \longmapsto
f(\lambda_1, \cdots, \lambda_N) / f(\lambda_1, \cdots, \lambda_i-\gamma, \cdots, \lambda_N).
\end{equation*}
We next define an operator $d_{\gamma} : \Omega^k \rightarrow \Omega^{k+1}$ mapping $\phi$ to $d_{\gamma}\phi$ given by:
\begin{equation*}
(d_{\gamma}\phi)_{i_1,\dots,i_{k+1}}(\lambda_1, \cdots, \lambda_N) = \prod_{s=1}^{k+1} \left ( \delta_{i_s} \phi_{i_1,\dots,i_{s-1}, i_{s+1}, \dots, i_{k+1}}(\lambda_1, \cdots, \lambda_N) \right )^{(-1)^{s+1}}.
\end{equation*}
A multiplicative $k$-form $\phi$ is {\it $\gamma$-closed} if $d_{\gamma}\phi = 0$. Obviously, $d^2_{\gamma} = 0$ because the zero element of $\Omega^k$ is the form $\{\phi_{i_1,\dots,i_k}(\lambda_1, \cdots, \lambda_N)\equiv 1 \}$.
Still following \cite[\S\S 1.4]{EV2}, we say that a smooth family $\phi(\gamma) =
\{\phi_{i_1,i_2,\dots,i_k}(\lambda_1, \lambda_2, \cdots, \lambda_N,\gamma) \}$ of multiplicative $k$-forms with
\[ \phi_I(\lambda,\gamma) = 1 - \gamma C_I(\lambda)+O(\gamma^2) \text{ for each } I = (i_1,i_2,\dots,i_k)\]
is a {\em quantization} of the differential form
\[C = \sum_{i_1 < i_2 < \cdots < i_k} C_{i_1,i_2,\dots,i_k}(\lambda) \, dx_{i_1}\wedge dx_{i_2} \wedge \cdots \wedge dx_{i_k}.\]
Conversely we will say that a differential form $C$ given as above is {\em quantizable} if there exists a power series in $\gamma$:
\[ \phi_I(\lambda,\gamma) = 1 - \gamma C_I(\lambda)+\sum_{n=2}^{\infty} \gamma^n C_{n;I}(\lambda) \text{ for each } I = (i_1,i_2,\dots,i_k)\]
convergent for small $|\gamma|$ and fixed $\lambda \in U$, where $U$ is an open polydisc in $\mathbb{C}^N$, in such a way that $\{\phi_{i_1,i_2,\dots,i_k}(\lambda_1, \lambda_2, \cdots, \lambda_N,\gamma) \}$ is a multiplicative $k$-form.
Here is Lemma 1.1 from \cite{EV2}:
\begin{lemma}
\label{L:Lemma1.1}
Every closed holomorphic differential $k$-form $C$ defined on an open polydisc is quantizable to a holomorphic multiplicative closed $k$-form $\phi(\gamma)$.
\end{lemma}
\noindent
The proof is included in \cite{EV2} and will not be repeated here.
\subsection{$R$-matrices of $\mathfrak{gl}(m,n)$ type}
\label{SS:glmntype}
Now let $\mathfrak{h}$ be a finite dimensional commutative Lie superalgebra over $\mathbb{C}$ and let $V = V_{\overline{0}} \oplus V_{\overline{1}}$ be a finite dimensional semisimple $\mathfrak{h}$-module whose weights $W= \{\omega_1, \omega_2, \cdots, \omega_N\}$ make up a basis for $\mathfrak{h}^*$. We label the elements of the dual basis for $\mathfrak{h}$ by $x_i$; clearly the $x_i$ are all even, and $\dim_{\mathbb{C}} V = \dim_{\mathbb{C}} \mathfrak{h} = N$. Let $\{v_1, v_2, \cdots, v_N\}$ be an $\mathfrak{h}$-eigenbasis for $V$ with $x_i v_j = \delta_{ij}v_j$. By relabeling as needed, we can assume that $v_1, v_2, \cdots v_m$ is a basis for $V_{\overline{0}}$, the even part of $V$, while $v_{m+1},v_{m+2} \cdots, v_{N}$ is a basis for $V_{\overline{1}}$, the odd part of $V$. Let $n = N-m$. We will say that a super dynamical $R$-matrix $R : \mathfrak{h}^* \rightarrow \operatorname{End}(V \otimes V)$ for such $\mathfrak{h}$ and $V$ is an {\em $R$-matrix of $\mathfrak{gl}(m,n)$ type}. The super dynamical $R$-matrices in this paper will all be of this kind unless explicitly noted otherwise.
In this setup $V \otimes V$ has the following weight decomposition:
\begin{equation}
\label{E:WeightDecomp}
V \otimes V = \left ( \bigoplus_{i=1}^N V_{ii} \right ) \oplus \left ( \bigoplus_{i <j} V_{ij} \right ).
\end{equation}
Here $V_{ii} = \mathbb{C} (v_i \otimes v_i)$ and $V_{ij} = \mathbb{C} (v_i \otimes v_j) \oplus \mathbb{C} (v_j \otimes v_i)$. It is clear that $V_{ii}$ will always belong to the even part of $V \otimes V$ while $V_{ij}$ may be even or odd. In particular, if exactly one of $v_i$ and $v_j$ is odd, then $V_{ij}$ will be odd, otherwise it will be even. In other words, if we introduce the notation
\[ \sigma(i) = \begin{cases}
0 & \text{ if } i \le m \\
1 & \text{ if } i > m,
\end{cases}\]
then $V_{ij}$ is odd if and only if $\sigma(i)+\sigma(j) = 1$.
We can introduce a basis $\{E_{ij} \, \vert 1 \le i,j\le N\}$ for $\operatorname{End}(V)$ by setting $E_{ij} v_k = \delta_{jk} v_i$. Recall that we require our dynamical $R$-matrices to satisfy the zero weight condition. Then we can write any $R$-matrix $R : \mathfrak{h}^* \rightarrow \operatorname{End}(V\otimes V)$ of $\mathfrak{gl}(m,n)$ type in the form:
\[ R(\lambda) = \sum_{i,j=1}^N \alpha_{ij}(\lambda) E_{ii} \otimes E_{jj} + \sum_{i \neq j} \beta_{ij}(\lambda) E_{ji}\otimes E_{ij} \]
for some meromorphic functions $\alpha_{ij}, \beta_{ij} : \mathfrak{h}^* \rightarrow \mathbb{C}$.
\subsection{Gauge transformations for super dynamical $r$-matrices}
\label{SS:Gauge}
Before we can prove Theorem \ref{T:QuantRat},
we will need to simplify the expression \eqref{rfor00thm}. In order to do that we first discuss briefly the appropriate gauge transformations for super dynamical $r$-matrices of the form $r : \mathfrak{h}^* \rightarrow \operatorname{End}(V \otimes V)$. Note that if we assume the setup of \S\S\ref{SS:glmntype}, then we can use the $\{E_{ij}\}$ basis for $\operatorname{End}(V)$. Then the zero weight condition on $r$ implies that $r$ has to be in the form:
\[ r(\lambda) = \sum_{i,j=1}^N \alpha_{ij}(\lambda) E_{ii} \otimes E_{jj} + \sum_{i \neq j} \beta_{ij}(\lambda) E_{ji}\otimes E_{ij} \]
for some meromorphic functions $\alpha_{ij}, \beta_{ij} : \mathfrak{h}^* \rightarrow \mathbb{C}$.
The following is a list of the gauge transformations for such $r$ which we will need in the rest of this note (cf.\ \cite{EV, EV2}):
\begin{enumerate}\addtolength{\itemsep}{0.4\baselineskip}
\item The transformation:
\[ r(\lambda) \longmapsto r(\lambda) + \sum_{i,j=1}^N D_{ij}(\lambda) E_{ii} \otimes E_{jj} \]
for some closed meromorphic differential $2$-form $D = \sum_{i<j} D_{ij} dx_i \wedge dx_j$ on $\mathfrak{h}^*$. $D_{ij}$ is then extended to all $i,j$ by setting
$D_{ij} = -D_{ji}$ for $i \ge j$. ($D_{ii} = 0$ for each $i$.)
\item The transformation:
\[ r(\lambda) \longmapsto r(\lambda+\mu) \]
for $\mu \in \mathfrak{h}^*$.
\item The transformation:
\[ r(\lambda) \longmapsto c r(c\lambda) \]
for a nonzero complex number $c \in \mathbb{C}$.
\item The transformation:
\[ r(\lambda) \longmapsto (\tau \otimes \tau) r(\tau^{-1} \cdot \lambda) (\tau^{-1} \otimes \tau^{-1}) \]
for some permutation $\tau \in S_N$
of the coordinates in $\mathfrak{h}^*$ and $V$.
\item The transformation:
\[ r(\lambda) \longmapsto r(\lambda) + c\operatorname{Id}\]
for a nonzero complex number $c \in \mathbb{C}$.
\end{enumerate}
Each of these transformations corresponds to a specific quantum gauge transformation allowed for super dynamical $R$-matrices (cf.\ \cite{EV2}).
We will briefly study
these quantum gauge transformations
in \S\S\ref{SS:SuperQuantGauge}, in the context of the classification problem for super dynamical $R$-matrices.
It is easy to show that the transformations (1-5) map a given super dynamical $r$-matrix to another. We omit the proofs here since they are straightforward modifications of those in \cite{EV}. We will say that two super dynamical $r$-matrices are {\em gauge equivalent} (or simply {\em equivalent} when the context is unambiguous) if one can be obtained from the other by a sequence of gauge transformations.
We can now simplify the expression in Theorem \ref{0couple0weighttheorem} using the above. Let $X \subset \{ 1, 2, \cdots, N\}$ be a subset of indices and write it as a disjoint union of subintervals $X = X_1 \sqcup X_2 \sqcup \cdots \sqcup X_n$. In other words, every subinterval $X_k$ should be of the form $X_k = [i_k,i_{k+1}, i_{k+2}, \cdots, j_k]$, and $j_k < i_{k+1}$ for each $k$. Define:
\begin{equation*}
A_{ij} = \left\{ \begin{array}{cl}
(-1)^{\sigma(i)+\sigma(j)} &\text{ if } i < j, \\
1 & \text{ if } i > j,
\end{array} \right.
\end{equation*}
(cf.\ Equation \eqref{defofA}). Now applying the above transformations
and using the $\{E_{ij}\}$ basis, we can show that the super dynamical $r$-matrix in Equation \eqref{rfor00thm} is (gauge-)equivalent to:
\[ r_{\text{rat}}(\lambda) = \sum_{k=1}^n \left ( \sum_{i,j \in X_k, i\neq j} \frac{A_{ij}}{\lambda_{ij}} E_{ij} \otimes (E_{ij})^* \right ). \]
Since $(E_{ij})^* = A_{ji}(-1)^{\sigma(i)}E_{ji}$, this further reduces to:
\begin{equation}
\label{E:rRatForm}
r_{\text{rat}}(\lambda) = \sum_{k=1}^n \left ( \sum_{i,j \in X_k, i\neq j} \frac{(-1)^{\sigma(j)}}{\lambda_{ij}} E_{ij} \otimes E_{ji} \right ).
\end{equation}
\subsection{The construction}
\label{SS:Constructions}
We are finally ready to construct the quantization necessary for Theorem \ref{T:QuantRat}. Let $\mathfrak{h}$ and $V$ be as in \S\S\ref{SS:glmntype}. We will once again use the basis $\{E_{ij}\}$ for $\operatorname{End}(V)$ and we will write $X \subset \{ 1, 2, \cdots, N\}$ as a disjoint union of subintervals $X = X_1 \sqcup X_2 \sqcup \cdots \sqcup X_n$
Consider:
\[ R_{\text{rat}}(\lambda,\gamma) = \text{Id} + \sum_{k=1}^n \, \sum_{i,j \in X_k,i\neq j} \frac{\gamma}{\lambda_{ij}} \left ( E_{ii} \otimes E_{jj} + (-1)^{\sigma(i)} E_{ji}\otimes E_{ij} \right )\]
Then $R_{\text{rat}}(\lambda,\gamma)$ satisfies Equation \eqref{E:QDYBE} (cf.\ Theorem \ref{T:ClassifyR1}), and its semi-classical limit is:
\[r^{\prime}_{\text{rat}}(\lambda) = \sum_{k=1}^n \, \sum_{i,j \in X_k,i\neq j} \frac{-1}{\lambda_{ij}} \left ( E_{ii} \otimes E_{jj} + (-1)^{\sigma(i)} E_{ji}\otimes E_{ij} \right ).\]
Using the gauge transformation of type (1) with the closed form:
\[ D =\sum_{k=1}^n \, \sum_{i,j \in X_k,i< j} D_{ij} dx_i \wedge dx_j = \sum_{k=1}^n \, \sum_{i,j \in X_k,i< j} \frac{-1}{\lambda_{ij}} dx_{i} \wedge dx_{j}, \]
we can show that $r^{\prime}_{\text{rat}}$ is (gauge-)equivalent to the $r_{\text{rat}}$ of Equation \eqref{E:rRatForm}.
Together with Lemma \ref{L:Lemma1.1} this proves Theorem \ref{T:QuantRat}. \qed
\section{The Quantum Picture}
\label{S:QuantPict}
In this section we define {\it the super Hecke condition} (\S\S\ref{SS:SuperHecke}) which is a generalized unitarity condition. Using this notion, we state and prove (\S\S\ref{SS:TheoremsStated}) a theorem in the spirit of Theorem 1.2 of \cite{EV2}. This is a result that provides a partial classification of all super dynamical $R$-matrices satisfying the super Hecke condition. It turns out that the super Hecke condition encodes the constraint on the coupling constant in the classical case.
\subsection{Some initial computations}
\label{SS:QuantumComputations}
Let $\mathfrak{h}$ and $V$ be as in \S\S\ref{SS:glmntype}. We will once again use the basis $\{E_{ij}\}$ for $\operatorname{End}(V)$ and throughout this section we will once again restrict ourselves to the study of $R$-matrices of $\mathfrak{gl}(m,n)$ type. Recall that this means, in particular, that the super vector space $V \otimes V$ has the weight decomposition given in \eqref{E:WeightDecomp}.
More specifically, a super dynamical $R$-matrix $R : \mathfrak{h}^* \rightarrow \operatorname{End}(V\otimes V)$ of $\mathfrak{gl}(m,n)$ type can be written in the form:
\[ R(\lambda) = \sum_{i,j=1}^N \alpha_{ij}(\lambda) E_{ii} \otimes E_{jj} + \sum_{i \neq j} \beta_{ij}(\lambda) E_{ji}\otimes E_{ij} \]
for some meromorphic functions $\alpha_{ij}, \beta_{ij} : \mathfrak{h}^* \rightarrow \mathbb{C}$.
If we now assume for simplicity (and for other reasons which will become clearer in \S\S\ref{SS:SuperHecke}) that our super dynamical $R$-matrices all satisfy $\alpha_{ii} = 1$ for all $i$, we can rewrite the above as:
\begin{equation}
\label{E:FormofR}
R(\lambda) = \sum_{i=1}^N E_{ii} \otimes E_{ii} + \sum_{i \neq j} \alpha_{ij}(\lambda) E_{ii} \otimes E_{jj} + \sum_{i \neq j} \beta_{ij}(\lambda) E_{ji}\otimes E_{ij} \end{equation}
for some meromorphic functions $\alpha_{ij}, \beta_{ij} : \mathfrak{h}^* \rightarrow \mathbb{C}$.
In this subsection we list a few conditions on these $\alpha$ and $\beta$ functions. We limit ourselves to simply summarizing the results of necessary computations; the explicit derivations can be found in Appendix \ref{A:AlphaBeta}.
By applying the two sides of Equation \eqref{E:QDYBE} for an $R$ of the form \eqref{E:FormofR} to a basis element $v_i \otimes v_i \otimes v_k$ of $V^{\otimes 3}$ with $i \neq k$ and setting the coefficients of like terms equal to one another, we obtain:
\begin{equation}
\label{E:coeffkii}
\alpha_{ki}(\lambda-\gamma \omega_i)\beta_{ik}(\lambda)\alpha_{ik}(\lambda-\gamma \omega_i) + (\beta_{ik}(\lambda-\gamma \omega_i))^2= \beta_{ik}(\lambda-\gamma \omega_i)
\end{equation}
and
\begin{eqnarray}
\label{E:coeffiki}
{(-1)^{\sigma(i)+\sigma(k)}} \beta_{ki}(\lambda-\gamma \omega_i)\beta_{ik}(\lambda) \alpha_{ik}(\lambda-\gamma \omega_i) &+& \alpha_{ik}(\lambda-\gamma \omega_i)\beta_{ik}(\lambda-\gamma \omega_i) \cr
&=& \beta_{ik}(\lambda)\alpha_{ik}(\lambda-\gamma \omega_i).
\end{eqnarray}
Note that Equation \eqref{E:coeffkii} is identical to \cite[Eqn.1.8.4]{EV2} while Equation \eqref{E:coeffiki} is a signed version of \cite[Eqn.1.8.5]{EV2}.
Similarly we can derive the following equations by applying the two sides of Equation \eqref{E:QDYBE} to a basis element $v_i \otimes v_j \otimes v_k$ with $i,j,k$ all distinct:
\begin{equation}
\label{E:coeffijk}
\alpha_{ij}(\lambda-\gamma \omega_k)\alpha_{ik}(\lambda)\alpha_{jk}(\lambda-\gamma \omega_i) = \alpha_{jk}(\lambda)\alpha_{ik}(\lambda-\gamma \omega_j)\alpha_{ij}(\lambda)
\end{equation}
which is precisely the same as \cite[Eqn.1.8.6]{EV2};
\begin{equation}
\label{E:coeffikj}
\alpha_{ik}(\lambda-\gamma \omega_j)\alpha_{ij}(\lambda) \beta_{jk}(\lambda-\gamma \omega_i)= \beta_{jk}(\lambda)\alpha_{ik}(\lambda-\gamma \omega_j)\alpha_{ij}(\lambda)
\end{equation}
which is precisely the same as \cite[Eqn.1.8.7]{EV2};
\begin{equation}
\label{E:coeffjik}
\beta_{ij}(\lambda-\gamma \omega_k)\alpha_{ik}(\lambda)\alpha_{jk}(\lambda-\gamma \omega_i) = \alpha_{ik}(\lambda)\alpha_{jk}(\lambda-\gamma\omega_i)\beta_{ij}(\lambda)
\end{equation}
which is precisely the same as \cite[Eqn.1.8.8]{EV2};
\begin{eqnarray}
\label{E:coeffjki}
(-1)^{\sigma(k)}\beta_{kj}(\lambda-\gamma \omega_i)\beta_{ik}(\lambda) \alpha_{jk}(\lambda-\gamma \omega_i)&+&(-1)^{\sigma(j)} \alpha_{jk}(\lambda-\gamma \omega_i)\beta_{ij}(\lambda) \beta_{jk}(\lambda-\gamma \omega_i)\cr
&=&(-1)^{\sigma(i)} \beta_{ik}(\lambda)\alpha_{jk}(\lambda-\gamma\omega_i)\beta_{ij}(\lambda)
\end{eqnarray}
which is a signed analogue of \cite[Eqn.1.8.9]{EV2};
\begin{eqnarray}
\label{E:coeffkji}
\alpha_{kj}(\lambda-\gamma \omega_i)\beta_{ik}(\lambda)\alpha_{jk}(\lambda-\gamma \omega_i) &+& \beta_{jk}(\lambda-\gamma \omega_i)\beta_{ij}(\lambda) \beta_{jk}(\lambda-\gamma \omega_i) =\cr
\alpha_{ji}(\lambda)\beta_{ik}(\lambda-\gamma \omega_j)\alpha_{ij}(\lambda) &+&(-1)^{\sigma(i)+\sigma(j)}\beta_{ij}(\lambda)\beta_{jk}(\lambda-\gamma\omega_i)\beta_{ij}(\lambda)\end{eqnarray}
which is a signed analogue of \cite[Eqn.1.8.10]{EV2}; and
\begin{eqnarray}
\label{E:coeffkij}
\beta_{ik}(\lambda-\gamma \omega_j)\alpha_{ij}(\lambda)\beta_{jk}(\lambda-\gamma \omega_i)&=&\cr
\beta_{ji}(\lambda)\beta_{ik}(\lambda-\gamma \omega_j)\alpha_{ij}(\lambda) &+&\alpha_{ij}(\lambda)\beta_{jk}(\lambda-\gamma\omega_i)\beta_{ij}(\lambda)\end{eqnarray}
which is precisely the same as \cite[Eqn.1.8.11]{EV2}.
\subsection{The Super Hecke Condition}
\label{SS:SuperHecke}
Let $p \neq -q$ be two complex numbers. Set $\check{R} = P_s R$ where $P_s \in \operatorname{End}(V \otimes V)$ is the element corresponding to $T_s$.
In a way analogous to \cite{EV2} we will say that a function $R : \mathfrak{h}^* \rightarrow \operatorname{End}(V \otimes V)$ satisfies the {\em strong super Hecke condition} if it has the following properties:
\begin{enumerate}\addtolength{\itemsep}{0.4\baselineskip}
\item The function preserves the weight decomposition given in \eqref{E:WeightDecomp}.
\item For any $i =1, 2, \cdots, N$, and $\lambda \in \mathfrak{h}^*$, $\check{R}(\lambda)(v_i \otimes v_i) = p (v_i \otimes v_i)$.
\item For any $i \neq j$, and $\lambda \in \mathfrak{h}^*$, the operator $\check{R}(\lambda)$ restricted to $V_{ij}$ has eigenvalues $(-1)^{\sigma(i)+\sigma(j)}p$ and $-(-1)^{\sigma(i)+\sigma(j)}q$.
\end{enumerate}
A function $R : \mathfrak{h}^* \rightarrow \operatorname{End}(V \otimes V)$ satisfies the {\em weak super Hecke condition} if it has the following properties (cf\ \cite[Eq.1.3.6]{EV2}) :
\begin{enumerate}\addtolength{\itemsep}{0.4\baselineskip}
\item The function preserves the weight decomposition given in \eqref{E:WeightDecomp}.
\item For any $\lambda \in \mathfrak{h}^*$ and $i,j \le N$, $(\check{R}(\lambda) -(-1)^{\sigma(i)+\sigma(j)} p)(\check{R}(\lambda)+(-1)^{\sigma(i)+\sigma(j)}q)= 0$ when restricted to $V_{ij}$.
\end{enumerate}
Just as in the non-graded case these two properties are intimately related. In fact whenever a continuous family $R_t : \mathfrak{h}^* \rightarrow \operatorname{End}(V\otimes V)$, $t \in [0,1]$, of meromorphic functions, analytic for $0 < t < 1$ and $R_0 = \operatorname{Id}$, satisfies the weak super Hecke condition for all $t$, then $R_t$ satisfies the strong super Hecke condition as well.
Hence we will simply assume that $R$ satisfies both whenever we say that $R$ satisfies the super Hecke condition.
Now we consider a super dynamical $R$-matrix $R(\lambda)$ with step $\gamma=1$ which satisfies the super Hecke condition with $p=1$ and $q$ arbitrary. Then we can see that $\alpha_{ii} = 1$ and $R$ has the form given by Equation \eqref{E:FormofR}. Furthermore, whenever $i \neq j$, we have:
\begin{equation}
\label{E:trace}
(-1)^{\sigma(i)}\beta_{ij}(\lambda) + (-1)^{\sigma(j)}\beta_{ji}(\lambda) = (-1)^{\sigma(i)+\sigma(j)}(1-q)
\end{equation}
and
\begin{equation}
\label{E:determinant}
(-1)^{\sigma(i)+\sigma(j)}\beta_{ij}(\lambda)\beta_{ji}(\lambda)-\alpha_{ij}(\lambda)\alpha_{ji}(\lambda)=-q
\end{equation}
obtained from the trace and determinant of $\check{R}$ on $V_{ij}$. Note that these are signed versions of \cite[Eqn.1.8.2]{EV2} and \cite[Eqn.1.8.3]{EV2}.
At this point it is easy to notice that if $i\neq j$, then assuming $\alpha_{ij} \equiv 0$ implies that
\[ \beta_{ij}(\lambda)\beta_{ji}(\lambda) = -(-1)^{\sigma(i)+\sigma(j)}q \]
by Equation \eqref{E:determinant}, and Equation \eqref{E:coeffkii} gives us:
\[ (\beta_{ij}(\lambda))^2= \beta_{ij}(\lambda) \text{ and } (\beta_{ji}(\lambda))^2= \beta_{ji}(\lambda). \]
These then contradict with Equation \eqref{E:trace}. Therefore $\alpha_{ij}$ cannot be identically zero. Similarly we can show that
\begin{equation}
\label{E:alphasbetas}
\alpha_{ij}(\lambda)\alpha_{ji}(\lambda) = ((-1)^{\sigma(i)}\beta_{ij}(\lambda)+(-1)^{\sigma(i)+\sigma(j)}q) ((-1)^{\sigma(j)}\beta_{ji}(\lambda)+(-1)^{\sigma(i)+\sigma(j)}q)
\end{equation}
and therefore the quantity $(-1)^{\sigma(i)}\beta_{ij}(\lambda)+(-1)^{\sigma(i)+\sigma(j)}q$ is also not identically zero.
Finally we consider a super dynamical $R$-matrix $R(\lambda)$ of the form \eqref{E:FormofR} with step $\gamma=1$, and assume that $R(\lambda)$ satisfies the super Hecke property with Hecke parameters $p=1$ and $q$. Then the collection of functions:
\begin{equation}
\label{E:phiDef}
\phi = \{ \phi_{ij}(\lambda) \} \text{ where } \phi_{ij}(\lambda) = \frac{(-1)^{\sigma(i)}\beta_{ij}(\lambda)+(-1)^{\sigma(i)+\sigma(j)}q}{\alpha_{ij}(\lambda)} \text{ for } i\neq j
\end{equation}
is a $\gamma$-closed multiplicative $2$-form with $\gamma = 1$. This follows from our earlier computations and in particular from Equation \eqref{E:alphasbetas}; just as in \cite{EV2}, Equations \eqref{E:coeffijk} and \eqref{E:coeffikj} are used to show that $d_{\gamma}\phi = 0$. We will use this $\phi$ in the next subsection.
\subsection{Gauge transformations for super dynamical $R$-matrices}
\label{SS:SuperQuantGauge}
Let us now assume that we have a super dynamical $R$-matrix of $\mathfrak{gl}(m,n)$ type and we write it in the form given by Equation \eqref{E:FormofR}. The following is a list of the gauge transformations for such $R(\lambda)$ which we will need in the rest of this note (cf.\ \cite[\S\S1.4]{EV2}):
\begin{enumerate}
\item The transformation:
\[ R(\lambda) \longmapsto \sum_{i=1}^N E_{ii} \otimes E_{ii} + \sum_{i \neq j} \phi_{ij}(\lambda) \alpha_{ij}(\lambda) E_{ii} \otimes E_{jj} + \sum_{i \neq j} \beta_{ij}(\lambda) E_{ji}\otimes E_{ij} \]
for some meromorphic s-multiplicative $\gamma$-closed multiplicative $2$-form $\{\phi_{ij}\}$ on $\mathfrak{h}^*$.
\item The transformation:
\[ R(\lambda) \longmapsto (\tau \otimes \tau) R(\tau^{-1} \cdot \lambda) (\tau^{-1} \otimes \tau^{-1}) \]
for some permutation $\tau \in S_N$
of the coordinates in $\mathfrak{h}^*$ and $V$.
\item The transformation:
\[ R(\lambda) \longmapsto cR(\lambda) \]
for a nonzero complex number $c \in \mathbb{C}$.
\item The transformation:
\[ R(\lambda) \longmapsto R(c\lambda+\mu) \]
for a nonzero complex number $c \in \mathbb{C}$ and an element $\mu \in \mathfrak{h}^*$.
\end{enumerate}
It is easy to see that transformations of type (1-3) transform a super dynamical $R$-matrix with step $\gamma$ to another one with step $\gamma$. In particular it suffices to check that the relevant equations in \S\S\ref{SS:QuantumComputations} and \S\S\ref{SS:SuperHecke} for $\alpha_{ij}(\lambda)$ and $\beta_{ij}(\lambda)$ are invariant with respect to them.
Transformations of type (4) modify the step $\gamma$ to $\gamma/c$.
In all cases the super Hecke property is preserved. While the transformations of type (3) modify the relevant Hecke parameters, the rest preserve them. Moreover, any super dynamical $R$-matrix $R(\lambda)$ of Hecke type can be shown to be (gauge-)equivalent to a super dynamical $R$-matrix $R(\lambda)$ with step $\gamma=1$ which satisfies the super Hecke condition with $p=1$ and $q$ arbitrary. This requires simply gauge transformations of types (3) and (4).
At this point we can specialize \eqref{E:FormofR} even further. Once again let $R(\lambda)$ be a super dynamical $R$-matrix of the form \eqref{E:FormofR} with step $\gamma$ satisfying the super Hecke condition. As justified by the above we can assume that the step $\gamma=1$ and the Hecke parameters are $p=1$ and $q$ arbitrary. Then if we apply the gauge transformation of type (1) to this $R(\lambda)$ using the reciprocal of the multiplicative $2$-form given in \eqref{E:phiDef}, we obtain a new super dynamical $R$-matrix (satisfying the super Hecke condition with the same parameters) whose coefficients now satisfy
\begin{equation}
\label{E:alphabeta}
{\alpha_{ij}(\lambda)} = (-1)^{\sigma(i)}\beta_{ij}(\lambda)+(-1)^{\sigma(i)+\sigma(j)}q \text{ for } i\neq j.
\end{equation}
\subsection{Statement of the main quantum theorem}
\label{SS:TheoremsStated}
We are now ready to state the main result of this section:
\begin{theorem}[Classification Theorem for Equal Parameters]
\label{T:ClassifyR1}
Let $\mathfrak{h}$ be a finite dimensional commutative Lie superalgebra over $\mathbb{C}$ and let $V = V_{\overline{0}} \oplus V_{\overline{1}}$ be a finite dimensional semisimple $\mathfrak{h}$-module whose weights make up a basis for $\mathfrak{h}^*$. Let $N = \dim_{\mathbb{C}} V = \dim_{\mathbb{C}} \mathfrak{h}$ .
\begin{enumerate}
\item
Let $X \subset \{ 1, 2, \cdots, N\}$ be a subset of indices written as a disjoint union of subintervals $X = X_1 \sqcup X_2 \sqcup \cdots \sqcup X_n$.
Fix a $\gamma$-quasiconstant $\mu : \mathfrak{h}^* \rightarrow \mathfrak{h}^*$ with $\gamma=1$. Define scalar meromorphic $\gamma$-quasiconstant functions $\mu_{ij} : \mathfrak{h}^* \rightarrow \mathbb{C}$ by $\mu_{ij}(\lambda) = x_1(\mu(\lambda)) - x_j(\mu(\lambda))$.
Then the meromorphic function $R_X : \mathfrak{h}^* \rightarrow \operatorname{End}(V \otimes V)$ defined by:
\begin{equation*}
R_X(\lambda) = \sum_{i,j=1}^N (-1)^{\sigma(i)+\sigma(j)}E_{ii} \otimes E_{jj} + \sum_{s=1}^n \left ( \sum_{i,j \in X_s, i \neq j} \frac{1}{\lambda_{ij} - \mu_{ij}(\lambda)} [E_{ii}\otimes E_{jj} + (-1)^{\sigma(i)}E_{ji} \otimes E_{ij}] \right )
\end{equation*}
is a super dynamical $R$-matrix of $\mathfrak{gl}(m,n)$ type satisfying the super Hecke condition with $p=q=1$ and step $\gamma=1$.
\item Every super dynamical $R$-matrix of $\mathfrak{gl}(m,n)$ type satisfying the super Hecke condition with $p=q$ is equivalent to a super dynamical $R$-matrix of this form.
\end{enumerate}
\end{theorem}
\subsection{Proof of Theorem \ref{T:ClassifyR1}}
\label{SS:QProof1}
Now we let $R(\lambda)$ be a super dynamical $R$-matrix satisfying the super Hecke condition with parameters $p=q$.
As we showed in the previous subsection, we can use appropriate gauge transformations to ensure that $\gamma = p=q=1$. Then Equation \eqref{E:alphabeta} becomes:
\[ {\alpha_{ij}(\lambda)} = (-1)^{\sigma(i)}\beta_{ij}(\lambda)+(-1)^{\sigma(i)+\sigma(j)} \text{ for } i\neq j. \]
Next look at Equation \eqref{E:coeffiki} for indices $i,j$. Clearly $\beta_{ij}(\lambda) = \beta_{ji}(\lambda) \equiv 0$ is one solution, so we assume that this is not the case. Since we showed earlier in \S\S\ref{SS:SuperHecke} that $\alpha_{ij}$ cannot be identically zero, we obtain from the two versions (for $i,i,j$ and $j,j,i$, reading the coefficients of $i,j,i$ and $j,i,j$ respectively), the following two conditions on $\beta_{ij}$:
\begin{equation}
\label{E:betaReciprocal1}
\frac{1}{\beta_{ij}(\lambda)} - \frac{1}{\beta_{ij}(\lambda-\omega_i)} = 1
\text{ for } i\neq j, \end{equation}
and
\begin{equation}
\label{E:betaReciprocal2}
\frac{1}{\beta_{ij}(\lambda)} - \frac{1}{\beta_{ij}(\lambda-\omega_j)} = -(-1)^{\sigma(i)+\sigma(j)} \text{ for } i\neq j. \end{equation}
where we are using $(-1)^{\sigma(i)}\beta_{ij}(\lambda) + (-1)^{\sigma(j)}\beta_{ji}(\lambda)=0$ or equivalently $A_{ij}\beta_{ij}(\lambda) + A_{ji}\beta_{ji}(\lambda)=0$
(obtained from Equation \eqref{E:trace} with $q=1$).
Rewriting these equations as:
\begin{equation*}
\beta_{ij}(\lambda-\omega_i) = \frac{\beta_{ij}(\lambda)}{1-\beta_{ij}(\lambda)}
\end{equation*}
and
\begin{equation*}
\beta_{ji}(\lambda-\omega_i) = \frac{\beta_{ji}(\lambda)}{1+ (-1)^{\sigma(i)+\sigma(j)}\beta_{ji}(\lambda)}
\end{equation*}
and using the description of $\alpha_{ij}(\lambda)$ in terms of the $\beta_{ij}(\lambda)$ given above, we see that solutions $\beta_{ij}(\lambda), \beta_{ji}(\lambda)$ to the above equations will also be solutions to Equation \eqref{E:coeffkii} (cf.\ \cite[Lemma 1.4]{EV2}.
Furthermore defining
\begin{equation*}
\mu_{ij}(\lambda) = \lambda_{ij} - \frac{(-1)^{\sigma(i)}}{\beta_{ij}(\lambda)}
\end{equation*}
we can show that $\mu_{ij}(\lambda-\omega_i) = \mu_{ij}(\lambda-\omega_j) = \mu_{ij}(\lambda)$ for all $i \neq j$.
Thus the meromorphic functions
\begin{equation*}
\beta_{ij}(\lambda) = \frac{(-1)^{\sigma(i)}}{\lambda_{ij} - \mu_{ij}(\lambda)} \text{ and }
\beta_{ji}(\lambda) = \frac{(-1)^{\sigma(j)}}{\lambda_{ji} - \mu_{ji}(\lambda)}
\end{equation*}
where $\mu_{ij}(\lambda)=-\mu_{ji}(\lambda)$ and $\mu_{ij}(\lambda)$ is a meromorphic function periodic with respect to shifts of $\lambda$ by $\omega_i$ and $\omega_j$ will be solutions to Equations \eqref{E:coeffkii} and \eqref{E:coeffiki} (cf.\ \cite[Lemma 1.4]{EV2}).
Note that Equations \eqref{E:coeffikj} and \eqref{E:coeffjik} imply that the function $\beta_{ij}(\lambda)$ is periodic with respect to shifts of $\lambda$ by $\omega_k$ for all $k$ distinct from $i$ and $j$. This periodicity then holds also for $\mu_{ij}(\lambda)$.
Next look at Equation \eqref{E:coeffjki} on functions $\beta_{ij}(\lambda)$, $\beta_{jk}(\lambda)$, $\beta_{ik}(\lambda)$, we note that if any one of these is identically zero, then at least one more has to be identically zero. This allows us to define an equivalence relation on the indices $\{1, 2, \cdots, N\}$: First assert that all $i$ are related to themselves. Then for $i \neq j$ let $i$ be related to $j$ if $\beta_{ij}(\lambda)$ is not identically zero. The symmetry property follows directly from the trace condition.
For the equivalence relation defined above, let $Y = Y_1 \cup Y_2 \cup \cdots \cup Y_n$ be the union of all $n$ equivalence classes $Y_i$ with more than one element. If pairwise distinct $i,j,k$ do not all belong in the same equivalence class, then at least two of $\beta_{ij}$, $\beta_{jk}$, $\beta_{ik}$ will be identically zero, thus the triple will be consistent with Equation \eqref{E:coeffjki}. If all three lie in the same equivalence class, then we get:
\begin{equation*}
(-1)^{\sigma(k)}\beta_{kj}(\lambda-\omega_i)\beta_{ik}(\lambda) + (-1)^{\sigma(j)}\beta_{ij}(\lambda) \beta_{jk}(\lambda-\omega_i) = (-1)^{\sigma(i)} \beta_{ik}(\lambda)\beta_{ij}(\lambda),
\end{equation*}
and by periodicity of $\beta_{kj}$ and $\beta_{jk}$ with respect to $\omega_i$, we reduce this further to:
\begin{equation*}
(-1)^{\sigma(k)}\beta_{kj}(\lambda)\beta_{ik}(\lambda) + (-1)^{\sigma(j)}\beta_{ij}(\lambda)\beta_{jk}(\lambda) = (-1)^{\sigma(i)} \beta_{ik}(\lambda)\beta_{ij}(\lambda).
\end{equation*}
We can rewrite this as:
\begin{align*}
(-1)^{\sigma(k)}\left ( \frac{(-1)^{\sigma(k)}}{\lambda_{kj} - \mu_{kj}(\lambda)}\right ) \left ( \frac{(-1)^{\sigma(i)}}{\lambda_{ik} - \mu_{ik}(\lambda)}\right )
&+ (-1)^{\sigma(j)}\left ( \frac{(-1)^{\sigma(i)}}{\lambda_{ij} - \mu_{ij}(\lambda)} \right ) \left ( \frac{(-1)^{\sigma(j)}}{\lambda_{jk} - \mu_{jk}(\lambda)} \right ) \\
&= (-1)^{\sigma(i)} \left ( \frac{(-1)^{\sigma(i)}}{\lambda_{ik} - \mu_{ik}(\lambda)}\right ) \left ( \frac{(-1)^{\sigma(i)}}{\lambda_{ij} - \mu_{ij}(\lambda)}\right ),
\end{align*}
which, after sign cancelations, yields $\mu_{ik}(\lambda) = \mu_{ij}(\lambda)+\mu_{jk}(\lambda)$. Therefore as in the non-graded case of \cite[\S\S1.11]{EV2}, we conclude that there exists a $1$-quasiconstant meromorphic function $\mu : \mathfrak{h}^* \rightarrow \mathfrak{h}^*$ such that $\mu_{ij}(\lambda) = x_i(\mu(\lambda)) - x_j(\mu(\lambda))$ for all $i,j$ with $\mu_{ij}$ not identically zero, and thus
Equation \eqref{E:coeffkji} is also satisfied.
Let $\tau$ be a permutation of $\{1, \dots, N\}$ that transforms the set $Y$ into a set $X$ which can now be written as a disjoint union of subintervals $X = X_1 \sqcup X_2 \sqcup \cdots \sqcup X_n$. In other words, every subinterval $X_k$ should be of the form $X_k = [i_k,i_{k+1}, i_{k+2}, \cdots, j_k]$, and $j_k < i_{k+1}$ for each $k$. Finally applying a gauge transformation of type (2) for this $\tau$ to the $R$-matrix $R$ will yield an $R$-matrix of the form desired. This completes the proof of Theorem \ref{T:ClassifyR1}. \qed
\section{Conclusion}
\label{S:Conclusion}
In this note we proved a quantization theorem for super dynamical $r$-matrices. More specifically we explicitly constructed quantizations for zero weight super dynamical $r$-matrices with zero coupling constant. We expect that the definitions and constructions here will also be helpful in the proof of an analogous quantization result for nonzero coupling constants, we plan to follow up on this thread in future work.
It must be clear that quantization in this note meant finding a solution to the quantum dynamical Yang-Baxter equation whose semi-classical limit was the original super dynamical $r$-matrix. In particular we have not explicitly constructed algebraic structures which should be the corresponding dynamical quantum groups associated to the resulting $R$-matrices. However, while working in the quantum picture, we have proposed and used a particular algebraic condition which we called {\em the super Hecke condition} (cf.\ Subsection \ref{SS:SuperHecke}). Finding the correct super Hecke condition is important because the Hecke condition in the non-graded case turns out to be the right pre-condition for a meaningful description of dynamical quantum groups in the language of Hopf algebroids (cf.\ \cite{EV2}).
Studying the proof of our main classification result for super dynamical $R$-matrices (Theorem \ref{T:ClassifyR1}), one can see that the building blocks fall into their right places when one defines the super Hecke condition as we do. In this framework, the super dynamical $R$-matrices with equal Hecke parameters correspond precisely to the zero weight super dynamical $r$-matrices with zero coupling constant. This is exactly analogous to the non-graded picture in \cite{EV2}. This observation may offer some support for our particular definition of the super Hecke condition.
The construction of the actual algebraic structures that correspond to the super dynamical $R$-matrices we study in Section \ref{S:QuantPict} involves the difficult problem of determining what the appropriate super analogue to dynamical quantum groups should be. This is beyond the scope of this note, but we believe that our work here will shed some light to it by contributing some evidence for the right way to superize the Hecke condition. We intend to address this issue in depth in our followup work. For various possible approaches to the theory of super dynamical quantum groups and some preliminary results, see \cite{Kar4, Kar5}.
|
\section{Introduction}
\noindent Field theories in (1+2)-dimensions often provide easy ways
to check ideas that are difficult to prove in actual
(1+3)-dimensions. However, it is not surprising to uncover other new
ideas as well that are specific to (1+2)-dimensions. Topologically
massive gravity \cite{deser} or BTZ black holes \cite{teitelboim}
are some of the best known examples to the latter. Other aspects
and the literature may be found in Ref.\cite{carlip}. In this paper
we couple a Dirac spinor in (1+2)-dimensions to Einstein-Cartan
gravity \cite{trautman}. The field equations are derived by a
variational principle. Then a family of circularly symmetric,
rotating solutions that are asymptotically $AdS_3$ are found and
discussed.
\bigskip
\noindent We specify the space-time geometry by a triplet $ \left (
M, g, \nabla \right )$ where $M$ is a 3-dimensional differentiable
manifold equipped with a metric tensor \begin{equation} g = \eta_{ab} e^a \otimes
e^b \end{equation} of signature $(-++)$. $\{e^a\}$ is an orthonormal co-frame
dual to the frame vectors $\{X_a\}$, that is $e^{a}(X_b) =
\delta^{a}_{b}$. A metric compatible connection $\nabla$ can be
specified in terms of connection 1-forms $\{\omega^{a}_{\;\;b}\}$
satisfying $\omega_{ba} = -\omega_{ab}$. Then the Cartan structure
equations \begin{equation} de^a + \omega^{a}_{\;\; b} \wedge e^b = T^a , \end{equation} \begin{equation}
d\omega^{a}_{\;\; b} + \omega^{a}_{\;\; c} \wedge \omega^{c}_{\;\; b} =
R^{a}_{\;\; b} \end{equation} yield the space-time torsion 2-forms $\{T^a\}$
and curvature 2-forms $\{R^{a}_{\;\;b}\}$, respectively. Here $d$
denotes the exterior derivative and $\wedge$ the wedge product. We fix
the orientation of space-time by choosing the volume 3-form $*1 =
e^0 \wedge e^1 \wedge e^3$ where $*$ is the Hodge star map. It is
possible to decompose the connection 1-forms in a unique way as \begin{equation}
\omega^{a}_{\;\; b} = \hat{\omega}^{a}_{\;\; b} +K^{a}_{\;\; b} \end{equation}
where $\{\hat{\omega}^{a}_{\;\; b} \}$ are the zero-torsion
Levi-Civita connection 1-forms satisfying \begin{equation} de^a +
\hat{\omega}^{a}_{\;\; b} \wedge e^b = 0 \end{equation} and $\{K^{a}_{\;\; b} \}$
are the contortion 1-forms satisfying \begin{equation} K^{a}_{\;\; b} \wedge e^b =
T^a .\end{equation} The curvature 2-forms are also decomposed in a similar way:
\begin{equation} R^{a}_{\;\;b} = \hat{R}^{a}_{\;\;b} + \hat{D}K^{a}_{\;\;b} +
K^{a}_{\;\;c} \wedge K^{c}_{\;\;b} \end{equation} with
$$
\hat{D}K^{a}_{\;\;b} = dK^{a}_{\;\;b} + \hat{\omega}^{a}_{\;\;c} \wedge
K^{c}_{\;\;b} - \hat{\omega}^{c}_{\;\;b} \wedge K^{a}_{\;\;c} .
$$
\medskip
\noindent The field equations of Einstein-Cartan theory of gravity
\cite{trautman} are obtained by varying the action \begin{equation} I = \int_M
\left ( {\cal{L}}_{EC} + {\cal{L}}_{M} \right ) \end{equation} where the
Einstein-Cartan Lagrangian density 3-form \begin{equation} {\cal{L}}_{EC} =
-\frac{1}{2 \kappa} R_{ab} \wedge *(e^a \wedge e^b) + \lambda *1 , \end{equation} with
the gravitational constant $\kappa$ and the cosmological constant
$\lambda$, and the matter Lagrangian density 3-form ${\cal{L}}_{M}$.
We write the infinitesimal variations (up to a closed form) as \begin{equation}
\dot{\cal{L}} = \dot{e}^a \wedge \left ( -\frac{1}{2 \kappa} R^{bc}
*e_{abc} + \lambda
*e_a + \tau_a \right) + \frac{1}{2} \dot{\omega}^{ab} \wedge \left ( -\frac{1}{\kappa}*e_{abc} T^c +
\Sigma_{ab} \right ) \end{equation} where the variations of the matter
Lagrangian yield the stress-energy 2-forms \begin{equation} \tau_a =
\frac{\partial {\cal{L}_{M}}}{\partial e^a} = T_{ab} *e^b \end{equation} and
the angular momentum 2-forms \begin{equation} \Sigma_{ab} = \frac{\partial
{\cal{L}_{M}}}{\partial \omega^{ab}} = S_{ab,c} *e^c . \end{equation} Therefore
the Einstein-Cartan field equations are given either as \begin{equation}
\frac{1}{2} R^{bc}
*e_{abc} - \kappa \lambda
*e_a = \kappa \tau_a , \end{equation}
\begin{equation} *e_{abc} T^c = \kappa \Sigma_{ab} \end{equation}
or equivalently as \begin{equation} R_{ab} =
\kappa \lambda ( e_a \wedge e_b) + \kappa *e_{abc} \tau^c \label{einstein} , \end{equation} \begin{equation} T_a = \frac{1}{2} \kappa *e_{abc} \Sigma^{bc} .
\label{cartan} \end{equation}
\section{Einstein-Cartan-Dirac Field Equations}
\noindent Let us consider a Dirac spinor field \begin{equation} \psi = \left (
\begin{array}{c} \psi_1 \\ \psi_2 \end{array}
\right ) \end{equation} and the conjugate spinor field \begin{equation} \bar{\psi} =
\psi^{\dagger} \gamma_0 = \left ( -\psi^{*}_2 \; \; \psi^{*}_1
\right ) \end{equation} where $\psi_1$ and $\psi_2$ are complex, odd Grassmann
valued functions. We use a real (i.e. Majorana) realization of the
gamma matrices $\{\gamma_a\}$ given explicitly as \begin{equation} \gamma_0 =
\left (
\begin{array}{cc} 0 & 1\\ -1 & 0 \end{array} \right ) \; ,
\;\gamma_1 = \left ( \begin{array}{cc} 0 & 1\\ 1 & 0
\end{array} \right ) \; , \;\gamma_2 = \left ( \begin{array}{cc} 1 & 0\\0& -1 \end{array}
\right ) . \end{equation} The exterior covariant derivative of the spinor
fields are defined to be \begin{equation} \nabla \psi = d \psi + \frac{1}{2}
\omega^{ab} \sigma_{ab} \psi \quad , \quad \overline{\nabla \psi} =
d \bar{\psi} -\frac{1}{2} \omega^{ab} \bar{\psi} \sigma_{ab} \end{equation}
with \begin{equation} \sigma_{ab} = \frac{1}{4} \left [ \gamma_{a} ,\gamma_{b}
\right ] = \frac{1}{2}
*e_{abc} \gamma^c . \end{equation}
We set $\gamma = \gamma_{a} e^{a} .$
\medskip
\noindent Next let us introduce the (Hermitian) Dirac Lagrangian
density 3-form \begin{equation} {\cal{L}}_D = \frac{i}{2} \left ( \bar{\psi}
*\gamma \wedge \nabla \psi - \overline{\nabla \psi} \wedge *\gamma \psi
\right ) + i m \bar{\psi} \psi
*1 . \end{equation}
Its infinitesimal variations are found to be (up to a closed form)
\begin{eqnarray} {\dot{\cal{L}}}_D &=& \dot{e}^a \wedge \left \{ \frac{i}{2}
*e^{b}_{\; \; a} \wedge ( \bar{\psi} \gamma_b \nabla \psi + \overline{\nabla \psi} \gamma_{b} \psi) + i m \bar{\psi} \psi *e_a \right
\} \nonumber \\ & & + \frac{1}{2} \dot{\omega}^{ab} \wedge \left \{
\frac{i}{2} \bar{\psi} (
*\gamma \sigma_{ab} + \sigma_{ab} *\gamma ) \psi
\right \} \nonumber \\
& & + i \dot{\bar{\psi}} \left \{ *\gamma \wedge \nabla \psi +
\frac{1}{2}
*e^{a}_{\;\; b} \wedge T^b \gamma_a \psi + m * \psi \right \} \nonumber \\
& & - i \left \{ \overline{\nabla \psi} \wedge *\gamma - \frac{1}{2}
*e^{a}_{\;\;b} \wedge T^b \bar{\psi} \gamma_a - m * \bar{\psi} \right \} \dot{\psi} .
\end{eqnarray}
From the above expression we identify the stress-energy 2-forms \begin{equation}
\tau_a = \frac{i}{2} *e^{b}_{\;\;a} \left ( \bar{\psi} \gamma_b
\nabla \psi + \overline{\nabla \psi} \gamma_b \psi \right ) + i m
\bar{\psi} \psi *e_a \end{equation} and the angular momentum 2-forms \begin{equation}
\Sigma_{ab} = \frac{i}{2} \left ( \bar{\psi} *\gamma \sigma_{ab}
\psi + \bar{\psi} \sigma_{ab} *\gamma \psi \right ) .\end{equation}
Substituting these into the Einstein-Cartan equations
(\ref{einstein}) and {\ref{cartan}) we obtain
\begin{eqnarray}
R_{ab} &=& \kappa \lambda e_a \wedge e_b + i m \kappa \bar{\psi} \psi
e_a \wedge e_b \nonumber \\
& & + i \frac{\kappa}{2} e_a \wedge ( \bar{\psi} \gamma_b \nabla \psi )
- i \frac{\kappa}{2} e_b \wedge ( \bar{\psi} \gamma_a \nabla \psi )
\nonumber \\ & & - i \frac{\kappa}{2} e_b \wedge ( \overline{\nabla
\psi} \gamma_a \psi ) + i \frac{\kappa}{2} e_a \wedge ( \bar{\nabla
\psi} \gamma_b \psi ) , \label{curvature}
\end{eqnarray}
\begin{equation} T_a = i \frac{\kappa}{2} \bar{\psi} \psi *e_a \label{torsion}
\end{equation} and the Dirac equation \begin{equation}
*\gamma \wedge \nabla \psi + m \psi *1 = 0 . \label{dirac}
\end{equation} We note that with the torsion 2-forms (\ref{torsion}),
$*e^{a}_{\;\;b} \wedge T^b = 0$ identically and the Dirac equation
simplifies to (\ref{dirac}).
\medskip
\noindent These field equations can be re-written in terms of the
Levi-Civita connection only. To see this, we solve (\ref{torsion})
for the contortion 1-forms \begin{equation} K_{ab} = -i \frac{\kappa}{4}
\bar{\psi} \psi
*(e_a \wedge e_b) \end{equation} and substitute these into (\ref{curvature}) which
simplify to \begin{equation} \hat{R}_{ab} = \kappa \lambda e_a \wedge e_b + *e_{abc}
{\hat{\tau}}^c -i \frac{\kappa}{4} d( \bar{\psi} \psi ) \wedge *(e_a \wedge
e_b) + \frac{3 \kappa^2}{16} (\bar{\psi} \psi )^2 e_a \wedge e_b .
\label{einstein2}\end{equation} The Dirac equation (\ref{dirac}) similarly
simplifies to \begin{equation}
*\gamma \wedge \bar{\nabla} \psi + m \psi *1 -i \frac{3 \kappa}{8} (\bar{\psi} \psi ) \psi *1 =
0 \label{dirac2} \end{equation}
\medskip
\noindent It is interesting to note that the Einstein-Dirac
equations (\ref{einstein2}) and (\ref{dirac2}) can be obtained from
an action by zero-torsion constrained variations using the method
of Lagrange multipliers \cite{dereli1988}. To this end, we consider
a modified Dirac Lagrangian density 3-form \begin{equation} {\cal{L}}_D^{\prime}
= \frac{i}{2} \left ( \bar{\psi}
*\gamma \wedge \hat{\nabla} \psi - \overline{{\hat{\nabla}} \psi} \wedge *\gamma \psi
\right ) + i m \bar{\psi} \psi *1 + \frac{3 \kappa^2}{16} (
\bar{\psi} \psi )^2
*1 \end{equation}
together with the constraint term \begin{equation} {\cal{L}}_{constraint} = (
de^a + \omega^{a}_{\;\;b} \wedge e^b ) \wedge \lambda_a \end{equation} where
$\{\lambda_a\}$ are the Lagrange multiplier 1-forms. The variation
of the total action
$$
I = \int_M \left ( {\cal{L}}_{EC} + {\cal{L}}_D^{\prime} +
{\cal{L}}_{constraint} \right )
$$
with respect to the Lagrange multipliers imposes the constraint that
the connection 1-forms are Levi-Civita. Then we solve the connection
variation equations under this constraint for the Lagrange
multiplier 1-forms as
$$
\lambda_a = -\frac{i}{4} \kappa \bar{\psi} \psi e_a .
$$
Substituting these in the remaining Einstein and Dirac equations
give precisely (\ref{einstein2}) and (\ref{dirac2}).
\section{A Circularly Symmetric Solution}
\noindent We consider the metric \begin{equation} g = - f^2(r) dt^2 + h^2(r)
dr^2 + r^2 ( d \phi + a(r) dt )^2 \end{equation} in plane polar coordinates
$(t, r, \phi)$ \cite{dereli2000}. The following choice of
orthonormal basis 1-forms \begin{equation} e^0 = f(r) dt \quad , \quad e^1 = h(r)
dr \quad , \quad e^2 = r ( d \pi + a(r) dt ) , \end{equation} leads to the
Levi-Civita connection 1-forms \begin{equation} \hat{\omega}^{0}_{\; \;1} =
\alpha e^0 - \frac{\beta}{2} e^2 \; , \; \hat{\omega}^{0}_{\; \; 2}
= -\frac{\beta}{2} e^1 \; , \; \hat{\omega}^{1}_{\; \; 2} = -\gamma
e^2 -\frac{\beta}{2} e^0 \end{equation} where we defined \begin{equation} \alpha =
\frac{f^{\prime}}{f h } \quad , \quad \beta = \frac{r a^{\prime}}{f
h } \quad , \quad \gamma = \frac{1}{r h} \end{equation} with ${ }^\prime$
denoting the derivative $\frac{d}{dr}$. On the other hand, assuming
$ i \frac{\kappa}{2} \bar{\psi} \psi = \tau(r)$, we calculate the
contortion 1-forms \begin{equation} K^{0}_{\; \;1} = \frac{\tau}{2} e^2 \; , \;
K^{0}_{\; \; 2} = -\frac{\tau}{2} e^1 \; , \; K^{1}_{\; \; 2} =
-\frac{\tau}{2} e^0 . \end{equation} As a first step towards a solution, we
take a Dirac spinor that depends only on $r$ and work out
(\ref{dirac}) in components as follows:
\begin{eqnarray}
\psi^{\prime}_{1} + \frac{h}{2} ( \alpha + \gamma ) \psi_1 +
\frac{h}{4} ( \beta + 3 \tau + 4m) \psi_2 &=& 0 , \nonumber \\
\psi^{\prime}_{2} + \frac{h}{2} ( \alpha + \gamma ) \psi_2 +
\frac{h}{4} ( \beta + 3 \tau + 4m) \psi_1 &=& 0 .
\end{eqnarray}
We take the combinations $\psi_{+} = \psi_1 + \psi_2$ and $\psi_{-}
= \psi_1 - \psi_2$ and write a decoupled system of equations
\begin{eqnarray}
\psi^{\prime}_{+} + (k_1 + k_2 ) \psi_{+} &=& 0 \nonumber \\
\psi^{\prime}_{-} + (k_1 - k_2 ) \psi_{-} &=& 0
\end{eqnarray}
where we set
$$
k_1 = \frac{h}{2} ( \alpha + \gamma ) \quad , \quad k_2 =
\frac{h}{4} ( \beta + 3 \tau + 4m).
$$
The formal solution to these equations are given by
\begin{eqnarray}
\psi_1 &=& e^{-\int^r k_1 dr} \left ( \xi_1 e^{-\int^r k_2 dr} +
\xi_2 e^{\int^r k_2 dr} \right ) \nonumber \\ \psi_2 &=&
e^{-\int^r k_1 dr} \left ( \xi_1 e^{\int^r k_2 dr} - \xi_2
e^{-\int^r k_2 dr} \right )
\end{eqnarray}
where $\xi_1$ and $\xi_2$ are complex, odd Grassmann valued
constants. It can easily be verified \begin{equation} \tau(r) = i \kappa (
\xi^{*}_2 \xi_1 - \xi^{*}_1 \xi_2 ) e^{-2 \int^r k_1 dr} . \end{equation}
\medskip
\noindent We next work out the Einstein field equations
(\ref{einstein}) which after simplifications reduce to the following
system of coupled first order differential equations:
\begin{eqnarray}
\frac{\beta^{\prime}}{2 h} + \beta \gamma & = & -
\frac{\tau^{\prime}}{2 h} - \tau \alpha \nonumber \\
\frac{\gamma^{\prime}}{h} + \frac{\beta^2}{4} + \gamma^2 + \kappa
\lambda &=& \frac{3 \tau^2}{4} + \frac{\beta \tau}{2}
\nonumber \\
\frac{\alpha^{\prime}}{h} - \frac{3\beta^2}{4} + \alpha^2 + \kappa
\lambda &=& \frac{3 \tau^2}{4} - \frac{\beta \tau}{2}
\nonumber \\
\frac{\beta^2}{4} + \alpha \gamma + \kappa \lambda & =& -\frac{3
\tau^2}{4} - 2 m \tau \nonumber \\
\frac{\beta^{\prime}}{2 h} + \beta \gamma & = &
\frac{\tau^{\prime}}{2 h} + \tau \gamma . \label{equations}
\end{eqnarray}
At this point we choose $$ \kappa \lambda = - \frac{1}{l^2} < 0 $$
and restrict our attention to those cases for which
$$ \gamma = \alpha = \frac{1}{r h} \quad , \quad \tau = \beta =
\frac{\beta_0}{r^2} .
$$
We can then integrate (\ref{equations}) for the metric functions \begin{equation}
f(r) = \frac{r}{l} \quad , \quad h(r) = \frac{l}{r \sqrt{ 1 -
\frac{2 m \beta_0 l^2}{r^2} - \frac{\beta^2_0 l^4}{r^4}}},\end{equation} and
\begin{equation} a(r) = \frac{1}{2l} \arcsin \left ( \frac{m}{\sqrt{m^2 +
\frac{1}{l^2}}} \right ) - \frac{1}{2l} \arcsin \left ( \frac{m +
\frac{\beta_0}{r^2}}{\sqrt{m^2 + \frac{1}{l^2}}} \right ) . \end{equation} It
now remains to integrate for the Dirac spinor and we find \begin{equation}
\psi_{\pm} = \frac{1}{r} \left |r^2 - m \beta_0 l^2 + \sqrt{r^4 - 2
m \beta_0 l^2 r^2 - \beta^2_0 l^4 }\right |^{\pm \frac{ml}{2}}
e^{\mp \frac{1}{2 \beta_0} \arcsin \left ( \frac{m +
\frac{\beta_0}{r^2}}{\sqrt{m^2 + \frac{1}{l^2}}} \right ) }
\xi_{\pm} \end{equation} where $\xi_{\pm} = \xi_1 \pm \xi_2 .$
\medskip
\noindent In order to understand the physical meaning of this
solution we write down the metric \begin{equation} g = - \frac{r^2}{l^2} dt^2 +
\frac{l^2 dr^2}{r^2 \left ( 1 -\frac{2 m \beta_0 l^2}{r^2} -
\frac{\beta_0^2 l^2}{r^4}\right ) } + r^2 (d\phi + a(r) dt)^2 . \end{equation}
We see that in the absence of a Dirac spinor ($\beta_0 = 0$) the
above metric collapses to the $AdS_3$ metric \begin{equation} g_0 =
-\frac{r^2}{l^2} dt^2 + \frac{l^2}{r^2} dr^2 + r^2 d\phi^2 . \end{equation}
Even when $\beta_0 \neq 0$, the metric $g \rightarrow g_0$
asymptotically as $r \rightarrow \infty $. We note a metric
singularity at \begin{equation} r_c = \left \{
\begin{array}{lr} l \sqrt{m \beta_0 + \beta_0 \sqrt{m^2 +
\frac{1}{l^ 2}}} \quad & , \quad \beta_0 > 0 \\
l \sqrt{|\beta_0| \sqrt{m^2 + \frac{1}{l^ 2}}- m |\beta_0|} \quad &
, \quad \beta_0 < 0 \end{array} \right . \end{equation} We further calculate
the curvature scalar \begin{equation} {\cal{R}} = -\frac{6}{l^2} + \frac{4 m
\beta_0}{r^2} \end{equation} and the quadratic curvature invariant \begin{equation} R_{ab}
\wedge *R^{ab} = \frac{6}{l^4} -\frac{8 m \beta_0}{l^2 r^2} +
\frac{\beta_0^2 (8m^2-\frac{4}{l^2})}{r^4} + \frac{16 m
\beta_0^2}{r^6} + \frac{8 \beta_0^4}{r^8} . \end{equation} The curvature
invariants have an essential singularity at $r = 0$. We also check
(See Ref.\cite{dereli2000}) the quasi-local angular momentum \begin{equation}
J(r) = \frac{r^3}{f(r) h(r)} \frac{da}{dr} = \beta_0 , \end{equation} the
quasi-local energy \begin{equation} E(r) = \frac{1}{h_0(r)} - \frac{1}{h(r)} =
\frac{r}{l} - \sqrt{\frac{r^2}{l^2} - 2 m \beta_0 -
\frac{\beta_0^2}{r^2}} \simeq \frac{\beta_0 m l}{r} ,\end{equation} and the
quasi-local mass \begin{eqnarray} M(r) & =& 2 f(r) E(r) = J(r)a(r)
\nonumber \\ &=& 2\frac{l^2}{r^2} - 2\frac{l^2}{r^2} \sqrt{1
-\frac{2 m \beta_0 l^2}{r^2} - \frac{\beta_0^2 l^2}{r^4}} \nonumber
\\ & & + \frac{1}{2l} \arcsin \left ( \frac{m}{\sqrt{m^2 +
\frac{1}{l^2}}} \right ) - \frac{1}{2l} \arcsin \left ( \frac{m +
\frac{\beta_0}{r^2}}{\sqrt{m^2 + \frac{1}{l^2}}} \right ) \\
&\simeq& 2 m \beta_0 . \nonumber
\end{eqnarray}
\newpage
\section{Conclusion}
\noindent We have formulated the Einstein-Cartan-Dirac theory in
(1+2)-dimensions using the algebra of exterior differential forms.
We coupled a Dirac spinor to Einstein-Cartan gravity and obtained
the field equations by a variational principle. The space-time
torsion is given algebraically in terms of the quadratic spinor
invariant. We then looked for rotating, circularly symmetric
solutions. We found a particular class of solutions that possesses
an essential curvature singularity at the origin $r=0$ that is
hidden behind an event horizon at some finite distance $r=r_c$ away
from the origin. The mass and the spin of this configuration can be
identified. Thus the space-time geometry we found exhibits all the
essential features of a black hole and we find it interesting that
in the absence of the Dirac spinor field collapses to the regular
$AdS_3$ geometry.
\bigskip
\noindent {\bf {\Large Acknowledgement}}
\medskip
\noindent TD is supported in part by the Turkish Academy of Sciences
(T\"{U}BA).
\bigskip
|
\section*{Introduction}
In the last years, fuzzy logics and fuzzy set theory have been widely applied to image processing tasks. In particular, the theory of \emph{fuzzy relation equations}, deeply investigated in~\cite{dinolasessa}, is involved in many algorithms for compression and reconstruction of digital images (see, for example,~\cite{dinolarusso,hirotapedrycz,nobuharatakama}).
In such techniques, however, the approach is mainly experimental and the algebraic context is seldom clearly defined. Basically, most of the fuzzy algorithms for image compression, make use of join-product operators; after all, a complete lattice order and a multiplication that is residuated w.r.t. the lattice-order are the fundamental ingredients of these operators.
On the other hand, there exists another class of operators acting on digital images that, although having a completely different scope, has the same algebraic form: \emph{mathematical morphological operators}. Mathematical Morphology is a technique for image processing and analysis whose birth can be traced back to the book~\cite{matheron} by G. Matheron and whose establishment is due mainly to the work of Heijmans and Serra (e.g. \cite{goutsiasheijmans,heijmans,serra}). The basic problem in mathematical morphology is to design nonlinear operators that extract relevant topological or geometric information from images. This requires the development of a mathematical model for images and a rigorous theory that describes fundamental properties of the desirable image operators.
Essentially, mathematical morphological operators analyse the objects in an image by ``probing'' them with a small geometric ``model-shape'' (e.g., line segment, disc, square) called the \emph{structuring element}. These operators are defined on spaces having both a complete lattice order (set or fuzzy set inclusion, in concrete applications) and an external action from another ordered structure (the set of translations); more, they are usually coupled in adjoint pairs.
Regarding these different classes operators, what is really outstanding from an algebraic point of view is the fact that they can both be expressed in terms of a complete lattice order and a residuated product. Our aim is to show how all these operators can be joined together in a common mathematical context: the categories of \emph{quantale modules} and the operators called \emph{$\mathcal{Q}$-module transforms}. We will also show that such operators
\begin{itemize}
\item are precisely the $\mathcal{Q}$-module homomorphisms between free modules,
\item are completely determined by the mathematical counterpart of the coder (for compression algorithms) and of the structuring element (in the case of mathematical morphology).
\end{itemize}
Throughout the paper, due to space constraint, we will omit all the proofs of propositions and theorems; however they can all be found in \cite{russo}.
\section{Preliminaries}
\label{prel}
In this section we will briefly recall some basic notions on several ordered algebraic structures.
\begin{definition}\label{residuated map}
Let $\langle X, \leq \rangle$ and $\langle Y, \leq \rangle$ be two posets. A map $f: X \longrightarrow Y$ is said to be \emph{residuated} iff there exists a map $g: Y \longrightarrow X$ such that, for all $x \in X$ and for all $y \in Y$, $f(x) \leq y \ \iff \ x \leq g(y)$.
It is immediate to verify that the map $g$ is uniquely determined; we will call it the \emph{residual map} or the \emph{residuum} of $f$, and denote it by $f_*$. The pair $(f, f_*)$ is called \emph{adjoint}; a residuated map preserves all existing joins and its residuum preserves all existing meets.
\end{definition}
The category $\mathcal{SL}$ of sup-lattices is the one whose objects are complete lattices and morphisms are maps preserving arbitrary joins or, that is the same, residuated maps. For a sup-lattice $\mathbf{L}$, we will use the notation $\mathbf{L} = \langle L, \vee, \bot \rangle$. For any sup-lattice $\mathbf{L} = \langle L, \vee, \bot \rangle$, it is possible to define a dual sup-lattice in an obvious way: if we consider the opposite partial order $\geq$, then $\mathbf{L}^{\operatorname{op}} = \langle L, \wedge, \top \rangle$ is a sup-lattice and, clearly, $(\mathbf{L}^{\operatorname{op}})^{\operatorname{op}} = \mathbf{L}$.
\begin{proposition}\label{residprop}
Let $\langle X, \leq \rangle$ and $\langle Y, \leq \rangle$ be posets, and let $(f,f_*)$ be an adjoint pair, with $f: X \longrightarrow Y$. Then the following hold:
\begin{enumerate}
\item[$(i)$]$f$ is surjective \quad $\iff$ \quad $f_*$ is injective \quad $\iff$ \quad $f \circ f_* = \operatorname{id}_Y$;
\item[$(ii)$]$f$ is injective \quad $\iff$ \quad $f_*$ is surjective \quad $\iff$ \quad $f_* \circ f = \operatorname{id}_X$.
\end{enumerate}
\end{proposition}
A binary operation $\cdot$ on a partially ordered set $\langle P, \leq \rangle$ is said to be \emph{residuated} iff there exist binary operations $\backslash$ and $/$ on $P$ such that for all $x, y, z \in P$, $x \cdot y \leq z \ \textrm{iff} \ x \leq z/y \ \textrm{iff} \ y \leq x \backslash z$. The operations $\backslash$ and $/$ are referred to as the left and right \emph{residua} of $\cdot$, respectively. In other words, a residuated binary operation over $\langle P, \leq \rangle$ is a map from $P \times P$ to $P$ that is residuated in both arguments. In the situations where $\cdot$ is a monoid operation with a unit element $e$ and the partial order is a lattice order, we can add the monoid unit and the lattice operations to the similarity type to get an algebraic structure $\mathbf{R} = \langle R, \vee, \wedge, \cdot, \backslash, /, e \rangle$ called \emph{residuated lattice}.
In the category $\mathcal{Q}$ of quantales, $\operatorname{Obj}(\mathcal{Q})$ is the class of complete residuated lattices and the morphisms are the maps preserving products, the unit, arbitrary joins and the bottom element. An alternative, yet equivalent, definition of quantale is the following
\begin{definition}\label{quantale}
A \emph{quantale}\index{Quantale} is an algebraic structure $\mathbf{Q} = \langle Q, \vee, \cdot, \bot, e \rangle$ such that
\begin{enumerate}
\item[$(Q1)$]$\langle Q, \vee, \bot \rangle$ is a sup-lattice,
\item[$(Q2)$]$\langle Q, \cdot, e \rangle$ is a monoid,
\item[$(Q3)$]$x \cdot \bigvee\limits_{i \in I} y_i = \bigvee\limits_{i \in I} (x \cdot y_i)$ \ and \ $\left(\bigvee\limits_{i \in I} y_i\right) \cdot x = \bigvee\limits_{i \in I} (y_i \cdot x)$ \ for all $x \in Q$, $\{y_i\}_{i \in I} \subseteq Q$.
\end{enumerate}
$\mathbf{Q}$ is said to be \emph{commutative} if so is the multiplication. Obviously, if $\mathbf{Q}$ is commutative then the two residua coincide and $x/y = y \backslash x$ is denoted by $y \to x$.
\end{definition}
Before giving examples of quantale structures interesting for the scope of this paper, we recall that a binary operation $\ast: [0,1]^2 \longrightarrow [0,1]$ is called a \emph{triangular norm}, \emph{t-norm} for short, provided it is associative, commutative, monotone in both arguments and has $1$ as the neutral element. A t-norm $\ast$ is called \emph{left-continuous} if, for all $\{x_n\}_{n \in \mathbb{N}}, \{y_n\}_{n \in \mathbb{N}} \in [0,1]^\mathbb{N}$,
\begin{center}
$\left(\bigvee_{n \in \mathbb{N}} x_n\right) \ast \left(\bigvee_{n \in \mathbb{N}} y_n\right) = \bigvee_{n \in \mathbb{N}} (x_n \ast y_n).$
\end{center}
In this case, clearly, $\ast$ is residuated and its residuum (unique, since $\ast$ is commutative) is given by \ $x \to y = \bigvee\{z \in [0,1] \mid z \ast x \leq y\}$.
\begin{exe}
If $\ast$ is any left-continuous t-norm on the real unit interval, then $\langle [0,1], \vee, \ast, 0, 1 \rangle$ is a commutative quantale.
\end{exe}
\section{Join-product operators in Image Processing}
\label{fuz}
\subsection{Fuzzy algorithms for image compression and reconstruction}
\label{fuzalgsec}
In the literature of image compression, the fuzzy approach is based essentially on the theory of fuzzy relation equations. The underlying idea is the following: a grey-scale image is a matrix in which every element represents a pixel and its value, included in the set $\{0, \ldots, 255\}$ in the case of a 256-bit encoding, is the grey-level. Then, if we normalize the set $\{0, \ldots, 255\}$ by dividing each element by $255$, grey-scale images can be modeled equivalently as fuzzy relations, fuzzy functions (i.e. $[0,1]$-valued maps), fuzzy subsets of a given set or $[0,1]$-valued matrices. A similar model is used for RGB colour images, where each image is represented by three fuzzy relations (respectively: functions, sets or matrices).
So, if we consider a grey-scale image $I$ of sizes $m \times n$ ($m, n \in \mathbb{N}$), we can see it as an $m \times n$ matrix $I_{ij}$ whose values are in $[0,1]$. Now we consider two natural numbers $a \leq m$ and $b \leq n$ and fix a $[0,1]$-valued map in four variables $C \in [0,1]^{m \times n \times a \times b}$ --- usually called \emph{coder} or \emph{codebook}; then we compress the image $I$ into an image $I' = I'_{hk}$ of sizes $a \times b$ by setting
\begin{equation}\label{com}
I'_{hk} = \bigvee_{i,j} I_{ij} \ast C_{ijhk},
\end{equation}
where $\ast$ is any left-continuous t-norm on $[0,1]$. The reconstructed image $I'' = I''_{ij}$ is defined by
\begin{equation}\label{rec}
I''_{ij} = \bigwedge_{h,k} C_{ijhk} \to_\ast I'_{hk},
\end{equation}
where $\to_\ast$ is the residuum of $\ast$.
Even if some fuzzy algorithms for image compression may look different at a first glance, most of them can be rewritten in a form similar to (\ref{com}), with a reconstruction process that will consequently look like (\ref{rec}).
\subsection{Dilation and erosion in Mathematical Morphology}
\label{mathmorphsec}
In Mathematical Morphology binary images are modeled, in the wake of tradition and intuition, as subspaces or subsets of a suitable space $E$, which is assumed to possess some additional structure (topological space, metric space, graph, etc.), usually depending on the kind of task at hand.
Concretely, the class of $n$-dimensional binary images is represented as $\wp(E)$, where $E$ is, in general, $\mathbb{R}^n$ or $\mathbb{Z}^n$. In the first case we have continuous binary images, otherwise we are dealing with discrete binary images. The basic relations and operations between images of this type are essentially those between sets, namely set inclusion, union, intersection and complementation. It is intuitively clear, then, that complete lattices are the algebraic structures required for abstracting the ideas introduced so far.
\begin{definition}\label{dilero}
Let $\mathbf{L}$, $\mathbf{M}$ be complete lattices. A map $\d: L \longrightarrow M$ is called a \emph{dilation} if it distributes over arbitrary joins: $\d\left({}^\mathbf{L}\bigvee X \right) = {}^\mathbf{M}\bigvee \d(X)$, for every $X \subseteq L$. A map $\varepsilon: M \longrightarrow L$ is called an \emph{erosion} if it distributes over arbitrary meets: $\varepsilon\left({}^\mathbf{M}\bigwedge Y \right) = {}^\mathbf{L}\bigwedge \varepsilon(Y)$, for every $Y \subseteq M$.
Two maps $\d: L \longrightarrow M$ and $\varepsilon: M \longrightarrow L$ are said to form an \emph{adjunction}, $(\d,\varepsilon)$, between $\mathbf{L}$ and $\mathbf{M}$ if $\d(x) \leq y \Longleftrightarrow x \leq \varepsilon(y)$, for all $x \in L$ and $y \in M$.\footnote{Notice that the notation used in the literature of Mathematical Morphology is slightly different. Indeed, an adjunction is presented with a dilation in the second coordinate and an erosion in the first. Here we use such a reversed notation because, as we will see, adjunctions are adjoint pairs in the sense of Definition \ref{residuated map}.}
\end{definition}
Assume that $\d: \mathbf{L} \longrightarrow \mathbf{M}$ is a dilation. For $x \in L$, we can write \ $\d(x) = \bigvee_{y \leq x} \d(y)$, \ where we have used the fact that $\d$ distributes over join. Every dilation defined on $\mathbf{L}$ is of the form above, and the adjoint erosion is given by \ $\varepsilon(y) = \bigvee_{\d(x) \leq y} x$.
Now, keeping in mind the models $\mathbb{R}^n$ and $\mathbb{Z}^n$, it is possible to introduce the concepts of \emph{translation} of an image and \emph{translation invariance} of an operator, by means of the algebraic operation of sum. Indeed, let $E$ be $\mathbb{R}^n$ or $\mathbb{Z}^n$ and consider the complete lattice $\wp(E)$; given an element $h \in E$, we define the $h$-\emph{translation} $\tau_h$ on $\wp(E)$ by setting, for all $X \in \wp(E)$, $\tau_h(X) = X + h = \{x + h \mid x \in X\}$, where the sum is intended to be defined coordinatewise.
Next, we consider the case of operators that are translation invariant; here the sets $\d(\{x\})$ are translates of a fixed set, called the \emph{structuring element}, by $\{x\}$. An operator $f: \wp(E) \longrightarrow \wp(E)$ is called \emph{translation invariant}, \emph{T-invariant} for short, if $\tau_h \circ f = f \circ \tau_h$ for all $h \in E$. It is proved in \cite{heijmans} that every T-invariant dilation on $\wp(E)$ is given by $\d_A(X) = \bigcup_{x \in X} A + x$, and every T-invariant erosion is given by $\varepsilon_A(X) = \{y \in E \mid A + y \subseteq X\} = \{y \in E \mid y \in X + \breve A\}$, where $A$ is an element of $\wp(E)$, called the \emph{structuring element}, and $\breve A = \{- a \mid a \in A\}$ is the reflection of $A$ around the origin.
Now we observe that the above expressions for erosion and dilation can also be written, respectively, as
\begin{equation}\label{trinv}
\d_A(X)(y) = \bigvee_{x \in E} A(y - x) \wedge X(x), \qquad \varepsilon_A(Y)(x) = \bigwedge_{y \in E} A(y - x) \to Y(y),
\end{equation}
where each subset $X$ of $E$ is identified with its (Boolean) membership function and $X \to Y =: X^c \vee Y$.
Moving from these expressions, and recalling that $\wedge$ is a residuated commutative operation (that is, a continuous t-norm) whose residuum is $\to$, it is possible to extend these operations from the complete lattice of sets $\wp(E) = \{0,1\}^E$ to the complete lattice of fuzzy sets $[0,1]^E$, by means of left-continuous t-norms and their residua. What we do, concretely, is to extend the morphological image operators of dilation and erosion, from the case of binary images, to the case of grey-scale images.
So let $\ast$ be a left-continuous t-norm and $\to$ be its residuum; a grey-scale image $X$ is a fuzzy subset of $E$. Given a fuzzy subset $A \in [0,1]^E$, called a \emph{fuzzy structuring element}, the operators
\begin{center}$
\d_A(X)(y) = \bigvee_{x \in E} A(y - x) \ast X(x), \qquad \varepsilon_A(X)(x) = \bigwedge_{y \in E} A(y - x) \to X(y)
$\end{center}
are, respectively, a translation invariant dilation and erosion on $[0,1]^E$.
\section{Quantale modules}
\label{qmod}
\begin{definition}\label{modules}
Let $\mathbf{Q}$ be a quantale and $\mathbf{M} = \langle M, \vee, \bot \rangle$ a sup-lattice. $\mathbf{M}$ is a (left) \emph{$\mathbf{Q}$-module} if there exists an external binary operation, called \emph{scalar multiplication}, \ $\star: (q,m) \in Q \times M \longmapsto q \star m \in M$, \ such that
\begin{enumerate}
\item[$(M1)$]$(q_1 \cdot q_2) \star m = q_1 \star (q_2 \star m)$, for all $q_1, q_2 \in Q$ and $m \in M$;
\item[$(M2)$]the external product is distributive with respect to arbitrary joins in both coordinates or --- that is the same --- it is residuated;
\item[$(M3)$]$e \star m = m$.
\end{enumerate}
\end{definition}
From $(M2)$ it follows that, for all $q \in Q$, there exists the residual map $(q^\star)_*$ of $q^\star$, and for all $m \in M$ there exists the residual map $(^\star m)_*$ of $^\star m$. In particular it is possible to define another external operation over $M$:
\begin{center}
$\under_\star: (q,m) \in Q \times M \longmapsto q \under_\star m = (q^\star)_*(m) \in M.$
\end{center}
\begin{exe}\label{freemodule}
Let $\mathbf{Q}$ be a quantale and $X$ be an arbitrary non-empty set. We can consider the sup-lattice $\mathbf{Q}^X = \langle Q^X, \vee^X, \bot^X \rangle$, where $\bot^X$ is the $\bot$-constant function from $X$ to $Q$ and the join and the scalar multiplication $\star$ are defined pointwisely from those in $\mathbf{Q}$.
Then $\mathbf{Q}^X$ is a left $\mathbf{Q}$-module and, for all $q \in Q$, $f \in Q^X$ and $x \in X$, $(q \under_\star f)(x) = q \backslash f(x)$.
\end{exe}
It is easy to show that the module in the previous example is the free $\mathbf{Q}$-module over the set of generators $X$.
Definition, and properties, of right $\mathbf{Q}$-modules are completely analogous. If $\mathbf{Q}$ is commutative, right and left $\mathbf{Q}$-modules coincide and we will say simply $\mathbf{Q}$-modules. If $\mathbf{Q}$ is a quantale and $\mathbf{M}$ is a left $\mathbf{Q}$-module, the dual sup-lattice $\mathbf{M}^{\operatorname{op}}$ is a right $\mathbf{Q}$-module (and vice versa) with the external multiplication $\under_\star$.\footnote{In what follows, in all the definitions and results that can be stated both for left and right modules, we will refer generically to ``modules''~--- without specifying left or right~--- and we will use the notations of left modules.}
Let $\mathbf{Q}$ be a quantale and $\mathbf{M}_1, \mathbf{M}_2$ be two $\mathbf{Q}$-modules. A map $f: M_1 \longrightarrow M_2$ is a $\mathbf{Q}$-module homomorphism if $f\left({}^{\mathbf{M}_1}\bigvee_{i \in I} m_i\right) = {}^{\mathbf{M}_2}\bigvee_{i \in I} f(m_i)$ for any family $\{m_i\}_{i \in I} \subseteq \mathbf{M}_1$, and $f(q \star_1 m) = q \star_2 f(m)$, for all $q \in Q$ and $m \in M_1$, where $\star_i$ is the external product of $\mathbf{M}_i$, for $i = 1,2$.
\begin{proposition}\label{dualhomomq}
Let $\mathbf{Q}$ be a quantale, $\mathbf{M}_1$, $\mathbf{M}_2$ be two $\mathbf{Q}$-modules and $f: \mathbf{M}_1 \longrightarrow \mathbf{M}_2$ be a homomorphism. Then $f$ is a residuated map and the residual map $f_*: M_2 \longrightarrow M_1$ is a $\mathbf{Q}$-module homomorphism between $\mathbf{M}_2^{\operatorname{op}}$ and $\mathbf{M}_1^{\operatorname{op}}$.
\end{proposition}
\begin{definition}\label{hommq}
Let $\mathbf{M}$ and $\mathbf{N}$ be two $\mathcal{Q}$-modules and $\hom_\mathbf{Q}(\mathbf{M},\mathbf{N})$, the set of all the homomorphisms from $\mathbf{M}$ to $\mathbf{N}$. Then the structure $\mathbf{Hom}_\mathbf{Q}(\mathbf{M},\mathbf{N}) = \langle \hom_\mathbf{Q}(\mathbf{M},\mathbf{N}), \sqcup, \bot^\bot \rangle$, with the pointwise join and the $\bot$-constant homomorphism as bottom element, is a sup-lattice; moreover, if $\mathbf{Q}$ is a commutative quantale, $\mathbf{Hom}_\mathbf{Q}(\mathbf{M},\mathbf{N})$ is a $\mathbf{Q}$-module with the scalar multiplication $\diamond$ defined, again, pointwisely: for all $q \in Q$ and $h \in \hom_\mathbf{Q}(\mathbf{M},\mathbf{N})$, $q \diamond h$ is the homomorphism defined by $(q \diamond h)(x) = q \star h(x) = h(q \star x)$, for all $x \in M$.
If $\mathbf{N} = \mathbf{M}$, $\mathbf{Hom}_\mathbf{Q}(\mathbf{M},\mathbf{M})$ is denoted by $\mathbf{End}_\mathbf{Q}(\mathbf{M}) = \langle \operatorname{End}_\mathbf{Q}(\mathbf{M}), \sqcup, \bot^\bot \rangle$.
\end{definition}
\section{Quantale module transforms}
\label{mqtransec}
In this section we introduce the $\mathcal{Q}$-module transforms and we list some results about them. Then we present a classification of these operators that have interesting theoretical and concrete consequences. For an extensive treatment of $\mathcal{Q}$-module transforms, the reader may refer to \cite{russo}.
\begin{definition}\label{qwtransform}
Let $\mathbf{Q} \in \mathcal{Q}$ and $X, Y$ be non-empty sets and let us consider the free $\mathbf{Q}$-modules $\mathbf{Q}^X$ and $\mathbf{Q}^Y$. We will call a \emph{$\mathcal{Q}$-module transform} between $\mathbf{Q}^X$ and $\mathbf{Q}^Y$, with \emph{kernel} $p$, the operator $H_p: Q^X \longrightarrow Q^Y$ defined by
\begin{center}
$H_p f(y) = \bigvee_{x \in X} f(x) \cdot p(x,y) \quad \textrm{for all } y \in Y,$
\end{center}
where $p \in Q^{X \times Y}$. Its \emph{inverse transform} $\Lambda_p: Q^Y \longrightarrow Q^X$ is the map defined by
\begin{center}
$\Lambda_p g(x) = \bigwedge_{y \in Y} g(y) / p(x,y) \quad \textrm{for all } x \in X.$
\end{center}
\end{definition}
\begin{theorem}\label{wadjointpair}
Let $\mathbf{Q} \in \mathcal{Q}$, $X, Y$ be two non-empty sets and $p \in Q^{X \times Y}$. If $H_p$ is the $\mathcal{Q}$-module transform, with kernel $p$, between $\mathbf{Q}^X$ and $\mathbf{Q}^Y$, and $\Lambda_p$ is its inverse transform, then the following hold:
\begin{enumerate}
\item[$(i)$]$(H_p,\Lambda_p)$ is an adjoint pair, i.e. $H_p$ is a residuated map and $\Lambda_p = {H_p}_*$;
\item[$(ii)$]$H_p \in \hom_\mathbf{Q}\left(\mathbf{Q}^X,\mathbf{Q}^Y\right)$ and $\Lambda_p \in \hom_\mathbf{Q}\left(\left(\mathbf{Q}^Y\right)^{\operatorname{op}},\left(\mathbf{Q}^X\right)^{\operatorname{op}}\right)$.
\end{enumerate}
\end{theorem}
The following classification of the kernels has a few interesting theoretical implications but it is important for applications to image processing. We refer to \cite{dinolarusso} (where an orthonormal transform is presented), \cite{russothesis} and \cite{russo} for details.
\begin{definition}\label{coders}
Let $\mathbf{Q} \in \mathcal{Q}$, and $X, Y$ be non-empty sets. Let us consider a map $p \in Q^{X \times Y}$; we set the following definitions:
\begin{enumerate}
\item[$(i)$]$p$ is called a \emph{coder} iff there exists an injective map $\varepsilon: Y \longrightarrow X$ such that $e \leq p(\varepsilon(y),y)$ for all $y \in Y$;
\item[$(ii)$]$p$ is said to be \emph{normal} iff there exists an injective map $\varepsilon: Y \longrightarrow X$ such that $p(\varepsilon(y),y) = e$ for all $y \in Y$;
\item[$(iii)$]$p$ is said to be \emph{strong} iff it is normal and \ $p(\varepsilon(y_1),y_2) = \bot$ \ for all $y_1 \neq y_2 \in Y$;
\item[$(iv)$]$p$ is said to be \emph{orthogonal} iff $p(x,y_1) \cdot p(x,y_2) = \bot$ for all $y_1, y_2 \in Y$ such that $y_1 \neq y_2$ and for all $x \in X$;
\item[$(v)$]$p$ is said to be \emph{orthonormal} iff it is orthogonal and normal.
\end{enumerate}
If $p$ is a coder, the $\mathcal{Q}$-module transform $H_p$ is called \emph{faithful} and, if $p$ is normal, strong, orthogonal or orthonormal, the corresponding transform will have the same adjective. Also, we observe that $(v) \Longrightarrow (iii) \Longrightarrow (ii) \Longrightarrow (i)$.
\end{definition}
\begin{theorem}\label{sadjointpair}
Let $\mathbf{Q} \in \mathcal{Q}$ and let $H_p$ be a $\mathcal{Q}$-module strong transform, by the coder $p \in Q^{X \times Y}$, with inverse transform $\Lambda_p$. Then \ $H_p \circ \Lambda_p = \operatorname{id}_{Q^Y}$; \ thus $H_p$ is onto and, by Proposition~\ref{residprop}, $\Lambda_p$ is one-one.
\end{theorem}
\begin{lemma}\label{transf=coder}
Let $\mathbf{Q} \in \mathcal{Q}$, $X$ be a non-empty set, $Y$ be a non-empty subset of $X$ and $p, p' \in Q^{X \times Y}$ be two maps. Then $H_p = H_{p'}$ if and only if $p = p'$.
\end{lemma}
The previous result ensures us that a $\mathcal{Q}$-module transform $H_p$ is completely determined by its kernel $p$, while next result is the converse of Theorem~\ref{wadjointpair}$(ii)$; it proves that all the homomorphisms between free modules are transforms.
\begin{theorem}\label{repr}
The sup-lattices $\mathbf{Hom}_\mathbf{Q}(\mathbf{Q}^X,\mathbf{Q}^Y)$ and $\mathbf{Q}^{X \times Y}$ are isomorphic. And, if $\mathbf{Q}$ is commutative, they are isomorphic also as $\mathbf{Q}$-modules. In particular $\mathbf{End}_\mathbf{Q}(\mathbf{Q}^X) \cong \mathbf{Q}^{X \times X}$.
\end{theorem}
Here we just ``scratched the surface'' of $\mathcal{Q}$-module transforms, especially in order to focus the attention on the main thesis of the present paper, i.e. the fact that fuzzy image compression and mathematical morphological operators fall within the same class of operators under an algebraic point of view. We, again, refer to \cite{russothesis} or \cite{russo} the reader who may be interested in $\mathcal{Q}$-module transforms.
\section{Conclusion}
\label{conc}
As the reader may have already noticed, the operators in Section \ref{fuz} are all special cases of $[0,1]$-module transforms. We will now analyse them in detail.
Let us consider the compression operator defined in Subsection \ref{fuzalgsec}. Its domain is $[0,1]^{m \times n}$ and its codomain is $[0,1]^{a \times b}$; in the light of the definitions and results presented so far, we get immediately that (\ref{com}) is the $[0,1]$-module transform $H_C: [0,1]^{m \times n} \longrightarrow [0,1]^{a \times b}$ with kernel $C \in [0,1]^{m \times n \times a \times b}$ and the reconstruction (\ref{rec}) is its inverse transform $\Lambda_C: [0,1]^{a \times b} \longrightarrow [0,1]^{m \times n}.$
We already observed in Subsection~\ref{mathmorphsec} that dilations are precisely the sup-lattice homomorphisms while erosions are their residua. In order to faithfully represent dilations and erosions that are translation invariant as $\mathcal{Q}$-module transforms from a free $[0,1]$-module to itself, we make the further assumption that the set over which the free module is defined has the additional structure of Abelian group. So let $\mathbf{X} = \langle X, +, -, 0 \rangle$ be an Abelian group, $\ast$ a t-norm on $[0,1]$ and consider the free $[0,1]$-module $[0,1]^X$. For any element $k \in [0,1]^X$, we define the two variable map $\overline k: (x,y) \in X \times X \longmapsto k(y - x) \in [0,1]$. Then, for all $k \in [0,1]^X$, the translation invariant dilation, on $[0,1]^X$, whose structuring element is $k$, is precisely the $\mathcal{Q}$-module transform $H_{\overline k}: [0,1]^X \longrightarrow [0,1]^X$, with the kernel $\overline k$ defined above. Obviously, the translation invariant erosion whose structuring element is $k$ is $\Lambda_{\overline k}$.
Then the representation of both classes of operators as $\mathcal{Q}$-module transforms is rather trivial. Actually, what we want to point out here is that, if we drop the assumption that our quantale is defined on $[0,1]$, the classes of transforms defined in this section become much wider. The purpose of this consideration is not to suggest purely speculative abstractions but, rather, to underline that suitable generalizations of these operators exist already and they may be useful provided their underlying ideas are extended to tasks involving other quantales.
|
\section{Introduction}
Fock space is an infinite dimensional vector space which is a representation of several important
algebras, as described in, for example, \cite[Chapter 14]{Kac:1990}. Here we consider the charge zero part of Fock space, which we denote by ${\bf F}$, and its $v$-deformation ${\bf F}_v$. The space ${\bf F}$ has a standard $\Bbb{Q}$-basis $\{\, |\lambda\rangle\ |\ \hbox{$\lambda$ is a partition}\}$ and ${\bf F}_v := {\bf F} \otimes_\Bbb{Q} \Bbb{Q}(v)$. Following Hayashi \cite{Hayashi:1990}, Misra and Miwa \cite{MM:1990} define an action of the quantized universal enveloping algebra $U_v(\widehat{\mathfrak{sl}}_\ell)$ on ${\bf F}_v$. The only non-zero matrix elements $\langle \mu| F_{\bar i} | \lambda \rangle$ of the Chevalley generators $F_{\bar i}$ in terms of the standard basis occur when $\mu$ is obtained by adding a single $\bar{i}$-colored box to $\lambda$, and these are powers of $v$.
We show that these powers of $v$ also appear naturally in the following context: Partitions with at most $N$ parts index polynomial Weyl modules $\Delta(\lambda)$ for the integral quantum group $U_q^\mathcal{A}(\mathfrak{gl}_N)$. Let $V$ be the standard $N$ dimensional representation of $U_q^\mathcal{A}(\mathfrak{gl}_N)$. If the matrix element
$\langle \mu| F_{\bar i} | \lambda \rangle$ is non-zero then, for sufficiently large $N$, $\left( \Delta^\mathcal{A}(\lambda) \otimes_\mathcal{A} V\right) \otimes_\mathcal{A} \Bbb{Q}(q)$ contains a highest weight vector of weight $\mu$. There is a unique such highest weight vector $v_\mu$ which satisfies a certain triangularity condition with respect to an integral basis of $\Delta^\mathcal{A}(\lambda) \otimes_\mathcal{A} V$. We show that the matrix element $\langle \mu| F_{\bar i} | \lambda \rangle$ is equal to $v^{\text{val}_{\phi_{2\ell}}(v_\mu, v_\mu)}$,
where $(\cdot, \cdot)$ is the Shapovalov form and $\text{val}_{\phi_{2\ell}}$ is
the valuation at the cyclotomic polynomial $\phi_{2\ell}$.
Our proof is computational, making use of the Shapovalov determinant \cite{Shapovalov:1972, KD:1991, KL:1997}. This is a formula for the determinant of the Shapovalov form on
a weight space of a Verma module. The necessary computation is most easily done
in terms of the universal Verma modules introduced in the classical case by Kashiwara \cite{Kashiwara:1985} and studied in the quantum case by Kamita \cite{Kamita}. The statement for Weyl modules is then a straightforward consequence.
Before beginning, let us discuss some related work.
In \cite{Kleshchev:1995}, Kleshchev carefully analyzed the $\mathfrak{gl}_{N-1}$ highest weight
vectors in a Weyl module for $\mathfrak{gl}_N$, and used this information to give modular branching rules for symmetric group representations.
Brundan and Kleshchev \cite{BK2} have explained that highest weight vectors in
the restriction of a Weyl module to $\mathfrak{gl}_{N-1}$ give information about highest weight vectors
in a tensor product $\Delta(\lambda)\otimes V$ of a Weyl module with the standard
$N$-dimensional representation of $\mathfrak{gl}_N$. Our computations
put a new twist on the analysis of the highest weight vectors
in $\Delta(\lambda)\otimes V$, as we study them
in their ``universal'' versions and by the use of the Shapovalov determinant. Our techniques can be viewed as an application of the theory of
Jantzen \cite{Jantzen:1974} as extended to the quantum case by Wiesner \cite{Wiesner}.
Brundan \cite{Brundan:1998} generalized
Kleshchev's \cite{Kleshchev:1995} techniques and used this information to give modular branching rules for Hecke algebras. As discussed in \cite{Ariki:1996, LLT:1996}, these branching rules are reflected in the fundamental representation of
$\widehat{\mathfrak{sl}}_p$ and its crystal graph, recovering much of the structure of the Misra-Miwa Fock space. Using Hecke algebras at a root of unity, Ryom-Hansen \cite{Ryom-Hansen:2004} recovered the full $U_v(\widehat{\mathfrak{sl}}_\ell)$ action on Fock space. To complete the picture one should construct a graded category, where multiplication by $v$ in the $\widehat{\mathfrak{sl}}_\ell$ representation corresponds to a grading shift.
Recent work of Brundan-Kleshchev \cite{BrKl:2009} and Ariki \cite{Ariki:2009} explains that one solution to this problem is through the representation theory of Khovanov-Lauda-Rouquier algebras \cite{KhLa:2009, Rouquier:2009}. It would be interesting to explicitly describe the relationship between their category and the present work. Another related construction due to Brundan-Stroppel considers the case when the Fock space is replaced by $\wedge^m V\otimes \wedge^n V$, where $V$ is the natural $\mathfrak{gl}_\infty$ module and $m, n$ are fixed natural numbers.
We would also like to mention very recent work of Peng Shan \cite{Shan:2010} which independently develops a similar story to the one presented here, but using representations of a quantum Schur algebra where we use representations of $U_\varepsilon(\mathfrak{gl}_N)$. The approach taken there is somewhat different, and in particular relies on localization techniques of Beilinson and Bernstein \cite{BB:1993}.
This paper is arranged as follows. Sections \ref{qua-group} and \ref{pFock} are background on the quantum group $U_q(\mathfrak{gl}_N)$ and the Fock space ${\bf F}_v$. Sections \ref{UV-section} and \ref{Sform} explain universal Verma modules and the Shapovalov determinant. Section \ref{makeops} contains the statement and proof of our main result relating Fock space and Weyl modules.
\subsection{Acknowledgments}
We thank M. Kashiwara, A. Kleshchev, T. Tanisaki, R. Virk and B. Webster for helpful discussions.
The first author was partly supported by NSF Grant DMS-0353038 and Australian Research Council Grants DP0986774 and DP0879951. The second author was partly supported by the Australia Research Council grant DP0879951 and NSF grant DMS-0902649.
\section{The quantum group $U_q(\mathfrak{gl}_N)$ and its integral form $U_q^\mathcal{A}(\mathfrak{gl}_N)$} \label{qua-group}
This is a very brief review, intended mainly to fix notation. With slight modifications the construction in this section works in the generality of symmetrizable Kac-Moody algebras. See \cite[Chapters 6 and 9]{CP} for details.
\subsection{The rational quantum group
} \label{the_q_groups}
$U_q(\mathfrak{gl}_N)$ is the associative algebra over the field of rational
functions $\Bbb{Q}(q)$ generated by
\begin{equation}
X_1,\ldots, X_{N-1}, \quad
Y_1,\ldots, Y_{N-1},\quad\hbox{and}\quad
L_1^{\pm1},\ldots, L_N^{\pm1},
\end{equation}
with relations
$$L_iL_j=L_jL_i, \quad
L_iL_i^{-1}= L_i^{-1}L_i =1,
\qquad
X_i Y_j - Y_j X_i = \delta_{i,j} \frac{L_iL_{i+1}^{-1} - L_{i+1} L_i^{-1}}{q-q^{-1}},
$$
\begin{equation} L_i X_j L_i^{-1}=
\begin{cases} qX_j, \quad \text{ if } i=j, \\
q^{-1}X_j, \quad \text{ if } i=j+1, \\
X_j \quad \text{ otherwise};
\end{cases}
\qquad\qquad
L_i Y_j L_i^{-1}=
\begin{cases} q^{-1}Y_j, \quad \text{ if } i=j, \\
q Y_j, \quad \text{ if } i=j+1, \\
Y_j, \quad \text{ otherwise};
\end{cases}
\end{equation}
$$X_iX_j = X_j X_i\quad\hbox{and}\quad Y_i Y_j=Y_jY_i,
\qquad\hbox{if $|i-j| \geq 2$,}$$
$$
X_i^{2} X_j -(q+q^{-1}) X_iX_jX_i + X_j X_i^2 = Y_i^{2} Y_j -(q+q^{-1})Y_iY_jY_i + Y_j Y_i^2=0,
\qquad
\hbox{if $|i-j|=1$.}
$$
The algebra $U_q(\mathfrak{gl}_N)$ is a Hopf algebra with coproduct and antipode given by
\begin{equation} \label{coproduct-antipode}
\begin{aligned}
& \Delta(L_i) = L_i\otimes L_i, \\
&\Delta(X_i)= X_i \otimes L_iL_{i+1}^{-1}+ 1 \otimes X_i, \\
& \Delta(Y_i)= Y_i \otimes 1+ L_i^{-1}L_{i+1} \otimes Y_i,
\end{aligned}
\qquad\hbox{and}\qquad
\begin{aligned}
& S(L_i) = L_i^{-1}, \\
&S(X_i)= -X_i L_i^{-1}L_{i+1}, \\
& S(Y_i)= -L_iL_{i+1}^{-1}Y_i,
\end{aligned}
\end{equation}
respectively (see \cite[Section 9.1]{CP}).
As a $\Bbb{Q}(q)$-vector space, $U_q(\mathfrak{gl}_N)$ has a triangular decomposition
\begin{equation} \label{rat-triangle}
U_q(\mathfrak{gl}_N) \cong U_q(\mathfrak{gl}_N)^{<0} \otimes U_q(\mathfrak{gl}_N)^0 \otimes U_q(\mathfrak{gl}_N)^{>0},
\end{equation}
where the inverse isomorphism is given by multiplication (see \cite[Proposition 9.1.3]{CP}). Here $U_q(\mathfrak{gl}_N)^{<0}$ is the subalgebra generated by the $Y_i$ for $i=1, \ldots, N-1$, $U_q(\mathfrak{gl}_N)^{>0}$ is the subalgebra generated by the $X_i$ for $i=1, \ldots, N-1,$ and $U_q(\mathfrak{gl}_N)^0$ is the subalgebra generated by the $L_i^{\pm1}$ for $i=1, \ldots, N$.
\subsection{The integral quantum group} \label{res-form}
Let $\mathcal{A}= \Bbb{Z}[q, q^{-1}]$. For $n,k \in \Bbb{Z}_{>0}$ and $c \in \Bbb{Z}$, let
\begin{align}
[n]:= \frac{q^n -q^{-n}}{q-q^{-1}},
\;
x^{(k)}:=\frac{x^k}{[k] [k-1] \cdots [2][1]},
\; \hbox{and}\;
\left[
\begin{array}{c}
x; c \\
k
\end{array}
\right] :=
\prod_{s=1}^k \frac{xq^{c+1-s}-x^{-1}q^{s-1-c}}{q^s-q^{-s}},
\end{align}
in $\Bbb{Q}(q,x)$.
The \emph{restricted integral form} $U_q^{\mathcal{A}}(\mathfrak{gl}_N)$ is the $\mathcal{A}$-subalgebra of $U_q(\mathfrak{gl}_N)$ generated by $X_i^{(k)}, Y_i^{(k)}$, $L_i^{\pm 1}$ and $\left[
\begin{array}{c}
L_i; c \\
k
\end{array}
\right]$ for $1 \leq i \leq N, c \in \Bbb{Z}, k \in \Bbb{Z}_{>0}$.
As discussed in \cite[Section 6]{Lusztig:1990}, this is an integral form in the sense that
\begin{equation} \label{int-form}
U_q^\mathcal{A}(\mathfrak{gl}_N) \otimes_{\mathcal{A}} \Bbb{Q}(q) = U_q(\mathfrak{gl}_N).
\end{equation}
As with $U_q(\mathfrak{gl}_N)$, the algebra $U_q^\mathcal{A}(\mathfrak{gl}_N)$ has a triangular decomposition
\begin{equation} \label{int-triangle}
U_q^\mathcal{A}(\mathfrak{gl}_N) \cong U_q^\mathcal{A}(\mathfrak{gl}_N)^{<0} \otimes U_q^\mathcal{A}(\mathfrak{gl}_N)^0 \otimes U_q^\mathcal{A}(\mathfrak{gl}_N)^{>0},
\end{equation}
where the isomorphism is given by multiplication (see \cite[Proposition 9.3.3]{CP}).
In this case, $U_q^\mathcal{A}(\mathfrak{gl}_N)^{<0}$ is the subalgebra generated by the $Y_i^{(k)}$, $U_q^\mathcal{A}(\mathfrak{gl}_N)^{>0}$ is the subalgebra generated by the $X_i^{(k)}$, and $U_q^\mathcal{A}(\mathfrak{gl}_N)^0$ is generated by $L_i^{\pm 1}$ and $\displaystyle \left[
\begin{array}{c}
L_i; c \\
k
\end{array}
\right]$
for $1 \leq i \leq N$, $c \in \Bbb{Z}$, and $ k \in \Bbb{Z}_{>0}$.
\subsection{Rational representations} \label{rational-reps}
The Lie algebra $\mathfrak{gl}_N = M_N(\mathbb{C})$ of $N\times N$ matrices has standard basis
$\{ E_{ij} \ |\ 1\le i,j\le N\},$ where $E_{ij}$
is the matrix with $1$ in position $(i,j)$ and $0$ everywhere else.
Let $\mathfrak{h}= \text{span} \{ E_{11}, E_{22}, \ldots, E_{NN} \}$. Let $\varepsilon_i \in \mathfrak{h}^*$ be the weight
of $\mathfrak{gl}_N$ given by $\varepsilon_i(E_{jj})= \delta_{i,j}$. Define
\begin{equation}
\begin{aligned}
&\mathfrak{h}^*_\Bbb{Z} := \{ \lambda=\lambda_1\varepsilon_1+\lambda_2 \varepsilon_2 + \cdots
+ \lambda_N \varepsilon_N \in \mathfrak{h}^* \ |\ \lambda_1,\ldots, \lambda_N\in \Bbb{Z}\}, \\
&(\mathfrak{h}^*_\Bbb{Z})^+
:= \{ \lambda=\lambda_1\varepsilon_1+\lambda_2 \varepsilon_2 + \cdots + \lambda_N \varepsilon_N
\in \mathfrak{h}_\Bbb{Z}^* \ |\ \lambda_1 \geq \lambda_2 \geq \cdots \geq \lambda_N
\}, \\
&P^+
:= \{ \lambda=\lambda_1\varepsilon_1+\lambda_2 \varepsilon_2 + \cdots + \lambda_N \varepsilon_N
\in (\mathfrak{h}_\Bbb{Z}^*)^+ \ |\ \lambda_N\ge 0\}, \\
&R^+ :=\{ \varepsilon_i-\varepsilon_j\ |\ 1 \leq i < j \leq N \}, \\
&Q:= \text{span}_\Bbb{Z}(R^+), \quad
Q^+:= \text{span}_{\Bbb{Z}_{\ge0}}(R^+), \quad \text{and}\quad Q^-:= \text{span}_{\Bbb{Z}_{\le0}}(R^+).
\end{aligned}
\end{equation}
to be the set of \emph{integral weights}, the set of \emph{dominant integral weights}, the set of \emph{dominant polynomial weights}, the set of \emph{positive roots}, the \emph{root lattice}, the \emph{positive part of the root lattice}, and the the \emph{negative part of the root lattice}, respectively.
For an integral weight $\lambda= \lambda_1 \varepsilon_1 + \cdots+\lambda_N\varepsilon_N$,
the \emph{Verma module} $M(\lambda)$ for $U_q(\mathfrak{gl}_N)$ of highest weight $\lambda$ is
\begin{equation}
M(\lambda) := U_q(\mathfrak{gl}_N) \otimes_{U_q(\mathfrak{gl}_N)^{\geq 0}} \Bbb{Q}(q)_\lambda,
\end{equation}
where
$\Bbb{Q}(q)_\lambda= \text{span}_{\Bbb{Q}(q)} \{ v_{\lambda} \}$ is the one dimensional vector space over
$\Bbb{Q}(q)$ with $U_q(\mathfrak{gl}_N)^{\geq 0}$ action given by
\begin{equation}
X_i \cdot v_\lambda =0 \quad\hbox{and}\quad
L_j \cdot v_\lambda =q^{\lambda_j}v_\lambda,
\qquad\hbox{for $1 \leq i \leq N-1$, $1 \leq j \leq N$.}
\end{equation}
\begin{Theorem} \label{Deltaconst} \emph{(see \cite[Chapter 10.1]{CP})}
If $\lambda\in (\mathfrak{h}_\mathbb{Z}^*)^+$ then $M(\lambda)$ has a unique finite dimensional quotient
$\Delta(\lambda)$ and the map $\lambda\mapsto \Delta(\lambda)$ is a bijection between
$(\mathfrak{h}_\mathbb{Z}^*)^+$ and the set of isomorphism classes of irreducible finite dimensional $U_q(\mathfrak{gl}_N)$-modules. \end{Theorem}
A \emph{singular vector} in a representation of $U_q(\mathfrak{gl}_N)$ is a vector $v$ such that $X_i \cdot v=0$ for all $i$.
\subsection{Integral representations}
The \emph{integral Verma module} $M^\mathcal{A}(\lambda)$ is the $U_q^{\mathcal{A}}(\mathfrak{gl}_N)$-submodule
of $M(\lambda)$ generated by $v_\lambda$. The \emph{integral Weyl module} $\Delta^\mathcal{A}(\lambda)$
is the $U_q^{\mathcal{A}}(\mathfrak{gl}_N)$-submodule of $\Delta(\lambda)$ generated by $v_\lambda$.
Using \eqref{int-form} and \eqref{rat-triangle},
\begin{equation}
M^{\mathcal{A}} (\lambda) \otimes_{\mathcal{A}} \Bbb{Q}(q)= M(\lambda),
\quad \text{and} \quad
\Delta^{\mathcal{A}}(\lambda) \otimes_{\mathcal{A}} \Bbb{Q}(q)=\Delta(\lambda).
\end{equation}
In general, $\Delta^{\mathcal{A}}(\lambda)$ is not irreducible as a $U^\mathcal{A}_q(\mathfrak{gl}_N)$ module.
\section{Partitions and Fock space} \label{pFock}
We now describe the $v$-deformed Fock space representation of $U_v(\widehat{\mathfrak{sl}}_\ell)$ constructed by Misra and Miwa \cite{MM:1990} following work of Hayashi \cite{Hayashi:1990}. Our presentation largely follows \cite[Chapter 10]{ariki:2000}.
\subsection{Partitions}
A partition $\lambda$ is a finite length non-increasing sequence of positive integers. Associated to a partition is its Ferrers diagram. We draw these diagrams as in Figure \ref{partition_bij} so that,
if $\lambda = (\lambda_1,\ldots, \lambda_N)$, then $\lambda_i$ is the number of boxes in row $i$
(rows run southeast to northwest
$\nwarrow$
). Say that $\lambda$ is contained in $\mu$ if the diagram
for $\lambda$ fits inside the diagram for $\mu$ and let $\mu/\lambda$ be the collection of boxes
of $\mu$ that are not in $\lambda$.
For each box $b \in \lambda$, the \emph{content} $c(b)$ is the horizontal position of $b$
and the \emph{color} $\overline c(b)$ is the residue of $c(b)$ modulo $\ell$.
In Figure \ref{partition_bij}, the numbers $c(b)$ are listed below the diagram.
The \emph{size} $|\lambda|$ of a partition $\lambda$ is the total number of boxes in its Ferrers diagram.
The set $P^+$ of dominant polynomial weights from Section \ref{rational-reps} is naturally identified with partitions with at most $N$ parts.
If $\lambda \in P^+$ then
\begin{equation} \label{plusboxes}
\Delta(\lambda) \otimes \Delta(\varepsilon_1) \cong
\bigoplus_{\tiny \begin{array}{l} 1 \leq k \leq N \\ \lambda+\varepsilon_k \in P^+
\end{array} } \Delta(\lambda+\varepsilon_k)
\end{equation}
as $U_q(\mathfrak{gl}_N)$-modules. The diagram of $\lambda+\varepsilon_k$ is obtained from the diagram of $\lambda$ by adding a box on row $k$, and $\Delta(\lambda+\varepsilon_k)$ appears in the sum on the right side of \eqref{plusboxes} if and only if $\lambda+\varepsilon_k$ is a partition.
See, for example, \cite[Section 6.1, Formula 6.8]{FH:1991} for the classical statement, and \cite[Proposition 10.1.16]{CP} for the quantum case.
\begin{figure}
\setlength{\unitlength}{0.4cm}
\begin{center}
\begin{picture}(20,12)
\put(0,1){
\begin{picture}(20,11)
\put(10,0){\line(1,1){10}}
\put(9,1){\line(1,1){8}}
\put(8,2){\line(1,1){7}}
\put(7,3){\line(1,1){7}}
\put(6,4){\line(1,1){5}}
\put(5,5){\line(1,1){5}}
\put(4,6){\line(1,1){3}}
\put(3,7){\line(1,1){1}}
\put(10,0){\line(-1,1){10}}
\put(11,1){\line(-1,1){7}}
\put(12,2){\line(-1,1){6}}
\put(13,3){\line(-1,1){6}}
\put(14,4){\line(-1,1){5}}
\put(15,5){\line(-1,1){5}}
\put(16,6){\line(-1,1){3}}
\put(17,7){\line(-1,1){3}}
\put(18,8){\line(-1,1){1}}
\put(3.4,6.7){{ \color{blue} ${\bar 0}$}}
\put(4.4,5.7){{ \color{green} ${\bar 2}$}}
\put(5.4,6.7){{ \color{red} ${\bar 1}$}}
\put(5.4,4.7){{ \color{red} ${\bar 1}$}}
\put(6.4,7.7){{ \color{blue} ${\bar 0}$}}
\put(6.4,5.7){{ \color{blue} ${\bar 0}$}}
\put(6.4,3.7){{ \color{blue} ${\bar 0}$}}
\put(7.4,6.7){{ \color{green} ${\bar 2}$}}
\put(7.4,4.7){{ \color{green} ${\bar 2}$}}
\put(7.4,2.7){{ \color{green} ${\bar 2}$}}
\put(8.4,7.7){{ \color{red} ${\bar 1}$}}
\put(8.4,5.7){{ \color{red} ${\bar 1}$}}
\put(8.4,3.7){{ \color{red} ${\bar 1}$}}
\put(8.4,1.7){{ \color{red} ${\bar 1}$}}
\put(9.4,8.7){{ \color{blue} ${\bar 0}$}}
\put(9.4,6.7){{ \color{blue} ${\bar 0}$}}
\put(9.4,4.7){{ \color{blue} ${\bar 0}$}}
\put(9.4,2.7){{ \color{blue} ${\bar 0}$}}
\put(9.4,0.7){{ \color{blue} ${\bar 0}$}}
\put(10.4,7.7){{ \color{green} ${\bar 2}$}}
\put(10.4,5.7){{ \color{green} ${\bar 2}$}}
\put(10.4,3.7){{ \color{green} ${\bar 2}$}}
\put(10.4,1.7){{ \color{green} ${\bar 2}$}}
\put(11.4,6.7){{ \color{red} ${\bar 1}$}}
\put(11.4,4.7){{ \color{red} ${\bar 1}$}}
\put(11.4,2.7){{ \color{red} ${\bar 1}$}}
\put(12.4,7.7){{ \color{blue} ${\bar 0}$}}
\put(12.4,5.7){{ \color{blue} ${\bar 0}$}}
\put(12.4,3.7){{ \color{blue} ${\bar 0}$}}
\put(13.4,8.7){{ \color{green} ${\bar 2}$}}
\put(13.4,6.7){{ \color{green} ${\bar 2}$}}
\put(13.4,4.7){{ \color{green} ${\bar 2}$}}
\put(14.4,7.7){{ \color{red} ${\bar 1}$}}
\put(14.4,5.7){{ \color{red} ${\bar 1}$}}
\put(15.4,6.7){{ \color{blue} ${\bar 0}$}}
\put(16.4,7.7){{ \color{green} ${\bar 2}$}}
\put(9.6,-1){ $\tiny{{}_{0}}$}
\put(10.1,-1){ $\tiny{{}_{-1}}$}
\put(11.1,-1){ $\tiny{{}_{-2}}$}
\put(12.1,-1){ $\tiny{{}_{-3}}$}
\put(13.1,-1){ $\tiny{{}_{-4}}$}
\put(14.1,-1){ $\tiny{{}_{-5}}$}
\put(15.1,-1){ $\tiny{{}_{-6}}$}
\put(16.1,-1){ $\tiny{{}_{-7}}$}
\put(17.1,-1){ $\tiny{{}_{-8}}$}
\put(18.1,-1){ $\tiny{{}_{-9}}$}
\put(8.6,-1){ $\tiny{{}_{1}}$}
\put(7.6,-1){ $\tiny{{}_{2}}$}
\put(6.6,-1){ $\tiny{{}_{3}}$}
\put(5.6,-1){ $\tiny{{}_{4}}$}
\put(4.6,-1){ $\tiny{{}_{5}}$}
\put(3.6,-1){ $\tiny{{}_{6}}$}
\put(2.6,-1){ $\tiny{{}_{7}}$}
\put(1.6,-1){ $\tiny{{}_{8}}$}
\put(0.6,-1){ $\tiny{{}_{9}}$}
\thicklines
\end{picture}}
\end{picture}
\end{center}
\vspace{0.4cm}
\caption{The partition $(7,6,6,5,5,3,3,1)$
with each box containing its color for $\ell =3$.
The content $c(b)$ of a box $b$ is the horizontal position of $b$ reading right to left. The contents
of boxes are listed beneath the diagram so that $c(b)$ is aligned with all boxes $b$ of that content.
\label{partition_bij}}
\end{figure}
\subsection{The quantum affine algebra}
Let $U'_v(\widehat{\mathfrak{sl}}_\ell)$ be the quantized universal enveloping algebra corresponding to the $\ell$-node Dynkin diagram
\vspace{0.2cm}
\begin{center}
\setlength{\unitlength}{0.4cm}
\begin{picture}(6,2)
\put(-1,0){\circle*{0.5}}
\put(0,0){\circle*{0.5}}
\put(1,0){\circle*{0.5}}
\put(5,0){\circle*{0.5}}
\put(6,0){\circle*{0.5}}
\put(7,0){\circle*{0.5}}
\put(3,2){\circle*{0.5}}
\put(3,2){\line(2,-1){4}}
\put(3,2){\line(-2,-1){4}}
\put(-1,0){\line(1,0){1}}
\put(0,0){\line(1,0){1}}
\put(1,0){\line(1,0){1}}
\put(4,0){\line(1,0){1}}
\put(5,0){\line(1,0){1}}
\put(6,0){\line(1,0){1}}
\put(2.6,0){\ldots}
\end{picture}
\end{center}
\vspace{0.1cm}
More precisely, $U'_v(\widehat{\mathfrak{sl}}_\ell)$ is the algebra generated by $ E_{\bar i}, F_{\bar i},
K_{\bar i}^{\pm1}$, for $\bar i \in \Bbb{Z}/\ell \Bbb{Z}$, with relations
$$K_{\bar i}K_{\bar j}=K_{\bar j}K_{\bar i}, \quad
K_{\bar i}K_{\bar i}^{-1}= K_{\bar i}^{-1}K_{\bar i} =1,
\qquad
E_{\bar i} F_{\bar j} - F_{\bar j} E_{\bar i}
= \delta_{\bar i,\bar j} \frac{K_{\bar i} - K_{\bar i}^{-1}}{v-v^{-1}},
$$
\begin{equation}
K_{\bar i} E_{\bar j} K_{\bar i}^{-1}=
\begin{cases} v^2 E_{\bar j}, \quad \text{ if } \bar i=\bar j, \\
v^{-1}E_{\bar j}, \quad \text{ if } \bar i=\bar j\pm1, \\
E_{\bar j} \quad \text{ otherwise};
\end{cases}
\qquad
K_{\bar i} F_{\bar j} K_{\bar i}^{-1}=
\begin{cases} v^{-2}F_{\bar j}, \quad \text{ if } \bar i=\bar j, \\
v F_{\bar j}, \quad \text{ if } \bar i=\bar j\pm1, \\
F_{\bar j}, \quad \text{ otherwise};
\end{cases}
\end{equation}
$$E_{\bar i}E_{\bar j} = E_{\bar j} E_{\bar i}
\quad\hbox{and}\quad
F_{\bar i} F_{\bar j}=F_{\bar j}F_{\bar i},
\qquad\hbox{if $|\bar i-\bar j| \geq 2$,}$$
$$
E_{\bar i}^{2} E_{\bar j} -(v+v^{-1}) E_{\bar i}E_{\bar j}E_{\bar i} + E_{\bar j} E_{\bar i}^2
= F_{\bar i}^{2} F_{\bar j} -(v+v^{-1}) F_{\bar i}F_{\bar j}F_{\bar i} + F_{\bar j} F_{\bar i}^2=0,
\qquad
\hbox{if $|\bar i-\bar j|=1$.}
$$
See \cite[Definition Proposition 9.1.1]{CP}.
The algebra $U_v'(\widehat{\mathfrak{sl}}_\ell)$ is the quantum group corresponding
to the non-trivial central extension
$\widehat{\mathfrak{sl}}_\ell' = \mathfrak{sl}_\ell[t,t^{-1}] \oplus \mathbb{C} c$ of the algebra of polynomial loops in $\mathfrak{sl}_\ell$.
\subsection{Fock space} \label{MM_section}
Define \emph{$v$-deformed Fock space} to be the $\Bbb{Q}(v)$ vector space
${\bf F}_v$ with basis
$\{ |\lambda \rangle \ |\ \lambda \text{ is a partition} \}$.
Our ${\bf F}_v$ is only the charge $0$ part of Fock space described in \cite{KMS:1995}.
Fix $\bar i \in \Bbb{Z}/\ell \Bbb{Z}$ and partitions $\lambda \subseteq \mu$ such that $\mu / \lambda$ is a single box. Define
\begin{equation}
\begin{array}{l}
\text{$A_{\bar i}(\lambda) \hspace{-0.1cm} := \hspace{-0.1cm} \{\text{boxes } b : b \notin \lambda, b \text{ has color }
\bar i \text{ and } \lambda \cup b \text{ is a partition} \},$} \\
\text{$R_{\bar i}(\lambda) \hspace{-0.1cm} := \hspace{-0.1cm} \{\text{boxes } b: b \in \lambda, b \text{ has color }
\bar i \text{ and } \lambda \backslash b \text{ is a partition} \},$} \\
\text{$N_{\bar i}^l(\mu / \lambda) \hspace{-0.1cm} := \hspace{-0.1cm} | \{ b \in R_{\bar i}(\lambda) : \hspace{-0.05cm} b \text{ to the left of } \mu / \lambda \}|
\hspace{-0.1cm} - \hspace{-0.1cm}
| \{ b \in A_{\bar i}(\lambda) : b \text{ to the left of } \mu / \lambda \}|
,$} \\
\text{$N_{\bar i}^r(\mu / \lambda) \hspace{-0.1cm} := \hspace{-0.1cm} | \{ b \in R_{\bar i}(\lambda) : b \text{ to the right of } \mu / \lambda \}|
\hspace{-0.1cm} - \hspace{-0.1cm}
| \{ b \in A_{\bar i}(\lambda) : b \text{ to the right of } \mu / \lambda \}|$}
\end{array}
\end{equation}
to be the set of \emph{addable boxes of color $\bar i$}, the set of \emph{removable boxes of color $\bar i$}, the \emph{left removable-addable difference}, and the \emph{right removable-addable difference}, respectively.
\begin{Theorem} \emph{(see \cite[Theorem 10.6]{ariki:2000})}
\label{MM_Fock_th}
There is an action of $U'_v(\widehat{\mathfrak{sl}}_\ell)$ on ${\bf F}_v$ determined by
\begin{align} \label{MM-a}
E_{\bar i}|\lambda\rangle &:=
\sum_{ \overline{c}(\lambda / \mu)={\bar i}} v^{-N_{\bar i}^r(\lambda / \mu )} |\mu\rangle
\qquad \quad \hbox{and}
& F_{\bar i}|\lambda\rangle &:= \sum_{ \overline{c}(\mu / \lambda)={\bar i}} v^{N_{\bar i}^l(\mu / \lambda )} |\mu\rangle,
\end{align}
where $\overline c(\lambda / \mu)$ denotes the color of $\lambda / \mu$ and the sum is over
partitions $\mu$ which differ from $\lambda$ by removing (respectively adding) a
single $\bar i$-colored box.
\end{Theorem}
As a $U'_v(\widehat{\mathfrak{sl}}_\ell)$-module, ${\bf F}_v$ is isomorphic to an infinite direct sum of copies of the basic representation $V(\Lambda_0)$. Using the grading of ${\bf F}_v$ where $|\lambda \rangle$ has degree $|\lambda|$, the highest weight vectors in ${\bf F}_v$ occur in degrees divisible by $\ell$, and the number of highest weight vectors in degree $\ell k$ is the number of partitions of $k$. Then ${\bf F}_v\cong V(\Lambda_0) \otimes \mathbb{C}[x_1, x_2, \ldots]$,
where $x_k$ has degree $\ell k$, and $U'_v(\widehat{\mathfrak{sl}}_\ell)$ acts trivially on the second factor (see \cite[Prop.\ 2.3]{KMS:1995}). Note that we are working with the `derived' quantum group $U_v'(\widehat{\mathfrak{sl}}_\ell)$, not the `full' quantum group $U_v(\widehat{\mathfrak{sl}}_\ell)$, which is why there are no $\delta$-shifts in the summands of ${\bf F}_v$.
\begin{Comment} Comparing with \cite[Chapter 10]{ariki:2000},
our $N_{\bar i}^l(\mu/\lambda)$ is equal to Ariki's $-N_{\bar i}^a(\mu /\lambda)$ and
our $N_{\bar i}^r(\mu/\lambda)$ is equal to Ariki's $-N_{\bar i}^b(\mu /\lambda)$. However, these numbers play a slightly different role in Ariki's work, which is explained by a different choice of conventions.
\end{Comment}
\section{Universal Verma modules} \label{UV-section}
The purpose of this section is to construct a family of representations which are
universal Verma modules in the sense that each can be ``evaluated" to obtain any given Verma module. This notion was defined by Kashiwara \cite{Kashiwara:1985} in the classical case, and was studied in the quantum case by Kamita \cite{Kamita}.
\subsection{Rational universal Verma modules} \label{rat-UV} \label{twisted-UV}
Let $\mathbb{K}:= \Bbb{Q}(q,z_1,z_2, \ldots, z_N)$. This field is isomorphic to the field of fractions of $U_q(\mathfrak{gl}_N)^0$ via the map
\begin{equation}\label{psidefn}
\psi: U_q(\mathfrak{gl}_N)^0 \rightarrow \mathbb{K}
\qquad\hbox{defined by}\qquad \psi(L_i^{\pm 1}) = z_i^{\pm 1}.
\end{equation}
For each $\mu \in \mathfrak{h}_\Bbb{Z}^*$, define a $\Bbb{Q}(q)$-linear
automorphism $\sigma_\mu\colon \mathbb{K} \to \mathbb{K}$ by
\begin{equation}\label{sigmadefn}
\begin{aligned}
\sigma_\mu(z_i):= q^{(\mu, \varepsilon_i)} z_i,
\qquad\hbox{for $1\le i\le N$,}
\end{aligned}
\end{equation}
where $(\cdot, \cdot)$ is the inner product on $\mathfrak{h}_\Bbb{Z}^*$ defined by $(\varepsilon_i, \varepsilon_j)= \delta_{i,j}$.
Let $ \mathbb{K}_\mu = \text{span}_\mathbb{K} \{ v_{\mu+} \}$ be the one dimensional vector space over $\mathbb{K}$ with
basis vector $v_\mu^+$ and
$U_q(\mathfrak{gl}_N)^{\geq 0}$ action given by
\begin{equation}
X_i \cdot v_{\mu +}=0,\quad\hbox{for $1 \leq i \leq N-1$, \quad and}
\qquad
a \cdot v_{\mu+}=\sigma_\mu(\psi (a)) v_{\mu+},\quad\hbox{for $a \in U_q(\mathfrak{gl}_N)^{0}$.}
\end{equation}
The {\it $\mu$-shifted rational universal Verma module} ${}^\mu \widetilde M$ is the $U_q(\mathfrak{gl}_N)$-module
\begin{equation} \label{ratUV-def}
{}^\mu \widetilde M:= U_q(\mathfrak{gl}_N) \otimes_{U_q(\mathfrak{gl}_N)^{\geq 0}} \mathbb{K}_\mu.
\end{equation}
The universal Verma module ${}^\mu \widetilde M$ is actually a module over $U_q(\mathfrak{gl}_N) \otimes_{U_q(\mathfrak{gl}_N)^0} \widetilde U_q(\mathfrak{gl}_N)^0 $, where $ \widetilde U_q(\mathfrak{gl}_N)^0$ is the field of fractions of $U_q(\mathfrak{gl}_N)^0$. However, if we identify $ \widetilde U_q(\mathfrak{gl}_N)^0$ with
$\mathbb{K}$ using the map $\psi$, the action of $\widetilde U_q(\mathfrak{gl}_N)^0 $ on ${}^\mu \widetilde M$
is not by multiplication, but rather is twisted by the automorphism $\sigma_\mu$. It is to keep track of
the difference between the action of $ U_q(\mathfrak{gl}_N)^0$ and multiplication that we use different
notation for the generators of $\mathbb{K}$ and $U_q(\mathfrak{gl}_N)^0$ (that is, $z_i$ versus $L_i$).
\subsection{Integral universal Verma modules} \label{int-UV}
The field $\mathbb{K}$ contains an $\mathcal{A}$-subalgebra
\begin{equation}
\mathcal{R} \quad\hbox{generated by}\qquad
z_i^{\pm 1}\quad\hbox{and}\quad
\left[
\begin{array}{c}
z_i; c \\
k
\end{array}
\right]
\qquad(1 \leq i \leq N, c \in \Bbb{Z}, k \in \Bbb{Z}_{>0}),
\end{equation}
which is isomorphic to $U_q^\mathcal{A}(\mathfrak{gl}_N)^0$ via the restriction of the map $\psi$ in
\eqref{psidefn}.
The {\it integral universal Verma module} ${}^\mu \widetilde M^\mathcal{R}$ is the
$U_q^\mathcal{A}(\mathfrak{gl}_N)$-submodule
of ${}^\mu \widetilde M$ generated by $v_{\mu+}$.
By restricting \eqref{ratUV-def},
\begin{equation}
{}^\mu \widetilde M^\mathcal{R}= U_q^\mathcal{A}(\mathfrak{gl}_N) \otimes_{U_q^\mathcal{A}(\mathfrak{gl}_N)^{\geq 0}} \mathcal{R}_\mu,
\end{equation}
where $\mathcal{R}_\mu$ is the $\mathcal{R}$-submodule of $ \mathbb{K}_\mu$ spanned by $v_{\mu+}$. In particular, ${}^\mu \widetilde M^\mathcal{R}$ is a free $\mathcal{R}$-module.
\subsection{Evaluation} \label{eval}
Let $\text{ev}_\lambda^\mathcal{R}: \mathcal{R} \rightarrow \mathcal{A}$ be the map defined by
\begin{equation} \label{ARE}
\text{ev}^\mathcal{R}_\lambda(z_i) =q^{(\lambda, \varepsilon_i)} \qquad\hbox{and}\qquad
\text{ev}^\mathcal{R}_\lambda
\left[
\begin{array}{c}
z_i; c \\
n
\end{array}
\right] =
\left[
\begin{array}{c}
q^{(\lambda, \varepsilon_i)}; c \\
n
\end{array}
\right],
\end{equation}
where $(\cdot, \cdot)$ is the inner product on $\mathfrak{h}^*$ defined by $(\varepsilon_i, \varepsilon_j)=\delta_{i,j}$.
There is a surjective $U_q^\mathcal{A}(\mathfrak{gl}_N)$-module homomorphism
``evaluation at $\lambda$''
\begin{equation}
\text{ev}_\lambda: {}^\mu \widetilde M^\mathcal{R} \rightarrow M^\mathcal{A}(\mu+\lambda)
\quad\hbox{defined by}\quad
\text{ev}_\lambda(a \cdot v_{\mu+}):= a \cdot v_{\mu+\lambda}, \quad \text{for all $a \in U^\mathcal{A}_q(\mathfrak{gl}_N)$.}
\end{equation}
For fixed $\lambda$, the maps $\text{ev}_\lambda^\mathcal{R}$ and $\text{ev}_\lambda$ extend to a map from the subspace of $\mathbb{K}$ and ${}^\mu \widetilde M ={}^\mu \widetilde M^\mathcal{R} \otimes_\mathcal{R} \mathbb{K}$ respectively where no denominators evaluate to 0. Where it is clear we denote both these extended maps by $\text{ev}_\lambda$.
\begin{Example} Computing the action of $L_i$ on $v_{\mu+}$ and $v_{\mu+\lambda}$,
\begin{equation}
L_i \cdot v_{\mu+}= q^{(\mu, \varepsilon_i)} z_i v_{\mu+},
\qquad\hbox{and}\qquad
\begin{array}{rl}
L_i \cdot v_{\mu+\lambda}
&= \text{ev}_\lambda (q^{(\mu, \varepsilon_i)} z_i ) v_{\mu+\lambda} \\
&= q^{(\mu, \varepsilon_i)}q^{(\lambda, \varepsilon_i)} v_{\mu+\lambda} = q^{(\mu+\lambda, \varepsilon_i)} v_{\mu+\lambda}.
\end{array}
\end{equation}
\end{Example}
\subsection{Weight decompositions} \label{dec}
Let $\widetilde V$ be a $U_q(\mathfrak{gl}_N) \otimes_\mathcal{A} \mathcal{R}$-module. For each $\nu \in \mathfrak{h}_\Bbb{Z}^*$, we define the \emph{$\nu$-weight space} of $\widetilde{V}$ to be
\begin{equation}
\widetilde{V}_\nu := \{ v \in \widetilde V : L_i \cdot v = q^{(\nu, \varepsilon_i)} z_i v \}.
\end{equation}
The universal Verma module ${}^\mu \widetilde M^\mathcal{R}$ is a $U_q(\mathfrak{gl}_N) \otimes_\mathcal{A} \mathcal{R}$-module, where the second factor acts as multiplication. The weight space ${}^\mu \widetilde M_\eta \neq 0$ if and only if $\eta=\mu-\nu$ with $\nu$ in the positive part $Q^+$ of the root lattice. These non-zero weight spaces and the weight decomposition of ${}^\mu \widetilde M$ can be described explicitly by
\begin{equation}
{}^\mu \widetilde M_{\mu-\nu}^\mathcal{R} = U^\mathcal{A}_q(\mathfrak{gl}_N)^{< 0}_{-\nu} \cdot \mathcal{R}_\mu
\qquad\hbox{and}\qquad
{}^\mu \widetilde M^\mathcal{R} = \bigoplus_{\nu\in Q^+} {}^{\mu} \widetilde M^\mathcal{R}_{\mu-\nu}.
\end{equation}
Here $U_q^\mathcal{A}(\mathfrak{gl}_N)^{< 0}_{-\nu}$ is defined using the grading of $U_q(\mathfrak{gl}_N)^{<0}$ with
$F_i\in U_q(\mathfrak{gl}_N)^{<0}_{-(\varepsilon_i-\varepsilon_{i+1})}$.
\subsection{Tensor products} \label{universal-tensor} Let $\widetilde V$ be a $U_q^\mathcal{A}(\mathfrak{gl}_N) \otimes_\mathcal{A} \mathcal{R}$-module and $W$ a $U_q^\mathcal{A}(\mathfrak{gl}_N)$-module. The tensor product $\widetilde V \otimes_\mathcal{A} W$ is a $U_q^\mathcal{A}(\mathfrak{gl}_N) \otimes_\mathcal{A} \mathcal{R}$-module, where the first factor acts via the usual coproduct and the second factor acts by multiplication on $\widetilde V$.
In the case when $\widetilde V$ and $W$ both have weight space decompositions, the weight spaces of $\widetilde V \otimes_\mathcal{A} W$ are
\begin{equation}
(\widetilde V \otimes_\mathcal{A} W)_\nu = \bigoplus_{\gamma+\eta=\nu} \widetilde V_\gamma \otimes_\mathcal{A} W_\eta.
\end{equation}
We also need the following:
\begin{Proposition} \label{rat_dec}
The tensor product of a universal Verma module with a Weyl module satisfies
\begin{equation}
\left({}^\mu \widetilde M^\mathcal{R} \otimes_\mathcal{A} \Delta^\mathcal{A}(\nu) \right)\otimes_\mathcal{R} \mathbb{K}
\cong \left( \bigoplus_{\gamma} ({}^{\mu + \gamma} \widetilde M^\mathcal{R})^{\oplus \mathrm{dim} \Delta^\mathcal{A}(\nu)_\gamma} \right) \otimes_\mathcal{R} \mathbb{K}.
\end{equation}
\end{Proposition}
\begin{proof}
Fix $\nu \in P^+$. In general, $M(\lambda+\mu) \otimes \Delta(\nu)$ has a Verma filtration
(see, for example, \cite[Theorem 2.2]{Jantzen:1979}) and
if $\lambda+\mu+\gamma$ is dominant for all $\gamma$ such that
$\Delta(\nu)_\gamma\ne 0$ then
\begin{equation}
M(\lambda+\mu) \otimes \Delta(\nu)
\cong \bigoplus_{\gamma} M(\lambda+\mu+\gamma)^{\oplus \mathrm{dim} \Delta(\nu)_\gamma},
\end{equation}
which can be seen by, for instance, taking central characters.
The proposition follows since this is true for a Zariski dense set of weights $\lambda$.
\end{proof}
\section{The Shapovalov form and the Shapovalov determinant} \label{Sform}
\subsection{The Shapovalov form} \label{S_over_A}
The \emph{Cartan involution} $\omega: U_q(\mathfrak{gl}_N) \rightarrow U_q(\mathfrak{gl}_N)$ is the $\Bbb{Q}(q)$-algebra anti-involution of $U_q(\mathfrak{gl}_N)$ defined by
\begin{equation}
\omega(L_i^{\pm 1})=L_i^{\pm 1},
\qquad
\omega(X_i)= Y_iL_iL_{i+1}^{-1},
\qquad
\omega(Y_i) = L_i^{-1}L_{i+1} X_i.
\end{equation}
The map $\omega$ is also a co-algebra involution.
An \emph{$\omega$-contravariant} form on a $U_q(\mathfrak{gl}_N)$-module $V$
is a symmetric bilinear form $(\cdot, \cdot)$ such that
\begin{equation} \label{contrav}
(u,a \cdot v)=(\omega(a) \cdot u, v),\qquad
\hbox{for $u, v \in V$ and $a \in U_q(\mathfrak{gl}_N)$.}
\end{equation}
It follows by the same argument used in the classical case \cite{Shapovalov:1972} that there is an
$\omega$-contravariant form (the Shapovalov form) on each Verma module $M(\lambda)$
and this is unique up to rescaling. The radical of $(\cdot, \cdot)$ is the maximal proper submodule of $M(\lambda)$, so $\Delta(\lambda)=M(\lambda)/\mathrm{Rad}( \cdot, \cdot)$ for all $\lambda \in P^+$. In particular, $(\cdot, \cdot)$ descends to an $\omega$-contravariant form on $\Delta(\lambda)$.
Since $\omega$ fixes $U_q^\mathcal{A}(\mathfrak{gl}_N) \subseteq U_q(\mathfrak{gl}_N)$, there is a well defined
notion of an $\omega$-contravariant form on a $U_q^\mathcal{A}(\mathfrak{gl}_N)$ module. In particular,
the restriction of the Shapovalov form on $\Delta(\lambda)$ to $\Delta^\mathcal{A}(\lambda)$
is $\omega$-contravariant.
\subsection{Universal Shapovalov forms}
There are surjective maps of $\mathcal{A}$-algebras $p_-: U_q^\mathcal{A}(\mathfrak{gl}_N)^{<0} \rightarrow \Bbb{Q}(q)$ and $p_+: U_q^\mathcal{A}(\mathfrak{gl}_N)^{>0} \rightarrow \Bbb{Q}(q)$ defined by $p_-(F_i)=0$ and $p_+(E_i)=0$, for $1 \leq i \leq N$. Using the triangular decomposition \eqref{int-triangle}, there is an $\mathcal{A}$-linear surjection
\begin{equation}
\pi_0 := p_- \otimes \text{Id} \otimes p_+: U_q^\mathcal{A}(\mathfrak{gl}_N) \cong U_q^\mathcal{A}(\mathfrak{gl}_N)^{<0} \otimes_\mathcal{A} U^\mathcal{A}_q(\mathfrak{gl}_N)^0 \otimes_\mathcal{A} U^\mathcal{A}_q(\mathfrak{gl}_N)^{>0} \rightarrow U_q^\mathcal{A}(\mathfrak{gl}_N)^0.
\end{equation}
The {\it standard universal Shapovalov form} is the $\mathcal{R}$-bilinear form $(\cdot, \cdot)_{{}^\mu \widetilde M^\mathcal{R}}: {}^\mu \widetilde M^\mathcal{R} \otimes {}^\mu \widetilde M^\mathcal{R} \rightarrow \mathcal{R}$ defined by
\begin{equation}
(a_1 \cdot v_{\mu+},a_2 \cdot v_{\mu+})_{{}^\mu \widetilde M^\mathcal{R}}
= \big(\sigma_\mu \circ \psi \circ \pi_0\big) (\omega(a_2) a_1)
\end{equation}
for all $a_1,a_2 \in U_q^\mathcal{R}(\mathfrak{gl}_N)^{<0}$.
Here $\psi$ and $\sigma_\mu$ are as in \eqref{psidefn} and \eqref{sigmadefn}.
Since
\begin{equation}
(a_1a_2 \cdot v_{\mu+},a_3 \cdot v_{\mu+})_{{}^\mu \widetilde M^\mathcal{R}}
= \big(\sigma_\mu \circ \psi \circ \pi_0\big) (\omega(a_2)\omega(a_1)a_3)
=(a_2 \cdot v_{\mu+},\omega(a_1)a_3 \cdot v_{\mu+})_{{}^\mu \widetilde M^\mathcal{R}}
\end{equation}
for $a_1,a_2,a_3 \in U_q(\mathfrak{gl}_N),$
the form
$(\cdot, \cdot)_{{}^\mu \widetilde M^\mathcal{R}}$ is $\omega$-contravariant. As with the usual Shapovalov form, distinct weight spaces are orthogonal, where weight spaces are defined as in Section \ref{dec}.
Evaluation
at $\lambda$ gives an $\mathcal{A}$-valued $\omega$-contravariant form
$(\cdot, \cdot)_{M^\mathcal{A}(\mu+\lambda)}$ on $M^\mathcal{A}(\mu+\lambda)$ by
\begin{equation} \label{eve}
(\text{ev}_\lambda(u_1), \text{ev}_\lambda(u_2))_{M^\mathcal{A}(\mu+\lambda)}
= \text{ev}_\lambda \left((u_1,u_2)_{{}^\mu \widetilde M^\mathcal{R}}\right),
\qquad\hbox{for $u_1,u_2 \in {}^\mu \widetilde M^\mathcal{R}$.}
\end{equation}
The form $(\cdot, \cdot)_{{}^\mu \widetilde M^\mathcal{R}}$ can be extended by linearity to an $\omega$-contravariant form $(\cdot, \cdot)_{{}^\mu \widetilde M}$ on ${}^\mu \widetilde M$.
\subsection{The Shapovalov determinant} \label{SD}
Let $\widetilde V$ be a $(U_q^\mathcal{A}(\mathfrak{gl}_N) \otimes_\mathcal{A} \mathcal{R})$-module with a chosen
$\omega$-contravariant form.
Let $B_\eta$ be an $\mathcal{R}$ basis for the $\eta$-weight space $\widetilde V_\eta$ of $\widetilde V$.
Let $\det \widetilde V_{B_\eta}$ be the determinant of the form
evaluated on the basis $B_\eta$. Changing the basis $B_\eta$ changes the determinant by a unit
in $\mathcal{R}$ and we sometimes write $\det\widetilde V_\eta$ to mean the determinant calculated on an unspecified basis ($\det\widetilde V_\eta$ which is only defined up to multiplication by unit in $\mathcal{R}$).
The {\it Shapovalov determinant} is
\begin{equation}
\det \widetilde M^\mathcal{R}_\eta := \det((b_i,b_j)_{\widetilde M^\mathcal{R}})_{b_i,b_j\in B_\eta}.
\end{equation}
Define the {\it partition function} $p\colon \mathfrak{h}^*\to \mathbb{Z}_{\ge 0}$ by
\begin{equation}
p(\gamma):= \mathrm{dim} M(0)_\gamma.
\end{equation}
Then
$p(\gamma)= \mathrm{dim} M(\lambda)_{\gamma+\lambda}$
for any $\lambda$, and $\eta \not\in Q^-$ implies that $p(\eta)=0$ and $\det \widetilde M^\mathcal{R}_\eta=1$.
\begin{Theorem} \label{fulldet} \emph{(see \cite[Proposition 1.9A]{KD:1991}, \cite[Theorem 3.4]{KL:1997}, \cite{Shapovalov:1972})}
For any weight $\eta$,
\begin{equation}
\det \widetilde M^\mathcal{R}_\eta
= c_\eta \prod_{\tiny \begin{array}{c} 1 \leq i < j \leq N \\ m >0 \end{array}}
{\Big( }z_iz_j^{-1} - q^{2m+2i-2j} z_i^{-1}z_j {\Big)}^{p(\eta+m\varepsilon_i-m\varepsilon_j)},
\end{equation}
where $c_\eta$ is a unit in $\mathcal{R} \otimes_\mathcal{A} \Bbb{Q}(q) = \Bbb{Q}(q)[ z_1^{\pm1}, \ldots, z_N^{\pm 1}]$.
\end{Theorem}
\begin{Proposition} \label{mu-s}
Fix $\mu, \eta \in \mathfrak{h}_\mathbb{Z}^*$ with $
\eta-\mu \in Q^-.$ Choose an $\mathcal{A}$-basis $B_{\eta-\mu}$ for $U_q^\mathcal{A}(\mathfrak{gl}_N)_{\eta-\mu}.$ Consider the $\mathcal{R}$-bases
$\widetilde B_{\eta-\mu} := \{ b \cdot v_+ \ |\ b \in B_{\eta-\mu} \}$ for $\widetilde M^{\mathcal{R}}_{\eta-\mu}$
and
${}^\mu \widetilde B_\eta : = \{ b \cdot {v_{\mu+}} \ |\ b \in B_{\eta-\mu} \}$
for $ {}^\mu \widetilde M^\mathcal{R}_\eta $. Then
$\det {}^\mu \widetilde M^\mathcal{R}_{({}^\mu \widetilde B_\eta)}
= \sigma_\mu \big(\det \widetilde M^\mathcal{R}_{\widetilde B_{\eta-\mu}}\big)$.
\end{Proposition}
\begin{proof}
For $b,b' \in B_{\eta-\mu}$,
\begin{align}
(b \cdot v_{\mu+},b' \cdot v_{\mu+})_{{}^\mu \widetilde M^\mathcal{R}}
= \sigma_\mu \circ \psi \circ \pi_0 (\omega(b') b)
= \sigma_\mu \big( (b \cdot v_{0+},b' \cdot v_{0+})_{\widetilde M^\mathcal{R}}\big).
\end{align}
The result follows by taking determinants.
\end{proof}
\subsection{Contravariant forms on tensor products} \label{SonT}
If $V$ and $W$ are $U_q^\mathcal{A}(\mathfrak{gl}_N)$-modules with $\omega$-contravariant forms
$(\cdot, \cdot)_V$ and $(\cdot, \cdot)_W$,
define an $\mathcal{A}$-bilinear form $(\cdot,\cdot)_{W \otimes V}$ by
$(w_1\otimes v_1, w_2\otimes v_2)_{W \otimes V}
= (w_1,w_2)_W(v_1,v_2)_{V}$.
Similarly, for
a $U_q^\mathcal{A}(\mathfrak{gl}_N) \otimes_\mathcal{A} \mathcal{R}$ module $\widetilde W$ with $\mathcal{R}$-bilinear
$\omega$-contravariant form $(\cdot, \cdot)_{\widetilde W}$, define
a $\mathcal{R}$-bilinear form $(\cdot, \cdot)_{\widetilde W \otimes_{\Bbb{Q}(q)} V}$
on $\widetilde W \otimes_{\Bbb{Q}(q)} V$ by
\begin{equation}\label{prodformdefn}
(u_1 \otimes v_1, u_2 \otimes v_2)_{\widetilde W \otimes_{\Bbb{Q}(q)} V}
= (u_1,u_2)_{\widetilde W} (v_1,v_2)_V.
\end{equation}
Since $\omega$ is a coalgebra involution (i.e., $\Delta(\omega(a))= (\omega \otimes \omega) \Delta(a)$, for $a \in U_q(\mathfrak{gl}_N)$),
the forms $(\cdot,\cdot)_{V \otimes W}$ and
$(\cdot, \cdot)_{{}^\mu \widetilde M \otimes_{\Bbb{Q}(q)} V}$
are $\omega$-contravariant.
In the case when $\widetilde W= {}^\mu \widetilde M^\mathcal{R}$, evaluation of the
$\omega$-contravariant form $(\cdot, \cdot)_{{}^\mu \widetilde M^\mathcal{R} \otimes_\mathcal{A} V}$
at $\lambda$ gives an $\omega$-contravariant form
$(\cdot, \cdot)_ {M^\mathcal{A}(\mu+\lambda) \otimes_\mathcal{A} V}$:
\begin{equation} \label{ev-tensor}
\begin{aligned}
(u_1 \otimes v_1, u_2 \otimes v_2)_ {M^\mathcal{A}(\mu+\lambda) \otimes_\mathcal{A} V} &= \text{ev}_\lambda \left((u_1 \otimes v_1, u_2 \otimes v_2)_ {{}^\mu \widetilde M^\mathcal{R} \otimes_\mathcal{A} V} \right) \\& = (\text{ev}_\lambda(u_1) \otimes v_1, \text{ev}_\lambda(u_2) \otimes v_2)_{M(\mu+\lambda) \otimes_\mathcal{A} V},
\end{aligned}
\end{equation}
for $u_1, u_2 \in {}^\mu \widetilde M$ and $v_1, v_2 \in V$.
As in Section \ref{eval}, this evaluation can be extended to the $\mathcal{A}$-submodule of the
rational module where no denominators evaluate to zero.
\section{The Misra-Miwa formula for $F_{\bar i}$ from $U_q^\mathcal{A}(\mathfrak{gl}_N)$ representation theory} \label{makeops}
Let us prepare the setting for our main result (Theorem 6.1). Fix $\ell \geq 2$ and a partition $\lambda$. Let $N$ a positive integer greater than the
number of parts of $\lambda$. All calculations below are in terms of representations
of $U_q^\mathcal{A}(\mathfrak{gl}_N)$.
$\bullet$ Let $V=\Delta^\mathcal{A}(\varepsilon_1)$ be the standard $N$-dimensional module. Since $\Delta^\mathcal{A}(\lambda) \otimes_\mathcal{A} \Bbb{Q}(q)= \Delta(\lambda)$, Equation \eqref{plusboxes} implies
\begin{equation} \label{tens}
\left( \Delta^{\mathcal{A}}(\lambda) \otimes_\mathcal{A} V \right) \otimes_{\mathcal{A}} \Bbb{Q}(q) \simeq \bigoplus \Delta^{\mathcal{A}}(\lambda+\varepsilon_{k_j}) \otimes_{\mathcal{A}} \Bbb{Q}(q),
\end{equation}
where the sum is over those indices $1 = k_1 < k_2 < \cdots < k_{m_\lambda} \leq N$ for which $\lambda+\varepsilon_{k_j}$ is a partition.
For ease of notation let $\mu^{(j)}= \lambda+\varepsilon_{k_j}$.
$\bullet$ Fix an $\mathcal{A}$-basis $\{ v_1, \ldots, v_N \}$ of $V$ where $v_k$ has weight $\varepsilon_k$ and $Y_i(v_k) = \delta_{i,k} v_{k+1}$. Recursively define singular weight vectors $v_{\mu^{(j)}}$ in $\left(\Delta^\mathcal{A}(\lambda) \otimes V \right) \otimes_\mathcal{A} \Bbb{Q}(q)$ by:
\begin{enumerate}
\item $v_{\mu^{(1)}}= v_\lambda \otimes v_1.$
\item For each $k$, the submodule $W_k$ of $(\Delta(\lambda) \otimes_\mathcal{A} V) \otimes_\mathcal{A} \Bbb{Q}(q)$ generated by $\{ v_\lambda \otimes v_i\ |\ 1 \leq i \leq k \}$ contains all weight vectors of $(\Delta(\lambda) \otimes_\mathcal{A} V) \otimes_\mathcal{A} \Bbb{Q}(q)$ of weight greater than or equal to $\lambda+\varepsilon_k$. Thus, using \eqref{tens}, for each $1 \leq j \leq m_\lambda$ there is a
one-dimensional space of singular vectors of weight $\mu^{(j)}$ in $W_{k_j}$,
and this is not contained in $W_{k_{j-1}}$
(since $k_j> k_{j-1}$).
This implies that there unique singular vector $v_{\mu^{(j)}}$ of weight $\mu^{(j)}$ in
\begin{equation} \label{eqs2}
v_\lambda \otimes v_{k_j} + \bigoplus_{1 \leq i <j} U_q(\mathfrak{gl}_N) v_{\mu^{(i)}}
\subseteq \left( \Delta^\mathcal{A}(\lambda)\otimes_\mathcal{A} V \right) \otimes_\mathcal{A} \Bbb{Q}(q),
\end{equation}
where we recall that $U_q(\mathfrak{gl}_N)= U_q^\mathcal{A}(\mathfrak{gl}_N) \otimes_\mathcal{A} \Bbb{Q}(q)$.
\end{enumerate}
$\bullet$ There is a unique $\omega$-contravariant form on $\Delta^\mathcal{A}(\lambda)$ normalized so that $(v_\lambda, v_\lambda)=1$ and a unique $\omega$-contravariant form on $V$ normalized so that $(v_1,v_1)=1$. As in section \ref{SonT}, define a $\omega$-contravariant form on $\left( \Delta^{\mathcal{A}}(\lambda) \otimes_\mathcal{A} V \right) \otimes_{\mathcal{A}} \Bbb{Q}(q)$ by $(u_1 \otimes w_1, u_2 \otimes w_2)= (u_1,u_2)(w_1,w_2)$. For each $ 1\leq j \leq m_\lambda$, define an element $r_{j}(\lambda) \in \Bbb{Q}(q)$ by
\begin{equation}
r_{j}(\lambda) := (v_{\mu^{(j)}}, v_{\mu^{(j)}}).
\end{equation}
\begin{Theorem} \label{main} The Misra-Miwa operators $F_{\bar i}$ from Section \ref{MM_section} satisfy
\begin{equation} \label{Fdeftwo}
\displaystyle F_{\bar i} |\lambda\rangle = \sum_{\bar c(b^{(j)})= \bar i}v^{\text{val}_{\phi_{2\ell}} r_{j}(\lambda)} |\mu^{(j)}\rangle,
\end{equation}
where $b^{(j)}$ is the box $\mu^{(j)} / \lambda$, $\bar c (b^{(j)})$ is the color of box $b^{(j)}$ as in Figure \ref{partition_bij}, $\phi_{2\ell}$ is the $2\ell^{th}$ cyclotomic polynomial in $q$ and $\text{val}_{\phi_{2\ell}} r$ is the number of factors of $\phi_{2\ell}$ in the numerator of $r$ minus the number of factors of $\phi_{2\ell}$ in the denominator of $r$.
\end{Theorem}
The proof of Theorem \ref{main} will occupy the rest of this section. We will first prove a similar statement, Proposition \ref{the_vals2}, where the role of the Weyl modules is played by the universal Verma modules from Section \ref{UV-section}. For ease of notation, let $\widetilde M^\mathcal{R}$ denote the module ${}^0 \widetilde M^\mathcal{R}$ from section \ref{int-UV}.
\begin{Definition} \label{make-the-high-vectors} \label{mhv}
Recursively define singular weight vectors $v_{\varepsilon_k+} \in \left( \widetilde M^\mathcal{R} \otimes_\mathcal{A} V \right) \otimes_\mathcal{R} \mathbb{K}$ and elements $s_k \in \mathbb{K}$ for $1 \leq k \leq N$ by
\begin{enumerate}
\item $v_{\varepsilon_1+}= v_+ \otimes v_1.$
\item Since $\{v_+ \otimes v_j\ |\ 1\le j\le N\}$ generates $\widetilde M^\mathcal{R} \otimes_\mathcal{A} V$ as a $U_q^\mathcal{A}(\mathfrak{gl}_N)^{\leq 0}$ module, Proposition \ref{rat_dec} implies that, for each $1 \leq k \leq N$, there is a unique singular vector $v_{\varepsilon_{k}\tiny +}$ in
$\displaystyle v_+ \otimes v_k
+ \bigoplus_{1 \leq j <k} U_q^\mathbb{K}(\mathfrak{gl}_N) v_{\varepsilon_j+}
\subseteq \left( \widetilde M^\mathcal{R} \otimes_\mathcal{A} V \right) \otimes_\mathcal{R} \mathbb{K}$, where
$U_q^\mathbb{K}(\mathfrak{gl}_N):= U_q(\mathfrak{gl}_N) \otimes_{\Bbb{Q}(q)} \mathbb{K}$ and the factor of $\mathbb{K}$ acts by multiplication on $\widetilde M^\mathcal{R}$.
\end{enumerate}
Let $s_k = (v_{\varepsilon_k\tiny+}, v_{\varepsilon_k \tiny+})$.
\end{Definition}
The $s_k$ are quantized versions of the Jantzen numbers
first calculated in \cite[Section 5]{Jantzen:1974} and quantized in \cite{Wiesner}. It follows immediately from the definition that $s_1=1$.
\begin{Lemma} \label{equal-dets}
For any weight $\eta$, up to multiplication by a power of $q$,
\begin{equation}
\prod_{1 \leq k \leq N} s_k^{p(\eta-\varepsilon_k)} =
\prod_{1 \leq k \leq N} \frac{\det \widetilde M^\mathcal{R}_{\eta-\varepsilon_k}}{\sigma_{\varepsilon_k} \det \widetilde M^\mathcal{R}_{\eta-\varepsilon_k}},
\end{equation}
where, as in Section \ref{SD}, $\det \widetilde M^\mathcal{R}_{\eta-\varepsilon_k}$ is the determinant of the Shapovalov form evaluated on an $\mathcal{R}$-basis for the $\eta-\varepsilon_k$ weight space of $\widetilde M^\mathcal{R}$.
\end{Lemma}
\begin{Comment}
In order for Lemma \ref{equal-dets} to hold as stated, for each $1 \leq k \leq N$, one must calculate the
$\det \widetilde M^\mathcal{R}_{\eta-\varepsilon_k}$ in the numerator and denominator with respect
to the same $\mathcal{R}$-basis.
The power of $q$ which appears depends on this choice of $\mathcal{R}$-bases.
\end{Comment}
\begin{proof}[Proof of Lemma \ref{equal-dets}]
For each $\gamma \in \hbox{span}_{\mathbb{Z}_{\le 0}}(R^+)$
fix an $\mathcal{R}$-basis $B_\gamma$ for $U_q^\mathcal{R}(\mathfrak{gl}_N)^{<0}_\gamma$.
Consider the following three $\mathbb{K}$-bases
for $\left( ( \widetilde M^\mathcal{R} \otimes_\mathcal{A} V)_\eta \right) \otimes_\mathcal{R} \mathbb{K}$:
\begin{equation}
\begin{aligned}
&A_\eta := \{ (b \cdot v_+) \otimes v_k \ |\ b \in B_{\eta-\varepsilon_k}, 1 \leq k \leq N \}, \\
&C_\eta := \{ b \cdot ( v_+ \otimes v_k) \ |\ b \in B_{\eta-\varepsilon_k}, 1 \leq k \leq N \},
\\
&D_\eta:= \{ b \cdot v_{\varepsilon_k\tiny+} \ |\ b \in B_{\eta-\varepsilon_k}, 1 \leq k \leq N \}.
\end{aligned}
\end{equation}
Let $\det (\widetilde M^\mathcal{R} \otimes_\mathcal{A} V)_B$ denote the determinant of
$(\cdot, \cdot)_{(\widetilde M^\mathcal{R} \otimes_\mathcal{A} V)_\eta}$ calculated on $B$, where
$B$ is one of $A_\eta, C_\eta$ or $D_\eta$. Let $\det {}^\nu \widetilde M^\mathcal{R}_{B_{\eta-\nu}}$
denote $\det{}^\nu \widetilde M^\mathcal{R}_\eta$ calculated with respect to the basis
$B_{\eta-\nu} \cdot v_{\nu+}$.
By the definition of the $\omega$-contravariant form on $\widetilde M^\mathcal{R} \otimes_\mathcal{A} V$ (see Section \ref{universal-tensor}),
\begin{equation} \label{det1}
\det (\widetilde M^\mathcal{R} \otimes V)_{A_\eta} = \prod_{k=1}^N (\det \widetilde M^\mathcal{R}_{B_{\eta-\varepsilon_k}})^{\mathrm{dim} V_{\varepsilon_k}} (\det V_{\varepsilon_k})^{\mathrm{dim} \widetilde M^\mathcal{R}_{\eta-\varepsilon_k}}.
\end{equation}
For $1 \leq k \leq N$, $V_{\varepsilon_k}$ is one dimensional and $\det V_{\varepsilon_k}$ is a power of $q$. Hence,
up to multiplication by a power of $q$, \eqref{det1} simplifies to
\begin{equation} \label{det2}
\det (\widetilde M^\mathcal{R} \otimes_\mathcal{A} V)_{A_\eta}
= \prod_{k=1}^N \det \widetilde M^\mathcal{R}_{B_{\eta-\varepsilon_k}}.
\end{equation}
Notice that $U_q^\mathcal{A} (\mathfrak{gl}_N)^{<0} \cdot v_{\varepsilon_k+}$ is isomorphic to
${}^{\varepsilon_k} \widetilde M$, and $D_\eta$ is the union of $\mathcal{R}$-bases for
each of these submodules.
For each $1 \leq k \leq N$, and each $\eta \in \mathfrak{h}_\Bbb{Z}^*,$ define an $\mathcal{R}$ basis of ${}^{\varepsilon_k} \widetilde M_\eta$ by
\begin{equation}
{}^{\varepsilon_k} \widetilde B_\eta:=
\{ b \cdot v_{\varepsilon_k+} \ |\ b \in B_{\eta-\varepsilon_k} \}.
\end{equation}
Using
$(v_{\varepsilon_k+}, v_{\varepsilon_k+} ) =s_k$,
\begin{equation} \label{newdet}
\det (\widetilde M^\mathcal{R} \otimes V)_{D_\eta}=
\prod_{k=1}^N s_k^{\mathrm{dim}( {}^{\varepsilon_k} \widetilde M^\mathcal{R}_\eta)} \det {}^{\varepsilon_k} \widetilde M^\mathcal{R}_{({}^{\varepsilon_k} \widetilde B_{\eta})} =
\prod_{k=1}^N s_k^{p(\eta-\varepsilon_k)}
\sigma_{\varepsilon_k}(\det \widetilde M^\mathcal{R}_{\widetilde B_{\eta-\varepsilon_k}}),
\end{equation}
where the last equality uses Proposition \ref{mu-s}. Here, as in Section \ref{SD}, $\det {}^{\varepsilon_k} \widetilde M^\mathcal{R}_{({}^{\varepsilon_k} \widetilde B_{\eta})}$ is the Shapovalov determinant calculated with respect to the basis ${}^{\varepsilon_k} \widetilde B_{\eta}$.
The change of basis from $A_\eta$ to $C_\eta$ is unitriangular and
the change of basis from $C_\eta$ to $D_\eta$ is unitriangular.
Thus $\det (\widetilde M^\mathcal{R} \otimes_\mathcal{A} V)_{A_\eta } =\det (\widetilde M^\mathcal{R} \otimes_\mathcal{A} V)_{D_\eta }$, and so the right sides of \eqref{det2} and \eqref{newdet} are equal. The lemma follows from this equality by rearranging.
\end{proof}
\begin{Lemma} \label{get_a_hold} Up to multiplication by a power of $q$,
\begin{equation} s_k = \prod_{1 \leq j < k} \left(
\frac{z_j^{}z_k^{-1}- q^{2 +2j-2k}z_j^{-1}z_k^{}}
{\sigma_{\varepsilon_j}\left( z_j^{}z_k^{-1}- q^{2 +2j-2k}z_j^{-1}z_k^{} \right)}
\right).
\end{equation}
\end{Lemma}
\begin{proof}
Fix $1 \leq k \leq N$. Setting $\eta=\varepsilon_k$ in Lemma \ref{equal-dets}
and applying Theorem \ref{fulldet} we see that, up to multiplication by a power of $q$,
\begin{equation}
\begin{aligned}
\prod_{1 \leq x \leq N} s_x^{p(\varepsilon_k-\varepsilon_x)} &= \hspace{-0.3cm}
\prod_{1 \leq x \leq N} \frac{\det \widetilde M^\mathcal{R}_{\varepsilon_k-\varepsilon_x}}{\sigma_{\varepsilon_x} \det \widetilde M^\mathcal{R}_{\varepsilon_k-\varepsilon_x}} \\
&=
\hspace{-0.3cm}
\prod_{1 \leq x \leq N} \prod_{\tiny \begin{array}{c} 1 \leq i < j \leq N \\ m >0 \end{array}} \hspace{-0.4cm} \left( \hspace{-0.1cm} \frac{c_{\varepsilon_k-\varepsilon_x} \left(z_i^{}z_j^{-1}- q^{2m+2i-2j } z_i^{-1}z_j^{}\right)}{\sigma_{\varepsilon_x}(c_{\varepsilon_k-\varepsilon_x}) \sigma_{\varepsilon_x} \left( z_i^{}z_j^{-1} - q^{2m+2i-2j } z_i^{-1}z_j^{}\hspace{-0.1cm} \right) } \right)^{ \hspace{-0.15cm} p(\varepsilon_k-\varepsilon_x + m\varepsilon_i-m\varepsilon_j)} \hspace{-1.8cm},
\end{aligned}
\end{equation}
where, for each $ 1 \leq x \leq N$, $c_{\varepsilon_k-\varepsilon_x}$ is a unit in
$\Bbb{Q}(q)[z_1^{\pm1}, \ldots, z_N^{\pm1}]$.
The value $p(\varepsilon_k-\varepsilon_x+m\varepsilon_i-m\varepsilon_j)$ is $0$ unless $m=1$ and $x \leq i < j \leq k$.
If $i>x$, then $\sigma_{\varepsilon_x}$ acts as the identity on $z_iz_j^{-1} - q^{2+2i-2j}z_i^{-1}z_j$, so the corresponding factors in the numerator and denominator cancel. Hence we need only consider factors on the right hand side where $m=1$, $i=x$, and $x< j \leq k$. If $x >k$ then $\varepsilon_k-\varepsilon_x \not\in Q^-$, and hence $p(\varepsilon_k-\varepsilon_x)=0$, so on the left hand since we only need to consider those factors where $1 \leq x \leq k$. Up to multiplication by a power of $q$, the expression reduces to
\begin{equation} \label{confusing-product}
\begin{aligned}
\hspace{-0.3cm}
\prod_{1 \leq x \leq k} s_x^{p(\varepsilon_k-\varepsilon_x)} &=
\hspace{-0.3cm}
\prod_{1 \leq x < k}
\hspace{-0.1cm}
\displaystyle \left( \frac{c_{\varepsilon_k-\varepsilon_x}}{\sigma_{\varepsilon_x}(c_{\varepsilon_k- \varepsilon_x})} \right)^{p(\varepsilon_k-\varepsilon_j)}
\hspace{-0.3cm}
\prod_{x< j \leq k}
\hspace{-0.1cm}
\left(
\frac{z_x^{}z_j^{-1}- q^{2 +2x-2j}z_x^{-1}z_j^{}}
{\sigma_{\varepsilon_x} \left( z_x^{}z_j^{-1}- q^{2 +2x-2j}z_x^{-1}z_j^{} \right)}
\right)^{p(\varepsilon_k-\varepsilon_j)}
\hspace{-0.7cm} \\
& = \hspace{-0.2cm} \prod_{1 < j \leq k} \left( \prod_{1 \leq x < j}\frac{z_x^{}z_j^{-1}- q^{2 +2x-2j}z_x^{-1}z_j^{}}{{\sigma_{\varepsilon_x}} \left( z_x^{}z_j^{-1}- q^{2 +2x-2j}z_x^{-1}z_j^{} \right)} \right)^{p(\varepsilon_k-\varepsilon_j)} \hspace{-0.1cm}.
\end{aligned}
\end{equation}
The last two expressions are equal because they are each a product over
pairs $(x,j)$ with $1 \leq x<j \leq k$, and the factors of
$ \displaystyle \frac{c_{\varepsilon_k-\varepsilon_x}}
{\sigma_{\varepsilon_x}(c_{\varepsilon_k- \varepsilon_x})}$ have been dropped because they are powers of $q$. Using the fact that $s_1=1$ and making the change of variables
$j \rightarrow x$ and $x \rightarrow j$ on the right side, \eqref{confusing-product} becomes
\begin{equation}
\begin{aligned}
\prod_{1 < x \leq k} s_x^{p(\varepsilon_k-\varepsilon_x)}
& = \hspace{-0.2cm} \prod_{1 < x \leq k} \left(
\prod_{1 \leq j < x}\frac{z_j^{}z_x^{-1}- q^{2 +2j-2x}z_j^{-1}z_x^{}}
{\sigma_{\varepsilon_j} \left( z_j^{}z_x^{-1}- q^{2 +2j-2x}z_j^{-1}z_x^{} \right)}
\right)^{p(\varepsilon_k-\varepsilon_x)} \hspace{-0.1cm}.
\end{aligned}
\end{equation}
For $k \geq 2$, the lemma now follows by induction. For $k=1$ the result simply says that $s_1=1$, which we already know.
\end{proof}
\begin{figure}
\setlength{\unitlength}{0.4cm}
\begin{center}
\begin{picture}(26,17)
\put(3.65,9.8){$g_1$}
\put(4.65,10.8){$g_2$}
\put(7.65,9.8){$g_3$}
\put(8.65,10.8){$g_4$}
\put(9.65,11.8){$g_5$}
\put(12.65,10.8){$g_6$}
\put(13.65,11.8){$g_7$}
\put(14.65,12.8){$g_8$}
\put(15.65,13.8){$g_9$}
\put(21.45,9.8){$g_{10}$}
\put(22.45,10.8){$g_{11}$}
\put(3,10){\begin{picture}(1,2)
\put(0,2){\line(1,-1){1}}
\put(0,2){\line(-1,-1){1}}
\put(-0.4,0.75){$a_1$}
\end{picture}}
\put(7,10){\begin{picture}(1,2)
\put(0,2){\line(1,-1){1}}
\put(0,2){\line(-1,-1){1}}
\put(-0.4,0.75){$a_3$}
\end{picture}}
\put(12,11){\begin{picture}(1,2)
\put(0,2){\line(1,-1){1}}
\put(0,2){\line(-1,-1){1}}
\put(-0.3,0.75){$b$}
\end{picture}}
\thinlines
\put(12,1){\line(1,1){11}}
\put(11,2){\line(1,1){9}}
\put(10,3){\line(1,1){9}}
\put(9,4){\line(1,1){9}}
\put(8,5){\line(1,1){9}}
\put(7,6){\line(1,1){5}}
\put(6,7){\line(1,1){5}}
\put(5,8){\line(1,1){2}}
\put(4,9){\line(1,1){2}}
\put(14,1){\line(-1,1){10}}
\put(15,2){\line(-1,1){8}}
\put(16,3){\line(-1,1){8}}
\put(17,4){\line(-1,1){8}}
\put(18,5){\line(-1,1){6}}
\put(19,6){\line(-1,1){6}}
\put(20,7){\line(-1,1){6}}
\put(21,8){\line(-1,1){6}}
\put(22,9){\line(-1,1){1}}
\put(23,10){\line(-1,1){1}}
\thicklines
\put(13,0){\line(1,1){13}}
\put(13,0){\line(-1,1){13}}
\put(24,11){\line(-1,1){1}}
\put(23,12){\line(-1,-1){2}}
\put(21,10){\line(-1,1){5}}
\put(16,15){\line(-1,-1){4}}
\put(12,11){\line(-1,1){2}}
\put(10,13){\line(-1,-1){3}}
\put(7,10){\line(-1,1){2}}
\put(5,12){\line(-1,-1){2}}
\put(13,0.05){\line(1,1){13}}
\put(13,0.05){\line(-1,1){13}}
\put(24,11.05){\line(-1,1){1}}
\put(23,12.05){\line(-1,-1){2}}
\put(21,10.05){\line(-1,1){5}}
\put(16,15.05){\line(-1,-1){4}}
\put(12,11.05){\line(-1,1){2}}
\put(10,13.05){\line(-1,-1){3}}
\put(7,10.05){\line(-1,1){2}}
\put(5,12.05){\line(-1,-1){2}}
\put(13,0.1){\line(1,1){13}}
\put(13,0.1){\line(-1,1){13}}
\put(24,11.1){\line(-1,1){1}}
\put(23,12.1){\line(-1,-1){2}}
\put(21,10.1){\line(-1,1){5}}
\put(16,15.1){\line(-1,-1){4}}
\put(12,11.1){\line(-1,1){2}}
\put(10,13.1){\line(-1,-1){3}}
\put(7,10.1){\line(-1,1){2}}
\put(5,12.1){\line(-1,-1){2}}
\put(10.5,10.75){{\tiny $[2]$}}
\put(9.5,9.75){{\tiny $[3]$}}
\put(8.5,8.75){{\tiny $[4]$}}
\put(7.5,7.75){{\tiny $[7]$}}
\put(6.5,6.75){{\tiny $[8]$}}
\put(12.6,-1){ $\tiny{{}_{0}}$}
\put(13.1,-1){ $\tiny{{}_{-1}}$}
\put(14.1,-1){ $\tiny{{}_{-2}}$}
\put(15.1,-1){ $\tiny{{}_{-3}}$}
\put(16.1,-1){ $\tiny{{}_{-4}}$}
\put(17.1,-1){ $\tiny{{}_{-5}}$}
\put(18.1,-1){ $\tiny{{}_{-6}}$}
\put(19.1,-1){ $\tiny{{}_{-7}}$}
\put(20.1,-1){ $\tiny{{}_{-8}}$}
\put(21.1,-1){ $\tiny{{}_{-9}}$}
\put(11.6,-1){ $\tiny{{}_{1}}$}
\put(10.6,-1){ $\tiny{{}_{2}}$}
\put(9.6,-1){ $\tiny{{}_{3}}$}
\put(8.6,-1){ $\tiny{{}_{4}}$}
\put(7.6,-1){ $\tiny{{}_{5}}$}
\put(6.6,-1){ $\tiny{{}_{6}}$}
\put(5.6,-1){ $\tiny{{}_{7}}$}
\put(4.6,-1){ $\tiny{{}_{8}}$}
\put(3.6,-1){ $\tiny{{}_{9}}$}
\end{picture}
$\mbox{}$
\end{center}
\caption{
The partition enclosed by the thick lines is $\lambda=(10, 10, 8, 8, 8, 6, 6, 6, 6,$ $1, 1)$.
If $k=6$ then $A(\lambda, <6)= \{ a_1, a_3 \}$, $R(\lambda, <6)= \{g_2, g_5\}$, and
$$\hspace{-0.7in} \text{ev}_\lambda(s_6)= \frac{[2]}{[3]}\frac{[3]}{[4]}\frac{[4]}{[5]} \frac{[7]}{[8]} \frac{[8]}{[9]} = \frac{[2][7]}{[5][9]}= \frac{[c(g_5)-c(b)][c(g_2)-c(b)]}{[c(a_3)-c(b)][c(a_1)-c(b)]}.
$$
The factors in the numerator of the first expression are displayed. These are
the $q$-integers corresponding to the hook lengths of the boxes in the same column
as the addable box $b$ in row 6.
\label{ar_fig}}
\end{figure}
\begin{Proposition} \label{the_vals1}
Let $\lambda$ be a partition. Let $A(\lambda,<k)$ (resp. $R(\lambda, <k)$) be the set of boxes which can be added to (resp. removed from) $\lambda$ on rows $\lambda_j$ with $j<k$ such that the result is still a partition. Let $b=(\lambda+\varepsilon_k) / \lambda$ and let
$c(\cdot)$ be as in Figure \ref{partition_bij}. Then, up to multiplication by a power of $q$,
\begin{equation} \label{canceled_eq}
\text{ev}_\lambda(s_k)=
\begin{cases}
\displaystyle{ \frac{
{\prod_{r \in R(\lambda,<k)} [c(r)-c(b)] }} { {\prod_{a \in A(\lambda,<k) } [c(a)-c(b)]}}
},
&\text{ if } \lambda+ \varepsilon_k \text{ is a partition}, \\
0, &\text{ if } \lambda+ \varepsilon_k \text{ is not a partition}.
\end{cases}
\end{equation}
\end{Proposition}
\begin{proof} For $1 \leq j \leq N$, let $g_j$ be the last box in row $j$ of $\lambda$. By Lemma \ref{get_a_hold}, up to multiplication by a power of $q$,
\begin{equation}
\text{ev}_\lambda^{}(s_k)= \text{ev}_\lambda^{} \left( \prod_{1 \leq j < k} \frac{z_jz_k^{-1}- q^{2 +2j-2k}z_j^{-1}z_k}{\sigma_{\varepsilon_j} ( z_jz_k^{-1}- q^{2 +2j-2k}z_j^{-1}z_k )} \right)^{}=
\prod_{1 \leq j < k} \frac{
[c(g_j)-c(b)]}{[c(g_j)-c(b)+1]},
\end{equation}
where the last equality is a simple calculation from definitions. The denominator on the right side is never zero, and the numerator is zero exactly when $\lambda_k=\lambda_{k-1}$, so that $\lambda+\varepsilon_k$ is no longer a partition. If $\lambda_j =\lambda_{j+1}$ for any $j<k$, then there is cancellation, giving \eqref{canceled_eq}. See Figure \ref{ar_fig}.
\end{proof}
\begin{Proposition} \label{the_vals2}
Let $N_{\bar j}^l(\mu / \lambda)$ be as in Section \ref{MM_section}.
For any partition $\lambda$,
\begin{equation}
\begin{cases}
val_{\phi_{2\ell}} \text{ev}_\lambda(s_k)= N_{\bar i}^l(\mu / \lambda),
& \hspace{-0.35cm} \text{ if } \mu =\lambda+\varepsilon_k \text{ is a partition, and } \mu / \lambda \text{ is an } {\bar i} \text{ colored box},\\
\text{ev}_\lambda(s_k)=0, & \hspace{-0.35cm} \text{ otherwise}.
\end{cases}
\end{equation}
\end{Proposition}
\begin{proof} By Proposition \ref{the_vals1}, $\text{ev}_\lambda(s_k)=0$ if $\lambda+\varepsilon_k$ is not a partition. If $\lambda+\varepsilon_k$ is a partition then
\begin{equation}
\begin{aligned}
\{ b \in A(\lambda, <k) : \bar c(b) & = \bar c(\mu /\lambda) \}
= \{ b \in A_{\bar i}(\lambda) \ |\ b \text{ is to the left of } \mu/\lambda \}, \quad \text{and}
\\
\{ b \in R(\lambda, <k) : \bar c(b) & = \bar c(\mu /\lambda) \}
= \{ b \in R_{\bar i}(\lambda) \ |\ b \text{ is to the left of } \mu/\lambda \},
\end{aligned}
\end{equation}
where the notation is as in Section \ref{MM_section}. Since
\begin{equation}
[x] = \frac{q^x-q^{-x}}{q-q^{-1}}= q^{-x} (q-q^{-1})^{-1} \prod_{d | 2x} \phi_d,
\end{equation}
$[x]$ is divisible by $\phi_{2\ell}$ if and only if $x$ is divisible by $\ell$, and $[x]$ is never divisible by $\phi_{2\ell}^2$. The result now follows from Proposition \ref{the_vals1}.
\end{proof}
\begin{proof}[Proof of Theorem \ref{main}]
Fix $\lambda$ and $1 \leq k \leq m_\lambda$. From definitions, $(\text{ev}_\lambda\otimes 1)v_{\varepsilon_{k_j} \hspace{-0.08cm}+}=v_{\mu^{(j)}}$. Thus, using \eqref{ev-tensor},
\begin{align}
r_{j}(\lambda) =(v_{\mu^{(j)}}, v_{\mu^{(j)}}) = ((\text{ev}_\lambda \otimes 1)v_{\varepsilon_{k_j} \hspace{-0.08cm}+}, (\text{ev}_\lambda \otimes 1)v_{\varepsilon_{k_j} \hspace{-0.08cm}+}) = \text{ev}_\lambda (v_{\varepsilon_{k_j} \hspace{-0.08cm}+}, v_{\varepsilon_{k_j} \hspace{-0.08cm}+}) = \text{ev}_\lambda(s_{k_j}).
\end{align}
The result now follows from Proposition \ref{the_vals2}.
\end{proof}
\def$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'${$'$}
\def$'$} \def\cprime{$'${$'$} \def$'$} \def\cprime{$'${$'$}
|
\section{Introduction} \label{sec:intro}
In the past two decades quantum computation has attracted growing interest,
encouraged by the development of tools for the manipulation of single quantum
objects as well as by several remarkable theoretical findings \cite{chuang}.
Different systems have been proposed as candidates for quantum computing; they
are based, for instance, on cavity-laser atoms, Bose-Einstein condensates, or NMR techniques (see e.g. \cite{daems,muller,holthaus}). At the same time, a
number of quantum algorithms have been designed and some have been shown to be
even exponentially faster than their best classical counterparts \cite{shor}. In
particular, quantum search algorithms, although able to achieve ``only'' a
polynomial speedup, have been proved to be very promising and of widespread use
in quantum computation \cite{williams,childs3}.
One of the best known quantum search algorithms is attributed to Grover \cite{grover}.
The algorithm can find a target within an unsorted database made up of $N$ items
using $O(\sqrt{N})$ queries. Due to its broad range of applications and its ability to be effectively used as a subroutine \cite{williams,roland2},
Grover's algorithm has been thoroughly investigated and a number of different
implementations have been proposed (see e.g.
\cite{AA,farhi,roland,childs,childs2,daems,dowling,williams,tulsi}). Recently,
implementations based on continuous-time \cite{farhi,childs} as well as on
discrete-time quantum walks \cite{AKR,SKW,tulsi} have been introduced: Given
that the application of random walks in classical algorithms provided
significant advantages for approximations and optimization, one is strongly
motivated to study quantum walks as algorithmic tools. At the current stage
the aim is not only to achieve similar computational improvements, but also to
understand the capabilities of quantum computations.
Here we focus on the
approach pioneered by Farhi and Gutmann \cite{farhi} and further developed by
Childs and Goldstone \cite{childs}, based on continuous-time quantum walks
(CTQWs). This implementation provides several important advantages: the algorithm does not
need auxiliary storage space and it makes it possible to take into account the geometrical
arrangement of the database (the latter being either a physical position space
or an efficiently encoded Hilbert space \cite{tulsi}). Indeed, while previous
studies just considered the cases of the translationally invariant $\mathbb{Z}^d$
(named $d$-dimensional periodic lattices by Childs and Goldstone \cite{childs}
and called hypercubic lattices in statistical physics, the term which we adopt here) and of complete graphs
\cite{farhi1}, here we extend the investigations to the case of generic
structures and analyze how geometrical parameters [e.g., (fractal)
dimension or (average) coordination number; see later in this article]
affect the
dynamics of the CTQW. In \cite{childs} it was shown
that quantum searches based on CTQWs recover the optimal quadratic speedup on
complete graphs and on high-dimensional hypercubic lattices (with dimension
$d>4$), while for low-dimensional ($d<4$) lattices CTQWs can not
outperform their classical counterpart. Hence, it may appear that $d=4$ works as a
``critical dimension'', separating highly performing structures from poorly
performing ones. We will show, both analytically and numerically, that the dimension of the substrate is
not sufficient for getting a sharp transition from the ground to the first excited state; we also take into account the success
probability $\pi_{w,s}(t)$, that is, the probability of finding the quantum
walker at the target site $w$ at time $t$, given as initial state the
equally weighted superposition $s$: An efficient quantum walk gives rise
to a success probability close to $1$ already at very small times $t$. In
particular, we will take into account several kinds of structures:
translationally invariant structures, such as $d$-dimensional hypercubic
lattices, complete graphs, fractals with low fractal dimension like dual
Sierpinski gaskets and T-fractals, hierarchical structures as
Cayley trees, and structures with (fractal) dimensions larger than four,
such as Cartesian products between Euclidean lattices and dual Sierpinski
gaskets (for the precise definitions of these structures and of the fractal
dimension considered here, see later in this article). In this way we are able to show
that for translationally invariant
structures with high dimensions, $\pi_{w,s}(t)$ displays sharp peaks,
while for fractals or low-dimensional structures
the peaks are low and broad so that the quantum walk is not particularly
effective in the sense that there exists only a low probability
$\pi_{w,s}(t)$ for any $t$. Moreover, for any structure, we evidence
interference phenomena which give rise to a non-monotonic time dependence
of the $\pi_{w,s}(t)$;
such effects can be significant and must be properly taken into account
when considering applications.
Interestingly, the CTQW Hamiltonian used in the quantum search also describes,
in solid-state physics, the dynamics of a tight-binding particle in the presence
of static, substitutional impurities. In this context our results show that in
regular, highly-connected geometries (such as the high-dimensional tori)
the probability of finding the moving particle at the impurity site is
(quasi-) periodic in time and that the localization can be very effective.
Our article is structured as follows. In Sec.~\ref{sec:quantum_search} we first
review basic principles concerning Grover's search and we explain how it can
be implemented by means of CTQWs. In Sec.~\ref{ssec:inho_graphs} we describe
the structures used as substrate for the CTQW. Then, in
Sec.~\ref{sec:phase_transition} we present several analytical results, later
corroborated and deepened in Sec.~\ref{sec:numerics}, where our numerical
results are shown. In Sec.~\ref{sec:success} we focus on the success
probability. Finally, Sec.~\ref{sec:conclusions} contains our conclusions and
discussion. In Appendix \ref{appe1}, Appendix \ref{appe2} and Appendix
\ref{appe3} we report the details of our analytical calculations.
\section{Quantum walks and Grover's search} \label{sec:quantum_search}
Grover's search algorithm \cite{grover} is meant to solve the unsorted search
problem under the assumption that there exists a computational oracle
working as a black-box function able to decide whether a candidate
solution is the true solution. Hence, the oracle knows which is the
target among the $N$ entries. The task is to find a target $w$ using the
fewest calls to the oracle. While the classical algorithm requires
exhaustive searches implying $O(N)$ queries, Grover's algorithm is able to
find $w$ using $O(\sqrt{N})$ queries, giving rise to a quadratic speed up
\cite{williams,bennett}.
A short outline of the idea behind Grover's algorithm in the presence of a
single marked target is as follows: First of all, one associates each of
the $N$ index integers with a unique orthonormal vector $|x \rangle = |1
\rangle, |2 \rangle, ..., |N \rangle$ in an $N$-dimensional Hilbert space.
Then, one chooses as initial state
\begin{equation}\label{eq:initial}
|s \rangle =
\frac{1}{\sqrt{N}} \sum_{x=1} ^{N} |x \rangle,
\end{equation}
which is delocalized over the entire set of states $|x \rangle$ with
equal weights at every site $x$. This is the least biased initialization
one can arrange, given the available information (since each of the $N$
nodes is in principle equally likely to be the target index, the initial
state is prepared as an equally weighted superposition of all $N$
indices).
Now, to perform the search, one needs to make the state $|s \rangle$ evolve into
a state that has almost all the amplitude concentrated in just the $| w \rangle$
component. In such a way a single final measurement will find the system in the
state $|w \rangle$, hence revealing the identity of the target index. More
precisely, we need an evolution operator whose repeated application makes the
amplitude of $| w \rangle$ grow with the number of iterations \cite{williams}.
This task was originally accomplished within the standard paradigm for quantum
computation \cite{grover}, namely using a discrete sequence of unitary logic
gates, while in the last years several different implementations have been
introduced in which the state of the quantum register evolves continuously under
the influence of a driving Hamiltonian \cite{farhi1,farhi2,roland}. In
particular, here we follow the approach developed by Childs and Goldstone
\cite{childs} which relies on CTQW \cite{farhi}. As
already mentioned, due to their versatility, CTQWs allow to model any discrete
database. In fact, a generic discretizable database can be represented by a
graph $\mathcal{G}=\{V,E\}$ made up of a set of nodes $V=\{1,2,...,N\}$, each
corresponding to a different item, and of a set of links $E$ joining nodes
pairwise in such a way that the topology of the graph mirrors the arrangement of
the database. The graph $\mathcal{G}$ can be algebraically described by the
adjacency matrix $\mathbf{A}$ whose entry $A_{ij}$ equals one if nodes $i$ and $j$ are
connected, otherwise it is zero (also for the diagonal elements). From $\mathbf{A}$ one can directly calculate
the degree matrix $\mathbf{Z}$, which is a diagonal matrix with elements
$Z_{ij}=z_i \delta_{ij}$, where $z_i = \sum_{j \in N} A_{ij}$ is the
coordination number (or degree) of the $i$-th node,
that is, the number of its nearest neighbors.
CTQWs on a graph are defined by the Laplacian
matrix $\mathbf{L}=\mathbf{Z}-\mathbf{A}$
and obey the following Schr\"{o}dinger equation for the transition
amplitude $\alpha_{k,j}(t)$ from state $| j \rangle$ to state $| k
\rangle$ \cite{kempe}:
\begin{equation}\label{eq:schrodinger}
\frac{d}{dt} \alpha_{k,j}(t)=-i \sum_{l=1}^{N} H_{kl} \alpha_{l,j}(t),
\end{equation}
where the Hamiltonian is just given by $\mathbf{H}=\mathbf{L}$, therefore, the time is dimensionless and given in units of the coupling elements $H_{kl}$.
The formal solution for $\alpha_{k,j}(t)$ can be written as
\begin{equation}\label{eq:formal_solution}
\alpha_{k,j}(t)=\langle k |
\exp(-i\mathbf{H}t) | j \rangle,
\end{equation}
and its magnitude squared provides the quantum mechanical transition
probability $\pi_{k,j}(t) \equiv |\alpha_{k,j}(t)|^2$. Notice that
$\mathbf{L}$ is symmetric and non-negative definite; its ground state,
corresponding to eigenvalue $0$, is given by $| s \rangle$.
Due to the unitary time-evolution generated by $\mathbf{H}$, CTQWs are
symmetric under time-inversion, which precludes $\pi_{k,j}(t)$ from
attaining equipartition at long times. This is different from the behavior
of classical continuous-time random walks. Moreover, CTQWs keep memory of
the initial conditions, as exemplified by the occurrence of (quasi-)
revivals \cite{revival,revival2}.
Now, the ``oracle Hamiltonian'' is given by
\begin{equation}\label{eq:oracle}
\mathbf{H}_w=-|w \rangle \langle w|,
\end{equation}
whose ground state, with energy $-1$, is just the marked item $|w
\rangle$; all other states have energy zero \cite{farhi1}. Then, the
Hamiltonian $\mathbf{H}$ governing the time evolution of the quantum walk
is
\begin{equation}\label{eq:hamiltonian}
\mathbf{H} = \gamma \mathbf{L} + \mathbf{H}_w,
\end{equation}
where $\gamma$ is a tunable parameter with units of inverse time, hence dimensionless here.
The \textit{success probability} $\pi_{w,s}^{\gamma}(t)$ is defined as
(see also \cite{childs})
\begin{equation}\label{eq:success_prob}
\pi_{w,s}^{\gamma}(t)
\equiv |\langle w | \exp(-i \mathbf{H} t) | s \rangle|^2,
\end{equation}
i.e., as the transtion probability to be at time $t$ at the target $w$
when starting in the delocalized state $|s\rangle$.
Here we study its dependence on time $t$ and on the parameter $\gamma$ and
we especially aim to evidence the existence of an optimization parameter
$\gamma_\mathrm{max}$, possibly depending on time, which maximizes
$\pi_{w,s}^{\gamma}(t)$. Notice that, having fixed $\gamma$, the time $t$
can be measured in terms of the number of queries of the discrete Grover
oracle \cite{childs2}.
\section{Fractal and hierarchical structures}\label{ssec:inho_graphs}
Before proceeding, it is worth introducing the geometrical structures on
which we are focusing, first the dual Sierpinski gasket (DSG), the
T-fractal (TF), and the Cayley tree (CT); examples of these structures are
shown in Fig.~\ref{fig:frattali}a, Fig.~\ref{fig:frattali}b, and
Fig.~\ref{fig:frattali}c, respectively. Notice that these structures
differ significantly from those analyzed previously: While hypercubic
lattices with periodic boundary conditions (toroids)
are translationally invariant, the aforementioned structures are not.
The DSG, TF and CT can be built iteratively; at the $g$-th iteration we
have the fractal of generation $g$ (see e.g. \cite{ABM2008,redner,cosenza}). The
DSG and the TF are examples of exactly decimable fractals for which the fractal
dimension, $d_f$,
and the spectral dimension, $\tilde{d}$, are exactly known.
While the fractal dimension relates the number of nodes inside a
sphere to the radius of the sphere \cite{fractals}, the
spectral dimension is obtained from the scaling of the eigenmodes of a given
structure (phonons for lattices, fractons for fractal substrates)
\cite{sdim}, most simply seen in the probability of a random
walker to be (still or again) at the original site \cite{aodim}.
Here, we have namely $d_f = \ln 3 / \ln 2 \approx 1.585$ and $\tilde{d}= 2
\ln 3 / \ln 5 \approx 1.365$ for the DSG, and $d_f = \ln 3 / \ln 2$ and
$\tilde{d}= 2 \ln 3 / \ln 6 \approx 1.226$ for the TF. We recall that for
the usual, translationally invariant lattices the spectral and fractal
dimensions coincide with the Euclidean
dimension $d$, namely $\tilde{d}=d_f=d=1$ for the chain, $\tilde{d}=d_f=d=2$ for the square
lattice and so on. On the other hand, on fractal
structures $\tilde{d}$ often replaces $d$ when dealing with dynamical and
thermodynamical properties \cite{hattori}. Also, for fractals
$\tilde{d}<d_f$ and $d_f$ is smaller
than the Euclidean dimension of the embedding space, that is, 2 for DSG and
TF. The CT is no fractal in the classical sense, since its growth with
increasing generation is exponential. However, the CT is built in a
hierarchical manner, analogous to the TF. We also notice that both, the TF and
the CT, are trees (and are hence devoid of loops) and exhibit a large
number $O(N/2)$ of end nodes.
\begin{figure} \includegraphics[width=55mm,height=48mm]{DSG4.pdf}
\includegraphics[width=55mm,height=48mm]{T4.pdf}
\includegraphics[width=55mm,height=48mm]{CT5.pdf}
\caption{\label{fig:frattali}Examples of fractal structures considered here; in
general the star indicates the position of the trap. Panel $a$: Dual Sierpinski
gasket of generation $4$ and volume $N=3^g=81$; due to symmetry, the
three corners are equivalent. Panel $b$: T-fractal of generation $3$ and volume
$N=3^g+1=28$; due to the symmetry all outmost sites (at distance $2^{g-1}$ from
the central node) are equivalent. Panel $c$: Cayley tree of generation $5$ and
volume $N=3 \times 2^{g}-2=94$; all of the $N/2+1$ outmost sites are
equivalent.} \end{figure}
As we will show in the following, CTQW on
the DSG, the TF and the CT can display a low
probability $\pi_{w,s}^\gamma(t)$ for any $t$ and $\gamma$ when compared
with the case of the translationally invariant lattices, especially when the
dimension of the lattice becomes larger than $4$.
Therefore one can ask whether the probabilities $\pi_{w,s}^{\gamma}(t)$ can be
improved by adopting (fractal or hierarchical) substrates which display a
large spectral dimension and a large coordination number.
For instance, we can build up such structures by combining the DSG and Euclidean lattices by means of Cartesian products. In general,
the Cartesian product of two graphs $\mathcal{G}_1=\{V_1,E_1\}$ and
$\mathcal{G}_2=\{V_2,E_2\}$ is a graph $\mathcal{G} = \mathcal{G}_1 \times
\mathcal{G}_2$ with the vertex set $V_1 \times V_2$, and such that two nodes
$(x_1 x_2, y_1 y_2)$ are adjacent if $(x_1,y_1) \in E_1$ and $x_2 = y_2$, or
$x_1=y_1$ and $(x_2, y_2) \in E_2$. It has been shown \cite{woess} that the
spectral dimension $\tilde{d}$ on the product graph $\mathcal{G}$ is then the sum of
the corresponding dimensions of the two initial graphs $\tilde{d}_1$ and
$\tilde{d}_2$.
We combine in this way the DSG with the chain $L_1$, with the square lattice $L_2$ and
with the cubic lattice $L_3$, to obtain more complex structures displaying spectral
dimensions approximately equal to $2.365$, $3.365$ and $4.365$, respectively.
Finally, it should be underlined that, dealing with such structures, the
location of the target, i.e. the node $w$, also (quantitatively) matters. In the
numerical analysis of Sec.~\ref{sec:numerics} we place the target on a
peripheral site, which means, without loss of generality, the apex for the DSG,
the leftmost site for the TF and an outmost site for the CT (see
Fig.\ref{fig:frattali}). We expect that a peripheral position for the target
site does not correspond to an optimal situation and, since a priori the target
position is unknown, this choice prevents an overestimation of the
probability of finding the target.
\section{Overlaps and transition}\label{sec:overlaps_transitions}
Let us consider the Hamiltonian of Eq.~(\ref{eq:hamiltonian}) and denote
the corresponding set of eigenstates and eigenvalues by $\{|\psi_k\rangle
\}$ and $\{ E_k \}$, respectively. Now, for large $\gamma$ the
contribution of $\mathbf{H}_w$ to $\mathbf{H}$ is negligible and the
ground state $|\psi_0\rangle$ is close to $|s\rangle$. On the other hand,
as $\gamma \rightarrow 0$ the ground state is close to $|w\rangle$ since
the weight of $\mathbf{L}$ is small and, from perturbation theory, we
expect that $| s \rangle$ is close to $| \psi_1 \rangle$, i.e. to the
first excited state of $\mathbf{H}$ \cite{childs}.
As pointed out by Childs and Goldstone \cite{childs}, there exists an
intermediate range of $\gamma$ where, for complete graphs and hypercubic
lattices with dimension $d>4$, $|s \rangle$ switches from the ground state $|
\psi_0 \rangle$ to the first excited state $| \psi_1 \rangle$; in the very same
region of $\gamma$ the target state $| w \rangle$ switches from a state with
large overlap with $| \psi_1 \rangle$ to the ground state $| \psi_0 \rangle$.
Therefore, by varying $\gamma$, we can find a particular value for $\gamma$ for which the Hamiltonian $\mathbf{H}$ can evolve the state $|s \rangle$ into a state close to the target state $|w \rangle$.
The existence and the narrowness of such a range for $\gamma$ are crucial for a
large success probability.
\subsection{Analytical results}\label{sec:phase_transition}
As anticipated, the CTQW under study can yield effective results if the
Hamiltonian $\mathbf{H}$ is able to rotate the state $| s \rangle$ to a
state with a large overlap with $| w \rangle$. For this to occur a first
condition which needs to be fulfilled is that there exists an intermediate range of $\gamma$ over which the state $| s \rangle$ has a substantial overlap with both $| \psi_0 \rangle$ and $| \psi_1 \rangle$. In this subsection the
occurrence of such a condition is investigated analytically for an
arbitrary structure, following arguments similar to those exploited in
\cite{childs} for $d$-dimensional hypercubic lattices; the details of the
calculations are given in Appendixes A and B
while here we just report the main results.
First of all, we define $| \phi_k \rangle$ and $\mathcal{E}(k)$, the
$k$-th eigenstate and eigenvalue of the Laplacian $\mathbf{L}$,
respectively. In the basis of the eigenstates $| \phi_k \rangle$ the
target site $|w \rangle$ can be written as
\begin{equation} \label{eq:expansion}
|w \rangle = \sum_{k=0}^{N-1}
a_k |\phi_k \rangle,
\end{equation}
where $a_k \equiv \langle w| \phi_k \rangle$.
As derived in Appendix~\ref{appe1}, once $\xi_j \equiv \sum_{k \neq 0}
|a_k|^2 /[\mathcal{E}(k)]^j$ is set, the overlap of $| s \rangle$ with the
ground state $| \psi_0 \rangle$ or with the first non-degenerate excited
state $| \psi_1 \rangle$ is (depending on $\gamma$) limited from below by
the same bound. We find namely
\begin{equation} \label{eq:s_0_short}
1> |\langle s | \psi_0 \rangle
|^2 > 1 - \frac{\xi_2}{N (\gamma - \xi_1 )^2},
\end{equation}
for $\gamma > \xi_1$, and also
\begin{equation} \label{eq:s_l_short}
1> |\langle s | \psi_1
\rangle |^2 > 1 - \frac{\xi_2}{N (\xi_1 - \gamma)^2},
\end{equation}
for $\gamma < \xi_1$.
Now, we define the (size dependent) critical value
$\tilde\gamma$ as that value of $\gamma$ for which
\begin{equation}\label{eq:intersection}
|\langle s | \psi_0\rangle |^2 = |\langle s|\psi_1\rangle|^2
\end{equation}
is satisfied, i.e., the $\gamma$-value for which the projection of state $s$ onto the ground and the first excited state has the same magnitude. This is particularly interesting in the limit $N \to \infty$, when a level crossing from state $| \psi_0 \rangle$ to state $| \psi_1 \rangle$ can occur. In fact, we find that if $\xi_2 /[N(\xi_1 - \gamma)^2]$ converges (at least in the limit $N
\rightarrow \infty$) to $0$ as $\gamma$ approaches $\xi_1$ (from different
sides), then $|\langle s | \psi_0 \rangle |$ and $|\langle s | \psi_1
\rangle |$ both approach $1$ (see Eqs.~(\ref{eq:s_0_short}) and
(\ref{eq:s_l_short})), namely a transition occurs at $\gamma = \xi_1$. Notice, however, that the
condition for this to occur is in general non-trivial as $\xi_1$ and
$\xi_2$ both depend on $N$. In Appendix~\ref{appe1} we find a sufficient
condition in the Laplacian spectrum and, in particular, in
Appendix~\ref{appe2} we prove that such a condition holds for the DSG for
which $\mathcal{E}(k)$ is exactly known \cite{cosenza,jurjiu,ABM2008}; in
this case we find that
\begin{equation}\label{eq:xi1_asy_short}
\xi_1 \leq C N^{\alpha + 2/\tilde{d}},
\end{equation}
and
\begin{equation}\label{eq:xi2_asy_short}
\xi_2 \leq C
N^{\alpha + 4/\tilde{d}},
\end{equation}
where $\alpha$ is a parameter
depending on the given network. In general, $-1 \leq \alpha < 0$; for
hypercubic lattices $\alpha=-1$, regardless of their dimension, while for
the DSG we can numerically estimate $\alpha$ as being $\alpha \approx
-0.9$ (see Appendix~\ref{appe1}). Therefore, we expect (for DSG)
$\tilde\gamma$ to be, approximately:
\begin{equation}\label{eq:gammacrit_short}
\tilde{\gamma} \approx C
N^{\alpha+2/\tilde{d}}.
\end{equation}
This result is consistent up to logarithmic corrections with the critical
points found in \cite{childs} for the linear chain and the square lattice,
namely $\tilde{\gamma} \sim N$ and $\tilde{\gamma} \sim \log N$,
respectively.
Finally, we point out that the critical parameter $\tilde{\gamma}$
provides interesting information in the context of quantum adiabatic computation
\cite{roland,santoro}: $\tilde{\gamma}$ represents a threshold below which we
can expect $| w \rangle$ to have a large overlap with the ground state.
\subsection{Numerical results}\label{sec:numerics}
We now consider the three examples of low-dimensional inhomogeneous
structures described previously, for which the overlaps of the initial state
$|s \rangle$ and of the target state $|w \rangle$ with $|\psi_1 \rangle$
and $|\psi_0 \rangle$ are shown in Figs.~\ref{fig:DS_ext}-\ref{fig:CT_ext}
for different generations $g$, as a function of the parameter $\gamma$.
These plots evidence that there exists an intermediate range of $\gamma$
where the state $|s \rangle$ changes from having a large overlap with the
first excited state to having a large overlap with the ground state. In
the same region of $\gamma$ the overlap $|\langle w | \psi_1 \rangle|^2$
is significant for structures of small size (top panels in
Figs.~\ref{fig:DS_ext}-\ref{fig:CT_ext}), while it is still very small
when the size is large (bottom panels in
Figs.~\ref{fig:DS_ext}-\ref{fig:CT_ext}). This is vastly distinct from the
situation found for hypercubic lattices \cite{childs} where close to
$\tilde{\gamma}$ the overlap $|\langle w | \psi_1 \rangle|^2$ is
significant.
\begin{figure} \includegraphics[width=80mm,height=70mm]{DS_G3_ext.pdf}\\
\includegraphics[width=80mm,height=70mm]{DS_G6_ext.pdf}
\caption{\label{fig:DS_ext}(Color online) Overlaps for a DSG of generation $g=3$
(up) and $g=6$ (bottom) as a function of (the dimensionless) $\gamma$, see text for details.} \end{figure}
\begin{figure} \includegraphics[width=80mm,height=70mm]{T_G3_ext.pdf}\\
\includegraphics[width=80mm,height=70mm]{T_G6_ext.pdf}
\caption{\label{fig:T_ext}(Color online) Overlaps for a T-fractal of generation
$g=3$ (up) and $g=6$ (bottom) as a function of $\gamma$. The symbols are as in
Fig.~\ref{fig:DS_ext}.} \end{figure}
\begin{figure} \includegraphics[width=80mm,height=70mm]{CT_G3_inset.pdf}\\
\includegraphics[width=80mm,height=70mm]{CT_G10_inset.pdf}
\caption{\label{fig:CT_ext}(Color online) Overlaps for a CT of generation $g=3$
(up) and $g=10$ (bottom) as a function of $\gamma$. The symbols are as in
Fig.~\ref{fig:DS_ext}.} \end{figure}
By comparing the plots obtained for the DSG (Fig.~\ref{fig:DS_ext}),
the TF (Fig.~\ref{fig:T_ext}), and the CT (Fig.~\ref{fig:CT_ext}), we
notice that $\tilde{\gamma}$ depends sensitively on the underlying
topology. In fact, going from DSG to CT and then to TF we notice an
amplification of the transition region, which is for TF most spread out,
requiring relatively large values of $\gamma$ in order for $| s \rangle$
to have a large overlap with the ground state. Such effects can be
ascribed to the absence of loops and to the existence for TF of a large
number of peripheral sites scattered throughout the whole structure which
give rise to localization effects \cite{ABM2008}.
We now calculate $\tilde{\gamma}$ according to Eq.~(\ref{eq:intersection})
and for several values of $g$; numerical data and relative best fits are
depicted in Fig.~\ref{fig:gamma_crit}. For the DSG, numerical points are
best fitted by the function $\tilde{\gamma} \approx 3^{0.55 g} \approx
N^{0.55}$, in very good agreement with the analytical findings. In fact,
according to our analytical investigations, $\tilde{\gamma}$ scales with
the size of the database like $N^{\alpha + 2/\tilde{d}}$ (see
Eq.~(\ref{eq:gammacrit_short})), where, for the DSG, $\alpha + 2/\tilde{d}
= \alpha + \log 5 / \log 3 \approx 0.57$ ($\alpha$ is taken to equal
$-0.9$). Let us now consider the case of the TF: From numerical data the
best fit turns out to be $\tilde{\gamma} \approx 3^{0.70 g} \approx
N^{0.70}$. Interestingly, this result is still in very good agreement
with the analytical approximation of Eq.~(\ref{eq:gammacrit_short}). In fact,
for the TF the exponent gets $\alpha + 2/\tilde{d} = \alpha + \log 6 /
\log 3 \approx 0.73$ (where, again, $\alpha$ is taken to equal $-0.9$,
consistently with the estimates given in Appendix \ref{appe1}). Such a
consistency might suggest that Eq.~(\ref{eq:gammacrit_short}) is valid not only
for the DSG but for any (exactly decimable) fractal with $\tilde{d}<2$.
As for the CT, not being a fractal, Eq.~(\ref{eq:gammacrit_short}) does
not hold. Indeed, we find that the value of $\tilde{\gamma}$ corresponding
to Eq.~(\ref{eq:intersection}) grows linearly with the generation of the
fractal, namely logarithmically with $N$. In Fig.~\ref{fig:gamma_crit}
numerical data are fitted by the curve $\tilde{\gamma} = g -1$ (notice the
semilogarithmic plot).
\begin{figure} \includegraphics[width=80mm,height=70mm]{g_crit.pdf}
\caption{\label{fig:gamma_crit}(Color online) Estimate of $\tilde{\gamma}$
for DSG ($\bullet$), TF ($*$) and CT (o) in a semi-logarithmic scale. The
continuous lines represent the best fits.} \end{figure}
Apart from this, the plots shown in
Figs.~\ref{fig:DS_ext}-\ref{fig:CT_ext} look rather similar. In
particular, for networks of large sizes $|\langle w| \psi_0 \rangle|^2$
decays with $\gamma$ more rapidly than $|\langle s| \psi_1 \rangle|^2$.
Hence, the range of $\gamma$ over which the transition occurs is wide,
analogously to what happens on low-dimensional hypercubic lattices (see
\cite{childs}). As shown in the next section, this has important effects
on the behaviour of the success probability and suggests that a sharp transition is associated with an effective search algorithm.
In order to sketch the role of the position of a target we show in the inset of
Fig.~\ref{fig:CT_ext} for the CT of $g=10$ the case of a target placed on a nearest neighbor of the central node. We see that the transition region is shifted
towards lower values of $\gamma$. This means that a more central placement of
the target is improving the probability for the CTQW to reach the target. Analogous results were also found for DSG and for TF. Thus, limiting our focus to peripheral nodes will prevent us from overestimating the success probabilities.
We now consider fractal structures exhibiting large spectral dimension; in
particular, we focus on fractals obtained from Cartesian products, such as
DSG $\times$ $L_1$, DSG $\times$ $L_2$ and DSG $\times$ $L_3$, as
introduced in the previous section. Again, we place the target on a
``peripheral site'', i.e. on a minimally connected site; this displays the
coordination number $5$, $7$ and $9$, for DSG $\times$ $L_1$, DSG $\times$
$L_2$ and DSG $\times$ $L_3$, respectively. We calculate for these
structures the overlaps of the initial state $|s \rangle$ and of the
target state $| w \rangle$ with $|\psi_0 \rangle$ and $|\psi_1 \rangle$;
results for DSG $\times$ $L_2$ ($\tilde{d} = 3.365$) and for DSG $\times$
$L_3$ ($\tilde{d} = 4.365$) are shown in Fig.~$6$
\begin{figure} \label{fig:cartesian}
\includegraphics[height=70mm]{CP2_G4L8_finer.pdf}
\includegraphics[height=70mm]{CP3_G4L4_finer.pdf} \caption{
(Color online) Upper panel: Overlaps for a DSG $\times$ $L_2$ structure given by
the Cartesian product of a DSG of generation $g=4$ and a square lattice of size
$L=8$; Bottom panel: Overlaps for a DSG $\times$ $L_3$ obtained from a DSG of
generation $g=4$ and a cubic lattice of size $L=4$. The lines are as in
Fig.~\ref{fig:DS_ext}.} \end{figure}
As stressed at the beginning of this section, these plots provide some
information about the sharpness of the transition from
state $| s \rangle$ to the state $|\psi_0 \rangle$ and from state $| w
\rangle$ to the state $|\psi_0 \rangle$. However,
around $\tilde{\gamma}$ also a significantly large overlap $|\langle w |
\psi_1 \rangle|^2$ is required.
Here, the transitions are still rather smooth although, by increasing
$\tilde{d}$, the region of $\gamma$ values, over which the curves
representing $|\langle w | \psi_0 \rangle|^2$, $|\langle s | \psi_0
\rangle|^2$ and $|\langle s | \psi_1 \rangle|^2$ intersect, is shrinking.
On the other hand, the overlaps between $| w \rangle$ and the first
excited state $| \psi_1 \rangle$ are negligible for all values of
$\gamma$.
\section{Success Probability} \label{sec:success}
We now turn to the success probability $\pi_{w,s}^{\gamma}(t)$,
Eq.~(\ref{eq:success_prob}), and we investigate numerically its dependence
on $t$ and on $\gamma$. Because of its dependence on time,
$\pi_{w,s}^{\gamma}(t)$ carries more information than the previously
discussed overlaps.
We first analyze the case of complete graphs and of $d$-dimensional hypercubic
lattices (for which the time dependences of $\pi_{w,s}^{\gamma}(t)$ have already
been determined for special choices of $\gamma$ in Ref. \cite{childs}) before
turning to the DSG, the TF and the CT.
\begin{figure} \includegraphics[width=85mm,height=70mm]{ampl_C_124.pdf}
\includegraphics[width=85mm,height=70mm]{ampl_C_3125.pdf} \caption{
\label{fig:Complete}(Color online) Contour plot of the success probability
$\pi_{w,s}(t,\gamma)$ as a function of (the dimensionless) time $t$ and of $\gamma$ for the complete
graph of size $N=124$ (top) and $N=3125$ (bottom). One can notice that for
$\gamma N =1$, namely $\gamma=8.1 \cdot 10^{-3}$ (top) and $\gamma=3.2 \cdot
10^{-4}$ (bottom), $\pi_{w,s}^{\gamma}(t)$ has a period $\tau=\pi
\sqrt{124}\approx 35$ and $\tau \approx 176$, respectively.} \end{figure}
We start our analysis from the complete graph $K_N$ for which, as shown in
\cite{childs}, at $\tilde{\gamma}=1/N$ the ground state changes sharply
from $|s\rangle$ to $|w\rangle$. This transition takes place at $t = \pi
\sqrt{N}/2$. Due to the special topology of $K_N$, we are able to
calculate $\pi_{w,s}^{\gamma}(t)$ exactly, obtaining
\begin{equation}\label{eq:success_completo}
\pi_{w,s}^{\gamma}(t) = \frac{1}{N} \left[ 1 + \frac{4 (N-1) \sin^2{(t(\sqrt{4
\gamma + (N \gamma -1)^2}/2)}}{4+ \gamma( N -1/\gamma)^2 } \right],
\end{equation}
see Appendix \ref{appe3} for details.
In Fig.~\ref{fig:Complete} we show $\pi_{w,s}(t,\gamma)$ for the complete
graph with $N=124$ and $N=3125$. We evaluated the figures both
numerically, by first diagonalizing $\mathbf{H}$ in
Eq.~(\ref{eq:success_prob}) and projecting on the states $| w \rangle$ and
$| s \rangle$ and also by making use of Eq.~(\ref{eq:success_completo}).
The results are numerically indistinguishable. From
Fig.~\ref{fig:Complete} we see that, around the values $\gamma = 8 \times
10^{-3}$ and $\gamma = 3.2 \times 10^{-4}$ for $N=124$ and $N=3125$
respectively, $\pi_{w,s}^{\gamma}(t)$ reaches values very close to $1$. In
fact, analyzing Eq.~(\ref{eq:success_completo}) (see Appendix
\ref{appe3}), one finds that $\pi_{w,s}^{\gamma}(t)$ attains its maximal
value of $1$ for $\gamma N =1$ and for $t=\pi \sqrt{N} (k+1/2)$, where $k
\in \mathbb{Z}$. Therefore, $\gamma_{\mathrm{max}} = 1/N$. Furthermore,
due to the fact that the period between maxima is $T= \pi \sqrt{N}$ for
$\gamma_{\mathrm{max}}$, it follows that the CTQW takes $O(\sqrt{N})$
queries to find the target, in agreement with previous results
\cite{childs,farhi1}. On the other hand, the exact dependence on $t$ and
on $\gamma$ also allows to highlight the oscillating behaviour of
$\pi_{w,s}^{\gamma}(t)$. This means that, although we properly select
$\gamma_{\mathrm{max}}=1/N$, the result for the walk depends sensitively
also on $t$. In particular, for $t=k \pi \sqrt{N}, k \in \mathbb{Z}$, the
success probability is minimal and equals $1/N$ (which also corresponds to
the absolute minimum).
For the hypercubic lattices the overlaps of the states $|s\rangle$ and
$|w\rangle$ with $|\psi_0 \rangle$ and $|\psi_1 \rangle$ have already been
analyzed in \cite{childs}; here we display analogous plots (but for larger
sizes) in order to compare them with the corresponding success probability
$\pi_{w,s}^{\gamma}(t)$.
\begin{figure} \includegraphics[width=80mm,height=70mm]{Torus5_L5.pdf}
\includegraphics[width=85mm,height=70mm]{ampl_To5_contour.pdf} \caption{
\label{fig:Torus5}(Color online) $5$-dimensional torus with linear size $L=5$;
Top: Overlaps (symbols are as in Fig.~\ref{fig:DS_ext}); Bottom: Contour plot of
the success probability $\pi_{w,s}(t,\gamma)$ as a function of $t$ and of
$\gamma$; the dashed white line represents $\tilde{\gamma} \approx 0.12$.}
\end{figure}
In Fig.~\ref{fig:Torus5} we consider the case of a $5$-dimensional torus (i.e. a five-dimensional cubic lattice with periodic boundary conditions)
of linear size $L \equiv N^{1/d}=5$: The transition at $\tilde{\gamma}
\approx 0.12$ is very clear (see the top panel) and the success
probability is sharply peaked just at $\gamma_\mathrm{max} \approx
\tilde{\gamma}$ (see the bottom panel).
Away of the critical point $\tilde{\gamma}$ the success probability
quickly decays as a function of $\gamma$: it is just in the region of
largest overlap between the initial state $| s \rangle$ and the target
state $|w\rangle$ (namely around $\tilde{\gamma}$) that one expects an
optimal success probability.
Again, we notice that $\pi_{w,s}^{\gamma}(t)$ oscillates in time; for
$L=5$ and $\gamma = \gamma_{\mathrm{max}}$ the success probability ranges
from about $0$ to about $0.8$. Moreover, for a given time $t$,
$\pi_{w,s}^{\gamma}(t)$ decays very fast as $|\gamma-\gamma_\mathrm{max}|$
increases. For instance, $\pi_{w,s}^{0.9
\gamma_\mathrm{max}}(70)/\pi_{w,s}^{\gamma_\mathrm{max}}(70)\approx0.05$.
As a result, the computational procedure for this structure can be very
efficient, provided that the parameter $\gamma$ can be sensitively
controlled.
\begin{figure} \includegraphics[width=80mm,height=70mm]{Torus2_L56.pdf}
\includegraphics[width=85mm,height=70mm]{ampl_To_contour.pdf}
\caption{\label{fig:Torus2}(Color online) $2$-dimensional square torus of linear
size $L=56$. Top: Overlaps (symbols are as in Fig.~\ref{fig:DS_ext}); Bottom:
Success probability as a function of time and $\gamma$; the dashed white line
represents $\tilde{\gamma}\approx 0.67$.} \end{figure}
For hypercubic lattices of dimension $d=4$, $d=3$ and $d=2$ (only the
latter case is depicted in Fig.~\ref{fig:Torus2}) the peaks get more and
more broadened and are of smaller magnitude. Notice that the low peaks
obtained are in agreement with the analytical results found in
\cite{childs} which predict for the $2$-dimensional torus a vanishing
success probability for $N \rightarrow \infty$.
We now turn to the analysis of the success probability for the DSG, the TF and
the CT, represented in Figures \ref{fig:ampl_DS}, \ref{fig:ampl_T} and
$12$, respectively; the upper panels display the situation for
small, the lower panels for larger networks. First of all, we notice that, as
previously found for regular lattices \cite{childs}, also for DGS, TF and CT,
$\pi_{w,s}^{\gamma}(t)$ exhibits peaks which are lower and lower as the size $N$
is enlarged. Moreover, for small sizes (upper panels) $\gamma_{\mathrm{max}}
\approx \tilde{\gamma}$, namely the maxima for the success probability occur for
values of $\gamma$ which are approximately equal to $\tilde{\gamma}$. On the
other hand, for large sizes (lower panels), in the temporal range considered
here, $\tilde{\gamma}$ provides an upper bound for $\gamma_{\mathrm{max}}$.
However, the most striking feature which emerges from the comparison between the
contour plot of the success probability for translationally invariant structures (see
Figs.~\ref{fig:Torus5} and \ref{fig:Torus2}) and for fractal/hierarchical structures (see
Figs.~ \ref{fig:ampl_DS}-$12$) is that for the latter
$\pi_{w,s}(t)$ is much more rough and broadened.
\begin{figure} \includegraphics[width=85mm,height=70mm]{ampl_DS_G3_contour.pdf}
\includegraphics[width=85mm,height=70mm]{ampl_DS_G6_contour.pdf} \caption{
\label{fig:ampl_DS}(Color online) Contour plot of $\pi_{w,s}^{\gamma}(t)$ for
the DSG of generation $g=3$ (top) and $g=6$ (bottom); the ``critical'' values
$\tilde{\gamma}$ are represented by the dashed line: $\tilde{\gamma}(3)\approx
1.20$ and $\tilde{\gamma}(6)\approx 7.38$ (see Fig.~\ref{fig:DS_ext}).}
\end{figure}
\begin{figure} \includegraphics[width=85mm,height=70mm]{ampl_T_G3_contour.pdf}
\includegraphics[width=85mm,height=70mm]{ampl_T_G6_contour.pdf}
\caption{\label{fig:ampl_T}(Color online) Contour plot of
$\pi_{w,s}^{\gamma}(t)$ for the TF of generation $g=3$ (top) and $g=6$ (bottom);
notice that $\tilde{\gamma}(3)\approx2.59$ and $\tilde{\gamma}(6)\approx21.31$
(see Fig.~\ref{fig:T_ext}).} \end{figure}
\begin{figure} \label{fig:ampl_CT}
\includegraphics[width=85mm,height=70mm]{ampl_CT_3_contour.pdf}
\includegraphics[width=85mm,height=70mm]{ampl_CT_10_contour.pdf}
\caption{(Color online) Contour plot of
$\pi_{w,s}^{\gamma}(t)$ for the CT of generation $g=3$ (top) and $g=10$
(bottom); notice that $\tilde{\gamma}(3)\approx 2.06$ and
$\tilde{\gamma}(10)\approx 8.87$ (see Fig.~\ref{fig:CT_ext}).} \end{figure}
Now, it is worth comparing the $5$-dimensional torus of
Fig.~\ref{fig:Torus5}, the square torus of Fig.~\ref{fig:Torus2} and the
CT of Fig.~$12$ since all three are of comparable size $N$.
From the computational point of view the $5$-dimensional torus corresponds
to the best situation: $\pi_{w,s}^{\gamma}(t)$ is sharp and reaches its
maximum value around $0.8$ after approximately $100$ unit steps; the CT
displays a success probability around $0.12$ after $150$ time steps; the
$2$-dimensional torus corresponds to an ineffective candidate situation:
after $300$ time steps the peak is still less than $0.1$.
In order to compare also to structures with spectral dimensions larger than
four, we show in Fig.~$13$ the success probabilities
$\pi_{w,s}^{\gamma}(t)$ for the cartesian products of a DSG of generation $g=3$
and a square lattice of size $L=8$ as well as a cubic lattice of size
$L=8$. Although the maxima of the success probabilities are in both cases
larger than the ones for the structures with low dimensions, they show
still a fairly unregular pattern. This is in contrast to the highly
regular structure of the 5-dimensional torus, see Fig.~\ref{fig:Torus5}.
\begin{figure} \label{fig:ampl_CP_DSGL}
\includegraphics[width=85mm,height=70mm]{contour_CP2_G3L8.pdf}
\includegraphics[width=85mm,height=70mm]{contour_CP3_G3L4.pdf}
\caption{(Color online) Contour plot of
$\pi_{w,s}^{\gamma}(t)$ for the cartesian product of a DSG of generation $g=3$
and a square lattice of size $L=8$ (top) and a cubic lattice of size $L=8$
(bottom).} \end{figure}
Finally, we stress that, as a result of interference phenomena,
$\pi_{w,s}^{\gamma}(t)$ oscillates with time. This has some important
consequences: Although we can determine and set the optimal
$\gamma_{\mathrm{max}}$, the probability of the CTQW reaching the target
depends on the instant of time at which it is calculated. In particular,
oscillations are ``faster'' for systems of smaller size; for example for
the complete graph we find a period $\tau(\gamma) = 2 \pi / \sqrt{(\gamma
N -1)^2 + 4 \gamma}$ (see Eq.~(\ref{eq:success_completo})), namely
$\tau(\gamma_{\mathrm{max}})= \pi \sqrt{N}$, while for the $5$-dimensional
torus the numerical analysis makes it possible to estimate a period $\tau$ which
grows exponentially with the lattice size $L$; for $L=5$ we get $\tau
\approx 200$ (see Fig.~\ref{fig:Torus5}).
\section{Conclusions and perspectives}\label{sec:conclusions}
In this work we considered CTQWs mimicking Grover's
quantum search problem; we especially focused on how the topology of the
space over which the walk takes place affects the position and sharpness of the transition of the ground state.
Previous studies \cite{childs} highlighted that for translationally invariant
structures, such as the hypercubic lattices, the quantum walk can be highly
efficient for sufficiently high dimensions,
i.e. $d>4$. However, here we evidence that on generic graphs the dimension
does not represent the key geometric parameter; indeed, both the (average) coordination number and the fact
that the structure is translationally invariant or not determine the sharpness of the transition.
In fact, on the one hand, a high coordination reduces the distance among
nodes and increases the possibility of interference effects, on the other
hand, (in the absence of a target) translational invariance prevents the emergence of localization
effects \cite{ABM2008}.
In particular, we considered the success probability
$\pi^{\gamma}_{w,s}(t)$ (here $|s\rangle$ and $|w\rangle$ are the initial
and the target state, respectively) as a function of the computation time
$t$ and of a properly tunable parameter $\gamma$ entering the Hamiltonian.
We showed that for highly dimensional ($d>4$)
translationally invariant structures ($5$-dimensional torus) there exists a
narrow range of $\gamma$ around a specific value
$\tilde{\gamma}$ where the ground state $| \psi_0 \rangle$ undergoes a
transition from having a large overlap with $|s\rangle$ to a state with a
large overlap with $|w\rangle$. This corresponds to a sharply peaked
success probability: $\pi^{\gamma}_{w,s}(t)$ displays a set of maxima just
at $\tilde{\gamma}$; the first one is reached after $O(\sqrt{N})$ queries.
Conversely, for structures with low coordination number and/or fractal or
hierarchical topology - such as the cubic ($L_3$) and square ($L_2$) tori,
the DSG, the TF, the CT
and the Cartesian products DSG $\times$ $L_2$ and DSG $\times$ $L_3$ - the
transition from the initial state to the target state takes place over a
wider region of $\gamma$ around the value $\tilde{\gamma}$. As a
consequence of such a spread-out transition, the success probability
displays broadened peaks whose locations depend on time. Therefore, for
the non-translationally invariant structures considered here, even with large
fractal dimension ($\tilde{d} > 4$), the large success probabilities found for
high-dimensional periodic lattices are not recovered.
These results, in agreement with previous findings \cite{childs}, highlight a possible connection between the sharpness of the transition occurring at $\tilde{\gamma}$ and the efficiency of the search algorithm. A mathematical, rigorous proof stating whether a sharp transition is a necessary condition for a good algorithm, which is beyond the aim of this article, could provide a very useful tool for further investigations on quantum search algorithms.
For the DSG we also proved that $\tilde{\gamma}$ scales like
$N^{2/\tilde{d}+\alpha}$ ($-1 \leq \alpha < 0$); interestingly our results
suggest that such a scaling might be generalized to all (exactly
decimable) fractals with spectral dimension $\tilde{d} < 2$.
Apart from the deterministic fractals considered here, it will be
extremely interesting to also consider disordered structures such as
percolation clusters and random graphs characterized by a degree
distribution $P(z)$. These networks display a tunable average degree
$\langle z \rangle$, which, in principle, can take values ranging from $0$
for totally disconnected networks up to $N$ for completely connected
networks. According to the results discussed here, we expect that for a
sufficiently large $\langle z \rangle$ and for a sufficiently peaked
$P(z)$ the transition from the ground state $|s \rangle$ to a state with
significant overlap with $|w \rangle$ occurs sharply and that,
consequently, the CTQW speeds up.
\section*{Acknowledgements} Support from the Deutsche Forschungsgemeinschaft
(DFG) and the Fonds der Chemischen Industrie is gratefully acknowledged. EA
thanks the Italian Foundation ``Angelo della Riccia'' for financial support.
|
\section{Introduction}
Topological solitons are important in many areas of physics, ranging
from high-energy (elementary particle) physics, condensed matter
physics and string theory to cosmology. In this letter, we shall focus
on a system possessing non-Abelian vortices and monopoles: a
supersymmetric gauge theory with $G=USp(2M)$ gauge group, which is
broken to $H=U(1)\times USp(2M-2)$ by the vacuum expectation value
(VEV) of an adjoint scalar field. This breaking gives rise to regular
non-Abelian 't Hooft-Polyakov monopoles. According to
Goddard-Nuyts-Olive-Weinberg
\cite{Goddard:1976qe,Bais:1978fw,Weinberg:1979zt,Weinberg:1982ev},
the non-Abelian monopoles transform according to the dual group of $H$
which in this case is $\tilde{H}= U(1)\times SO(2M-1)$. Several
difficulties in the na\"{i}ve idea of non-Abelian monopoles have been
known for some time, i.e.~the global $H$ group suffering from a
topological obstruction and non-normalizable zero-modes do not allow
the standard quantization and construction of the $H$ multiplets of
monopoles
\cite{Abouelsaood:1982dz,Nelson:1983bu,Balachandran:1982gt,Horvathy:1984yg,Horvathy:1985bp}.
These problems arise in the Coulomb phase of the theory.
As was done in a series of investigations
\cite{Auzzi:2003fs,Auzzi:2003em} for $SU(N)$ gauge theories, we take
one step further here and break the remaining gauge symmetry
completely at a much lower mass scale. This can be realized by the
introduction of an ${\cal N}=2$ breaking term in the superpotential,
giving rise to an effective Fayet-Iliopoulos term. In systems with
such a hierarchical gauge symmetry breaking, the homotopy group-maps
relate the regular monopoles to the non-Abelian vortices arising at
low energies, allowing for a better understanding of the concept of
the non-Abelian monopole itself. Also, this kind of system provides
a (dual) model of a non-Abelian color-confining superconductor,
further motivating its study.
Besides the cases of $SU(N)$ gauge theories extensively studied in
the last several years, this type of analysis has so far been made
only in the case of $SO(N)$ gauge theories \cite{Ferretti:2007rp},
i.e.~with a hierarchical breaking,
$SO(N) \to U(1)\times SO(N-2)\to {\mathbbm 1}$. In the $SO(N)$ systems
the adjoint matter in the high-energy system yields at low energies
exactly the right matter content -- a system with light fundamental
matter, all charged with respect to a common $U(1)$ factor.
As gauge systems with hierarchical symmetry breaking
$G \to H \to {\mathbbm 1}$ and a color-flavor locked symmetry
$H_{C+F}$, have been constructed to date only for the $SU(N)$ and
$SO(N)$ gauge groups \cite{Auzzi:2003fs,Auzzi:2003em,Ferretti:2007rp},
one might wonder to which extent our idea of defining non-Abelian
monopoles through better-understood non-Abelian {\it vortices}
is general.
The central aim -- and the result -- of the present note is to
construct explicitly an analogous system with the unitary symplectic
gauge group, strengthening further our belief that this kind of
approach is of quite a general validity.
Among the many remarkable developments which followed the discovery of
genuine non-Abelian vortices (vortices with continuous non-Abelian
moduli) in Refs.~\cite{Hanany:2003hp,Auzzi:2003fs} is the moduli
matrix formalism \cite{Isozumi:2004jc,Eto:2005yh} (see review
\cite{Eto:2006pg}), first constructed
for domain walls. This formalism made it possible to uncover the full
moduli space of these non-Abelian vortices, first in the
$U(N)\sim (U(1)\times SU(N))/\mathbb{Z}_N$ theories and subsequently
in models with generic gauge groups \cite{Eto:2008yi}. Finally, in
Ref.~\cite{Eto:2009bg} an in-depth study of the non-Abelian vortices
including the cases of the $(U(1)\times USp(2M))/\mathbb{Z}_2$ gauge
group, has been carried out.
The system considered in this note reduces at low energies, as we
shall show, to the
$(U(1)\times USp(2M))/\mathbb{Z}_2$ models investigated in
Ref.~\cite{Eto:2009bg}; the properties of the vortex moduli space
found there then give detailed exact information about the massive
non-Abelian monopoles.
\section{$USp(2M)$ theory with matter in the fundamental
representation}
Let us first briefly review the superpotential for $N_{\rm f}$ fundamental
hypermultiplets in the $USp(2M)$ gauge theory with $\mathcal{N}=2$
extended supersymmetry in $3+1$ dimensions
\begin{eqnarray} \sqrt{2}\sum_{i=1}^{N_{\rm f}}\tilde{q}_a^i\Phi^{ab}q_b^i \ , \end{eqnarray}
where $i$ denotes the flavor index and $a,b=1,\ldots,2M$ denote the
color indices.
Due to the pseudo-real nature of $USp$ matter fields, we can by a
change of basis
\begin{eqnarray} q^i = \frac{1}{\sqrt{2}}\left(Q^i+iQ^{N_{\rm f}+i}\right) \ , \ \
\tilde{q}^i = \frac{1}{\sqrt{2}}\left(Q^i-iQ^{N_{\rm f}+i}\right) \ ,
\label{eq:basischange}
\end{eqnarray}
write the superpotential as
\begin{eqnarray} \frac{1}{\sqrt{2}}\sum_{i=1}^{2N_{\rm f}} Q_a^i\Phi^{ab}Q_b^i \ , \end{eqnarray}
where we have used the fact that $\Phi^{ab}=\Phi^{ba}$ is symmetric
and we use a notation where the color indices are raised and lowered
with the invariant rank-two tensor of $USp(2M)$
\begin{eqnarray} J^{\rm T} = - J \ , \ \ J^\dag J = \mathbf{1}_{2M} \ , \end{eqnarray}
which we choose to be the skew-diagonal matrix as usual.
The (global) flavor symmetry which the theory at hand possesses is
$O(2N_{\rm f})$.
The mass term is
\begin{eqnarray} \sum_{i,j=1}^{2N_{\rm f}}\frac{m_{ij}}{2}Q_a^iJ^{ab}Q_b^j \ , \end{eqnarray}
where $m_{ij}=\hat{m}_iJ_{ij}$ is anti-symmetric. The flavor symmetry
is now $O(2N_{\rm f})\cap USp(2N_{\rm f})\sim U(N_{\rm f})$.
\section{$USp(2M)$ theory with matter in the adjoint representation }
To construct a system with a hierarchical gauge symmetry breaking as
explained in the Introduction we use the matter fields (squarks) in the
adjoint representation rather than in the fundamental
representation. As in the previous case we start with the matter
fields in the basis
\begin{eqnarray} \sqrt{2}\sum_{i=1}^{N_{\rm f}}{\rm Tr}\left\{
\tilde{q}^i\left[\Phi,q^i\right]\right\} \ , \end{eqnarray}
while by the change of basis (\ref{eq:basischange}) we obtain
\begin{align}
\mathcal{W}_{\rm Adj,Yukawa} &=
\frac{i}{\sqrt{2}}
\sum_{i,j=1}^{2N_{\rm f}}J_{ij}{\rm Tr}\left\{Q^i\left[\Phi,Q^j\right]\right\}
\nonumber\\
&=i\sqrt{2}\sum_{i,j=1}^{2N_{\rm f}}J_{ij}{\rm Tr}\left\{Q^i\Phi Q^j\right\}
\ ,
\end{align}
with $J^{\rm T}=-J, \ J^\dag J = \mathbf{1}_{2N_{\rm f}}$ being the rank-two
invariant tensor of
$USp(2N_{\rm f})$ \footnote{Notice that the flavor symmetry of this system is
different with respect to the usual $O(2N_{\rm f})$ as possessed by the
system with fundamental matter multiplets. }, whereas
the mass term is now
\begin{align}
\sum_{i,j=1}^{2N_{\rm f}}\frac{m_{ij}}{2}
{\rm Tr}\left\{Q^i Q^j\right\} \ ,
\end{align}
and needs to be symmetric in order not to vanish. We shall choose
$m_{ij}=\hat{m}_i \,\tilde{J}_{ij}$, where $\tilde{J}$ is the
symmetric invariant tensor of $SO(2N_{\rm f})$
\begin{eqnarray} \tilde{J}^{\rm T} = \tilde{J} \ , \ \ \tilde{J}^\dag \tilde{J} =
\mathbf{1}_{2N_{\rm f}} \ , \end{eqnarray}
where we again use the skew-diagonal basis. The global flavor symmetry
of our system is thus $USp(2N_{\rm f})\cap O(2N_{\rm f})\sim U(N_{\rm f})$.
\section{$\mathcal{N}=1$ deformation}
Finally, we will add a soft supersymmetry breaking term as
$\mu\,{\rm Tr}\,\Phi^2$ to the adjoint theory and hence we have the
superpotential
\begin{align}
\mathcal{W}_{\rm Adj} = &\
i\sqrt{2}
\sum_{i,j=1}^{2N_{\rm f}}J_{ij}{\rm Tr}\left\{Q^i\Phi Q^j\right\}
+\!\sum_{i,j=1}^{2N_{\rm f}}\frac{m_{ij}}{2}
{\rm Tr}\left\{Q^i Q^j\right\} \nonumber\\
&+\frac{\mu}{2}{\rm Tr}\left\{\Phi^2\right\} \ ,
\end{align}
which gives rise to the following vacuum equations
\begin{align}
J_{ij}\left[Q^j,\Phi\right] + \frac{i}{\sqrt{2}}m_{ij}Q^j &= 0 \ ,
\ \ i=1,\ldots,2N_{\rm f} \ , \\
J_{ij}\left[Q^i,Q^j\right] + i\sqrt{2}\mu\Phi &= 0 \ ,
\end{align}
(repeated indices are summed over) together with the $D$-term
conditions.
First a word on what we expect. From group theory we know that the
adjoint representation of $USp(2M)$ splits as
\cite{Slansky:1981yr}
\begin{align}
USp(2M) &\supset SU(2)\times USp(2M-2) \ , \\
{\rm Adj} &= ({\rm Adj},\mathbbm{1}) + (\mathbbm{1},{\rm Adj})
+(\square,\square) \ , \nonumber
\end{align}
($M>1$).
Actually, we are interested only in the $U(1)$ subgroup of $SU(2)$ so
the relevant decomposition reads
\begin{align}
USp(2M) &\supset U(1)\times USp(2M-2) \ , \label{eq:fund} \\
{\rm Adj} &= 3\, (0, {\mathbbm 1}) + (0,{\rm Adj})
+(1,\square) + (-1,\square) \ . \nonumber
\end{align}
We require the system to be such that only the fields in the
fundamental representation in the low-energy $USp(2M-2)$ remain light,
other fields with no $U(1)$ charges all becoming massive, with a mass
of the order $\mathcal{O}(m)$. Furthermore, only one set of
fundamentals will remain light, either the one with positive $U(1)$
charge or the one with negative charge in Eq.~(\ref{eq:fund}).
We choose the VEV of $\Phi$ as
\begin{eqnarray} \langle\Phi\rangle = \epsilon \
{\rm diag}(m,\underbrace{0,\ldots,0}_{M-1},-m,\underbrace{0,\ldots,0}_{M-1})
\equiv \epsilon\, \Phi_0 \ , \ \end{eqnarray}
and the mass parameters as
\begin{eqnarray} m_{ij}= -i\sqrt{2}\,m \, \tilde{J}_{ij} \ , \ \
\mu = -i\sqrt{2} \, \nu \ , \label{massterms} \end{eqnarray}
where again $\tilde{J}=\tilde{J}^{\rm T}$ is the invariant tensor of
$SO(2N_{\rm f})$.
In order to have a separation of scales in the hierarchical gauge
symmetry breaking, we take $m\gg \nu$.
$\epsilon=\pm$ is the sign that will select which
fundamental fields will become light, with positive or
negative $U(1)$ charge, respectively.
Accordingly, we make an Ansatz $Q^{N_{\rm f}+i}=(Q^i)^\dag$, which solves
the $D$-flatness conditions.
This Ansatz together with the masses taken as in Eq.~(\ref{massterms})
reduces the vacuum equations to
\begin{align}
\left[\Phi_0,Q^i\right] + \epsilon \, m \, Q^i &= 0 \ , \ \ i=1,\ldots,N_{\rm f}\ ,
\label{eq:vac1}\\
\sum_{i=1}^{N_{\rm f}}\left[Q^i, Q^{i\dag}\right] + \epsilon\, \nu\, \Phi_0 &= 0 \ .
\label{eq:vac2}
\end{align}
The light fields are then seen to correspond to the non-trivial
eigenvectors of $[ \Phi_0, \cdot\,]$ with eigenvalue $-\epsilon\, m$ and
they in turn condense by Eq.~(\ref{eq:vac2}).
Without loss of generality, we can choose the light
fields to be the ones with positive $U(1)$ charge and set
$\epsilon:=+$.
Such eigenvectors are found to be $h^{i}(x)$ where
\begin{align}
Q^i = Q_a^i \, t^a = h_{\alpha}^{i} \, K^\alpha +
h_{M-1+\alpha}^{i}\, L^\alpha \ ,
\end{align}
where $i=1,\ldots,N_{\rm f}$ is the flavor index and $a=1,\ldots,M(2M+1)$
is the adjoint color index and finally $\alpha=1,\ldots,M-1$ is half of
the fundamental color index for $USp(2M-2)$. The matrices
$K,L\in\mathfrak{usp}(2M)^{\mathbb{C}}$ are
\begin{align}
{(K^\alpha)_i}^j &= \frac{1}{2}
\left({\delta_{1+\alpha,i}} \delta^{1,j} -
{\delta_{M+1,i}} \delta^{M+1+\alpha,j}\right) \ ,
\\
{(L^\alpha)_i}^j &= \frac{1}{2}
\left({\delta_{M+1+\alpha,i}} \delta^{1,j} +
{\delta_{M+1,i}} \delta^{1+\alpha,j}\right) \ .
\end{align}
If we instead wanted the fundamental fields with negative $U(1)$
charge to be the light fields, we should set $\epsilon:=-$ and the
eigenvectors would be
\begin{align}
Q^i = Q_a^i \, t^a = h_{\alpha}^{i} \, \left(K^\alpha\right)^{\rm T} +
h_{M-1+\alpha}^{i}\, \left(L^\alpha\right)^{\rm T} \ ,
\end{align}
see Appendix for details.
Calculating now explicitly the commutator, Eq.~(\ref{eq:vac2}) gives
rise to the $D$-flatness conditions of the $U(1)\times USp(2M-2)$
low-energy
theory with fundamental matter content.
Let us make the following definition
\begin{eqnarray} h^i = \begin{pmatrix} h^i_{\alpha}
\\ h^i_{M-1+\alpha} \end{pmatrix} \equiv \begin{pmatrix}
k^i_{\alpha} \\ \ell^i_{\alpha} \end{pmatrix} \ , \end{eqnarray}
with $k,\ell$ being $(M-1)$-vectors of color and $i$ is the flavor
index. Then, independently of the choice of the sign $\epsilon$,
($4\times$) Eq.~(\ref{eq:vac2}) reads
\begin{align}
\begin{pmatrix}
-h^{i\dag}h^i + 4\nu m & 0 &
0 & 0\\
0 & \mathbf{A} & 0 & \mathbf{B}^\dag \\
0 & 0 & h^{i\dag}h^i - 4\nu m & 0\\
0 & \mathbf{B} & 0 & -\mathbf{A}^{\rm T}
\end{pmatrix} = 0 \ ,
\label{eq:vacfinal1}
\end{align}
from which the Abelian $D$-term
constraint (in the low-energy $\mathcal{N}=1$ theory) is easily read
off. Now for the non-Abelian part, we find the form of the matrices
${\bf A}, {\bf B}$:
\begin{align}
\mathbf{A} \equiv k^i k^{i\dag} -
\left(\ell^i \ell^{i\dag}\right)^{\rm T} \ , \ \
\mathbf{B} \equiv \ell^i k^{i\dag} +
\left(\ell^i k^{i\dag}\right)^{\rm T} \ ,
\end{align}
where ${\bf B}^{\rm T}={\bf B}$ is manifest.
Using that
\begin{eqnarray} h^i h^{i\dag} =
\begin{pmatrix}
k^i k^{i\dag} & k^i \ell^{i\dag} \\
\ell^i k^{i\dag} & \ell^i \ell^{i\dag}
\end{pmatrix} \ , \label{eq:hhdag_explicit}
\end{eqnarray}
together with the explicit form of the generators
\begin{eqnarray}
t^n =
\begin{pmatrix}
\alpha & \beta_{S} \\
\beta_{S}^\dag & -\alpha^{\rm T}
\end{pmatrix} \ , \ \
\alpha^\dag = \alpha \ , \ \
\beta_{S}^{\rm T} = \beta_{S} \ ,
\end{eqnarray}
we obtain
\begin{align}
0 &= {\rm Tr}\left\{h^i h^{i\dag}t^n\right\} \nonumber\\
&= {\rm Tr}\left\{{\bf A}\alpha\right\} +
\frac{1}{2}{\rm Tr}\left\{{\bf B}\beta_{S}\right\} +
\frac{1}{2}{\rm Tr}\left\{{\bf B}^\dag\beta_{S}^\dag\right\} \ ,
\label{eq:Dterm2}
\end{align}
for \emph{all} $\alpha,\beta_{S}$, which forces ${\bf A}={\bf B}=0$,
where we have used the fact that ${\bf B}$ is symmetric.
Now as a check, we can count the number of constraints of
${\bf A}={\bf B}=0$ yielding $M'(2M'+1)$ with $M'\equiv M-1$, which
indeed
coincides with the number of constraints in Eq.~(\ref{eq:Dterm2}).
Hence, using a color-flavor matrix notation
${(h h^\dag)_{\alpha}}^{\alpha'} = {h_\alpha}^i {(h^\dag)_i}^{\alpha'}
= h^i h^{i\dag}$ we can write the
Eqs.~(\ref{eq:vac1})-(\ref{eq:vac2}) as
\begin{align}
{\rm Tr}\left\{h h^\dag\right\} &= 4 \nu m \ , \label{eq:Dflat1} \\
{\rm Tr}\left\{h h^\dag t^n\right\} &= 0 \ , \label{eq:Dflat2}
\end{align}
which are the $D$-term conditions appropriate for constructing
non-Abelian BPS vortices and $t^n\in\mathfrak{usp}(2M-2)$ and
$n=1,\ldots,(M-1)(2(M-1)+1)$ and specifically for the fundamental
representation, as we intended.
These vortices have already been studied in the low-energy theory in
Ref.~\cite{Eto:2009bg}.
A comment in store is to emphasize the importance of identifying the
``light mass'' degrees of freedom in the symmetry breaking.
In order to have a vacuum that breaks completely the local gauge
symmetry, allowing at the same time for an intact global color-flavor
symmetry, we shall choose the number of flavor multiplets to be
$N_{\rm f}=2M-2$. Thus $h$ is a square matrix with the following VEV
\begin{eqnarray} \langle h\rangle =
\frac{\sqrt{\xi}}{\sqrt[4]{M-1}}\mathbf{1}_{2M-2} \ . \end{eqnarray}
For completeness, let us write down the low-energy effective action
for the light fundamental fields
\begin{align}
\mathcal{L} =\;&
-\frac{1}{4g^2}F_{\mu\nu}^n F^{n\mu\nu}
-\frac{1}{4e^2}F_{\mu\nu}^0 F^{0\mu\nu}
+{\rm Tr}\left(\mathcal{D}_\mu h\right)\left(\mathcal{D}^\mu h\right)^\dag \nonumber\\ &
-\frac{e^2}{2}\left|{\rm Tr}\left(h h^\dag t^0\right) - \xi\right|^2
-\frac{g^2}{2}\left|{\rm Tr}\left(h h^\dag t^n\right)\right|^2 \ ,
\end{align}
where we have rescaled the fields $h\to \sqrt{2}gh$ and
$\xi\equiv \nu m \to e\,\xi/(\sqrt{2M-2})$ and defined the $U(1)$
generator
\begin{eqnarray} t^0 \equiv \frac{\mathbf{1}_{2M-2}}{2\sqrt{M-1}} \ , \end{eqnarray}
and finally the index $n=1,\ldots,(M-1)(2(M-1)+1)$.
Due to different renormalization effects of the subgroups after
the gauge symmetry breaking, we use $e$ to denote the coupling for
$U(1)$ and $g$ for $USp(2M-2)$. Note that we have neglected higher
order terms in $\nu/m$ which will give rise to non-BPS terms in the
low-energy action for vortices, hence as already mentioned it is a
near-BPS system.
As a final remark, let us note that in the strictly BPS limit our
low-energy system would have a large vortex-moduli space including the
so-called semi-local vortices \cite{Eto:2009bg}. The latter do not
have a definite transverse size, and would not confine the monopole at
their ends.
However, our system (with $m \gg \nu$) is almost, but not exactly,
BPS. When small non-BPS corrections arising from the high-energy gauge
symmetry breaking are taken into account, we expect the vortex moduli,
which are not related to the exact global symmetry of the system, to
disappear.
This has been explicitly shown \cite{Auzzi:2008wm} in the case of the
vortex moduli in the $SU(N+1)$ theory with $N_{f}> N$,
spontaneously broken at two scales,
$SU(N+1) \to U(N) \to {\mathbbm 1}.$
\section{Conclusion}
Our system is characterized by the hierarchical gauge symmetry
breaking
\begin{eqnarray} G \ \mathop{\longrightarrow}^{m}\ H \
\mathop{\longrightarrow}^{2\sqrt{\nu m}}\ \mathbbm{1} \;. \end{eqnarray}
As all the fields in the underlying theory are in the adjoint
representation, we actually have $G=USp(2M)/\mathbb{Z}_2 $. The
(light) matter content of the low-energy theory shows also that
$H=\left(U(1)\times USp(2M-2)\right)/\mathbb{Z}_2$. Since
$\pi_1(G)=\mathbb{Z}_2$, the exact homotopy sequence tells us that
\begin{eqnarray} \pi_2\left(G/H\right) \sim \pi_1\left(H\right)/ \mathbb{Z}_2
\;: \end{eqnarray}
the regular monopoles arising at the high-mass scale breaking
are confined by the doubly-wound vortices of the low-energy theory.
The results of Ref.~\cite{Eto:2009bg}, which hold in a vacuum with the
color-flavor locked phase, indicate that the minimal winding vortices
of the low-energy $U(1)\times USp(2M-2)$ system, which are stable in
the full theory as $\pi_1(G)=\mathbb{Z}_2$, appear classified
according to the {\it spinor representation} of a dual (color-flavor)
$SO(2M-1)$ symmetry group. The regular monopoles of our system,
associated with the doubly-wound vortices, are then predicted
to transform according to various representations including the
{\it vector representation} of the $SO(2M-1)$ group, reminiscent of
the GNO duality. These group-theoretic features of our vortex-monopole
complex are under a careful scrutiny at present, and will be presented
elsewhere.
\section*{Acknowledgments}
We thank Minoru Eto, Muneto Nitta and Keisuke Ohashi for useful discussions.
|
\section{Introduction}
The \emph{Minkowski sum} of two polytopes is defined as
$P_1+P_2=\{x_1+x_2\;:\;x_1\in P_1,\;x_2\in P_2\}$. Minkowski sums are
of interest in various fields of theoretical and applied
mathematics. While some applications only require sums of two
polytopes in low dimensions (e.g. motion planning~\cite{Halperin04}\cite{Lozano79}),
others use iterative sums of many polytopes in higher
dimensions~\cite{Pachter05}\cite{Zhang09}. It is therefore desirable to
study the complexity of such
sums.
A trivial bound on the number of vertices of a sum is found as
follows. Every vertex of a Minkowski sum decomposes into a sum
of vertices of the summands. Therefore, there cannot be more vertices
in the sum than there are possible decompositions. Thus, a trivial
bound on the number of vertices in a Minkowski sum is the product of
the number of vertices in the summands. That is, if $P_1,\ldots,P_r$
are polytopes, and $f_0(P)$ is the number of vertices of a polytope,
then
$f_0(P_1+\cdots+P_r) \leq f_0(P_1)\cdots f_0(P_r)$.
If we sum $r$ polytopes in dimension $d$, with $r<d$, then the trivial
bound is tight, that is, it is possible to choose summands with any
number of vertices so that their sum has as many vertices as the trivial
bound~\cite{Fukuda07}.
However, if we sum $r$ polytopes in dimension $d$ with $r\geq d$, the
trivial bound cannot be reached, except when summing $d$
segments~\cite{Sanyal07}\cite{Weibel07}. We assume here and in the
rest of the article that all polytopes have at least two vertices,
since a summand of only one vertex can be ignored without changing the
properties of the sum.
The f-vector of a polytope encodes its number of faces of different
dimensions. Maximal f-vectors are obtained for a particular class of
Minkowski sums, called sums of polytopes \emph{in general
orientations}. We will therefore restrict our study to such sums.
We recently presented in~\cite{Fogel09} a result on sums of
$3$-dimensional polytopes in general orientations. We showed that the
number of vertices in a sum of $r$ summands can be deduced from the
number of vertices in the summands and the number of vertices in sums
of each of the ${r\choose 2}$ pairs of summands. Using this result, we
found a tight upper bound on the number of vertices and facets in sums
of $3$-dimensional polytopes.
The basic reasoning of this previous result is to define a unique
witness, called \emph{western-most corner}, for all but two vertices of
a polytope. These witnesses have the property that a western-most corner
for a Minkowski sum of any number of summands is also necessarily a
western-most corner for the sum of some pair of the summands. So the
number of western-most corners in the total sum, and thus its number
of vertices, can be found by examining sums of one or two summands only.
This prompted us to examine the possibility of extending the reasoning
to higher dimensions and other faces, which resulted in this
article. Our main result is presented in Theorem \ref{thm:main}. It is
a linear relation between the number of faces of a sum of $r$
polytopes and the number of faces in the sums of less than $d$ of the
summand polytopes:
\begin{thm}\label{thm:main}
Let $P_1,\ldots,P_r$ be $d$-dimensional polytopes in general
orientations, $r\geq d$, and each polytope full-dimensional. For any
$k$ in $0,\ldots,d-1$,
$$
f_k(P_1+\cdots+P_r)-\alpha=\sum_{j=1}^{d-1}(-1)^{d-1-j} {{r-1-j} \choose {d-1-j}} \sum_{S\in \mathcal{C}_j^r}(f_k(P_S)-\alpha),
$$
where $\mathcal{C}_j^r$ is the family of subsets of $\{1,\ldots,r\}$ of
cardinality $j$, $P_S$ is the sum of polytopes $\sum_{i\in S}P_i$;
$\alpha=2$ if $k=0$ and $d$ is odd, $\alpha=0$ otherwise.
\end{thm}
A slightly more general result also applies when summands are not
full-dimensional. The intuitive explanation of the theorem is that for
any face of the whole sum, we can find a witness of its existence by
examining the faces of the same dimension in sums of $d-1$
summands. However, if that witness exists in some sum of $d-2$
summands, we will find it in many different sums of $d-1$ summands. So
we need to offset this by removing an appropriate number of times the
witnesses in sums of $d-2$ summands. But that in turn removes too many
times witnesses that exist in some sum of $d-3$ summands, so we need
to add them back, etc. This implies that the total sum is smaller
than the term for $j=d-1$:
\begin{cor}\label{cor}
Let $P_1,\ldots,P_r$ be $d$-dimensional polytopes in general
orientations, $r\geq d$, and each polytope full-dimensional. For any
$k$ in $0,\ldots,d-1$,
$$
f_k(P_1+\cdots+P_r)\leq \sum_{S\in \mathcal{C}_{d-1}^r}f_k(P_S).
$$
\end{cor}
From this result, we deduce bounds for the maximum possible
number of vertices in a Minkowski sum of polytopes, for fixed number of
vertices in the summands.
We find in particular that a sum of $r$ polytopes in dimension $d$,
$r\geq d$, where summands have $n$ vertices in total, has less than ${n
\choose {d-1}}$ vertices, which is in $O(n^{d-1})$. In the case
where each summand has at most $n$ vertices, then the number of
vertices of the sum is less than ${r \choose {d-1}} n^{d-1}$, which is
in $O(r^{d-1}n^{d-1})$. This is better than the previous known bound
which was in $O(r^{d-1}n^{2(d-1)})$~\cite{Gritzmann93}.
In the rest of the article, we shortly present the theory in
Section~\ref{sec:mink}. We first give an introduction to the concepts
of west and western-most corner in three dimensions in
Section~\ref{sec:3d}, then extend them formally to higher dimensions
in Section~\ref{sec:wit}. We examine in Section~\ref{sec:max} what are
the maximum possible number of faces of a Minkowski sum and conclude
in Section~\ref{sec:con}.
\section{Minkowski sums}\label{sec:mink}
Let $P_1,\ldots,P_r$ be given polytopes. Their Minkowski sum is the polytope
defined as $P_1+\cdots+P_r = \{x_1+\cdots+x_r\::\:x_i\in P_i,\;\forall i\}$.
We assume in the following, and in the rest of the article, that
every polytope is full-dimensional.
A nontrivial \emph{face} of a polytope $P$ in dimension $d$ is the
intersection of $P$ with a support hyperplane of $P$. Vertices, edges,
facets, ridges are the faces of dimension $0$, $1$, $d-1$ and $d-2$
respectively. Thus, we can associate to each vector in the unit sphere
$S^{d-1}$ a face of the polytope, which is the intersection of the polytope
with the support hyperplane to which the vector is outwardly normal:
$\mathcal{S}(P;l) = \{x\in P\;:\:\langle l,x\rangle \geq \langle l,y\rangle,\;\forall y\in P\}$.
Conversely, each face $F$ of a $d$-dimensional polytope $P$ can be
associated with a region of the sphere $S^{d-1}$, called the
\emph{normal region}, which is the set of unit vectors outwardly
normal to some support hyperplane of $P$ whose intersection with $P$
is $F$: $\mathcal{N}(F;P)=\{l\in
S^{d-1}\;|\,F=\mathcal{S}(P;l)\}=\{l\in S^{d-1}\;|\;\langle
l,x\rangle>\langle l,y\rangle,\;\forall x\in F,\;y\in P\setminus
F\}$. The normal region of a facet of $P$ is thus a single point of
$S^{d-1}$, corresponding to the unit vector outwardly normal to the
facet. The normal region of a face of dimension $k$ is a relatively
open subset of $S^{d-1}$ of dimension $d-1-k$.
We call a subset of the sphere $S^{d-1}$ \emph{spherically convex} if
for any two points in the subset, any shortest arc of great circle
between the two points is inside the subset.\footnote{There exist
different definitions of convexity on a sphere. Note that according
to this one, the only convex set containing antipodal points is the
whole sphere.} If the polytope $P$ is full-dimensional, the normal
regions of faces of $P$ are all disjoint, relatively open and
spherically convex. They determine a subdivision of $S^{d-1}$ into a
spherical cell complex which we call the \emph{Gaussian map} of the polytope:
$\mathcal{G}(P)=\{\mathcal{N}(F;P)\;:\;F\mbox{ face of }P\}$.
A property of Minkowski sums is that faces of the sum have a unique
\emph{decomposition} in faces of the summand. Let $F$ be a face of the
Minkowski sum $P=P_1+\cdots+P_r$, and $l$ be in $\mathcal{N}(F;P)$.
Then $F=F_1+\cdots+F_r$, where $F_i= \mathcal{S}(P_i;l)$ is a face of
$P_i$. We can deduce that the normal region of a face of the sum is
equal to the intersection of the normal regions of the faces in its
decomposition:
$\mathcal{N}(F;P)=\mathcal{N}(F_1;P_1)\cap\cdots\cap\mathcal{N}(F_r;P_r)$.
Thus the Gaussian map of the Minkowski sum is the \emph{common
refinement} of the Gaussian map of the summands:
$$
\mathcal{G}(P_1+\cdots+P_r)=
\{\mathcal{N}(F_1;P_1)\cap\cdots\cap\mathcal{N}(F_r;P_r)\;:\;
F_i\mbox{ face of }P_i\}.
$$
A polytope and its Gaussian map being dual structures, it is possible
to study the number of faces of a polytope by studying the number of
cells of its Gaussian map.
We say that a face of a Minkowski sum has an \emph{exact
decomposition} $F=F_1+\cdots+F_r$ when its dimension is the sum of
the dimension of the faces in its decomposition:
$\dim(F)=\dim(F_1)+\cdots+\dim(F_r)$. That is, the decomposition is
exact when there are no two parallel segments inside different faces
in the decomposition. We say that polytopes are \emph{in general
orientations} when all faces of their Minkowski sum have an exact
decomposition.
For fixed f-vectors of summands, the maximum number of faces of any
dimension in the sum can always be reached by summands in general
orientations~\cite{Fukuda08}. Therefore, we can assume summands are in
general orientations when looking for sums with maximum number of
faces.
Let $F$ be a face of the Minkowski sum $P=P_1+\cdots+P_r$ of
$d$-dimensional polytopes in general orientations. The face $F$
decomposes into a sum $F_1+\cdots+F_r$ of faces of the summands, with
$\dim(F)=\dim(F_1)+\cdots+\dim(F_r)$. Even if $r\geq d$, there are
therefore at most $\dim(F)$ faces in the decomposition that have a
dimension of more than $0$. Let the \emph{support} $I_F\subseteq
\{1,\ldots,r\}$ of $F$ be the set of indices of these faces, with
$|I_F|\leq \dim(F)$. Note that for any subface $G$ of $F$, $G$
decomposes into a sum $G_1+\cdots+G_r$, where $G_i\subseteq F_i$ for
all $i$; and so, $I_G\subseteq I_F$.
For any $S=\{i_1,\ldots,i_s\}$ subset of $\{1,\ldots,r\}$, let us
define the \emph{partial sum} $P_S=P_{i_1}+\cdots+P_{i_s}$.
\begin{lemma}
Let $F$ be a facet of a Minkowski sum $P=P_1+\cdots+P_r$ of
$d$-dimensional polytopes in general orientations. Its normal region
$\mathcal{N}(F;P)$ is a node of $\mathcal{G}(P)$. Then
$\mathcal{N}(F;P)$ is also a node of the Gaussian map
$\mathcal{G}(P_S)$ of a partial sum if and only if $I_F\subseteq S$.
\end{lemma}
\begin{proof}
Let $F_1+\cdots+F_r$ be the decomposition of $F$, with $\dim(F_i)>0$
if and only if $i\in I_F$. Since the summands are in general
orientations, the decomposition is exact and
$\dim(F)=d-1=\dim(F_1)+\cdots+\dim(F_r)=\sum_{i\in
I_F}\dim(F_i)$. The normal region $\mathcal{N}(F;P)$ contains a
single unit vector $l$; and $\mathcal{N}(F;P)$ is a node of
$\mathcal{G}(P_S)$ if and only if $\dim(\mathcal{S}(P_S;l))=d-1$.
Again, the decomposition is exact and
$\dim(\mathcal{S}(P_S;l))=\sum_{i\in
S}\dim(\mathcal{S}(P_i;l))=\sum_{i\in S}\dim(F_i)$. Since
$\dim(F_i)>0$ if and only if $i\in I_F$, and $\sum_{i\in
I_F}\dim(F_i)=d-1$, the result is obvious.
\end{proof}
\section{Sums of polytopes in dimension 3}\label{sec:3d}
To facilitate the comprehension of the proof of
Theorem~\ref{thm:main}, we present informally in this section the
argument for three dimensions, where it is more readily understood. We
present the full proof for general dimensions in
Section~\ref{sec:wit}. The result for three dimensions has already
been published, though only for vertices~\cite{Fogel09}.
In dimension $3$, the Gaussian map of a polytope is a spherical cell
complex of $S^2$, which can be described as a planar graph embedded in
$S^2$. The normal regions of facets, edges and vertices of the
polytope are nodes, edges and faces of the graph respectively. Note
that the normal regions of edges, edges of the graph, are arcs of
great circles of $S^2$. The Gaussian map of a Minkowski sum is the
\emph{overlay} of the Gaussian maps of the summands.
\begin{wrapfigure}{r}{3.4cm}
\begin{center}
\psset{unit=1.6cm,shortput=nab,linewidth=0.5pt,arrowsize=2pt 3}
\pspicture(-1,-1)(1,1)%
\psline[linecolor=blue,fillstyle=solid,fillcolor=blue](0,-.446)(0.1,-.38)(0.1,-.58)
\psline[linecolor=red,fillstyle=solid,fillcolor=red](0.43,-.08)(0.52,-.13)(0.52,-.05)
\psline[linecolor=blue,fillstyle=solid,fillcolor=blue](0.43,-.08)(0.52,0.02)(0.52,-.05)
\psline[linecolor=blue,fillstyle=solid,fillcolor=blue](-0.92,0.08)(-0.85,0)(-0.85,0.1)
\psline[linecolor=red,fillstyle=solid,fillcolor=red](-0.92,0.08)(-0.85,0.22)(-0.85,0.1)
\psline[linecolor=red,fillstyle=solid,fillcolor=red](-0.13,0.24)(-0.03,0.2)(-0.03,0.28)
\psline[linecolor=green,fillstyle=solid,fillcolor=green](-0.13,0.24)(-0.03,0.34)(-0.03,0.28)
\psline[linecolor=blue,fillstyle=solid,fillcolor=blue](-0.52,-0.255)(-0.43,-0.3)(-0.43,-0.23)
\psline[linecolor=green,fillstyle=solid,fillcolor=green](-0.52,-0.255)(-0.43,-0.13)(-0.43,-0.23)
\psset{linecolor=blue}
\parametricplot{-39.5}{65}{45 dup 30 dup t cos exch sin mul exch cos t sin mul 3 -1 roll cos mul add exch sin t sin mul}%
\parametricplot{-29}{64}{-30 dup -60 dup t cos exch sin mul exch cos t sin mul 3 -1 roll cos mul add exch sin t sin mul}%
\parametricplot{31}{95}{-60 dup -17 dup t cos exch sin mul exch cos t sin mul 3 -1 roll cos mul add exch sin t sin mul}%
\psset{linecolor=red}
\parametricplot{-30}{35}{55 dup -71 dup t cos exch sin mul exch cos t sin mul 3 -1 roll cos mul add exch sin t sin mul}%
\parametricplot{-33}{-75}{-60 dup -31 dup t cos exch sin mul exch cos t sin mul 3 -1 roll cos mul add exch sin t sin mul}%
\parametricplot{-70}{75}{-30 dup 18 dup t cos exch sin mul exch cos t sin mul 3 -1 roll cos mul add exch sin t sin mul}%
\psset{linecolor=green}
\parametricplot{-80}{100}{50 dup -20 dup t cos exch sin mul exch cos t sin mul 3 -1 roll cos mul add exch sin t sin mul}%
\psset{linecolor=black}
\parametricplot{-180}{180}{90 dup 90 dup t cos exch sin mul exch cos t sin mul 3 -1 roll cos mul add exch sin t sin mul
\psdots(0,0.9)
\psdots[linecolor=lightgray](0,-0.9)
\psellipticarc{<-}(0,1.2)(0.3,0.1){110}{45}
\endpspicture%
\end{center}
\caption{Example of a Gaussian map, overlay of three Gaussian maps. Each western-most corner exists in the overlay of one or two Gaussian maps
}\label{fig:wmc}
\end{wrapfigure}
Let $P=P_1+\cdots+P_r$ be a sum of $3$-dimensional polytopes in
general orientations. We choose on $S^2$ two antipodal points in
generic position, so that no edge of $\mathcal{G}(P)$ is aligned with
them. In particular, the points are inside two distinct
faces of $\mathcal{G}(P)$. We call these two points \emph{north pole}
and \emph{south pole}. We define \emph{west} in the usual way with
respect to the poles, as a direction turning around the poles,
clockwise from the north pole.
For any spherically convex subset $C$ of $S^2$ that does not contain
either pole, we define the \emph{western-most point} of $C$ as the
point in the closure of $C$ that is to the west of all points in
$C$. We also define as \emph{western-most corner} of $C$ the subset of
$C$ at distance less than $\varepsilon$ of the western-most point, for
a small $\varepsilon>0$. Note that if $C$ is a cell of
$\mathcal{G}(P)$, the western-most point is a node of $\mathcal{G}(P)$
incident to $C$, which is unique because no edge of $\mathcal{G}(P)$
is aligned with the poles.
The normal region $\mathcal{N}(F;P)$ of a facet $F$ of the Minkowski
sum $P$ is a node of $\mathcal{G}(P)$. Because the summands are in
general orientations, there are no three edges of different
$\mathcal{G}(P_i)$ intersecting in a single point. Thus, a node in
$\mathcal{G}(P)$ is either a node in some $\mathcal{G}(P_i)$, in which
case $I_F=\{i\}$, or it is the intersection of two edges in some
$\mathcal{G}(P_i+P_j)$, in which case $I_F=\{i,j\}$ (See
Figure~\ref{fig:wmc}). So a western-most corner in $\mathcal{G}(P)$ is
always a western-most corner in some $\mathcal{G}(P_i)$, or a
western-most corner in some $\mathcal{G}(P_i+P_j)$ whose western-most
point is the intersection of two edges.
So we can find the number of western-most corners in $\mathcal{G}(P)$
by counting those in all $\mathcal{G}(P_i)$ and those in all
$\mathcal{G}(P_i+P_j)$ whose western-most point is an intersection of
edges. But then, the western-most corners in $\mathcal{G}(P_i+P_j)$
also include those whose western-most point is a node of
$\mathcal{G}(P_i)$ or $\mathcal{G}(P_j)$. Denoting as $w_k(.)$ the
number of western-most corners of $k$-dimensional cells in a Gaussian
map, this means that
$$
w_k(\mathcal{G}(P))=\sum_{i=1}^rw_k(\mathcal{G}(P_i))+ \sum_{1\leq i<j\leq
r}(w_k(\mathcal{G}(P_i+P_j))-w_k(\mathcal{G}(P_i))-w_k(\mathcal{G}(P_j)))
$$
$$
=\sum_{1\leq i<j\leq r}w_k(\mathcal{G}(P_i+P_j))
-(r-2)\sum_{i=1}^rw_k(\mathcal{G}(P_i)),\quad k=0,1,2.
$$
Intuitively, we sum the number of western-most corners in different
$\mathcal{G}(P_i+P_j)$, and subtract $(r-2)$ times the western-most
corners in the $\mathcal{G}(P_i)$, since they are each counted $(r-1)$
times in the first sum.
But since there is one distinct western-most corner for every cell of
a Gaussian map except the two $2$-dimensional cells that contain a
pole, and the cells of a Gaussian map correspond to faces of the
underlying polytope, we have that for any polytope $\mathcal{P}$,
$w_0(\mathcal{G}(\mathcal{P}))=f_2(\mathcal{P})$,
$w_1(\mathcal{G}(\mathcal{P}))=f_1(\mathcal{P})$ and
$w_2(\mathcal{G}(\mathcal{P}))=f_0(\mathcal{P})-2$. Replacing $w_k$ in
the above equation by these, we get Theorem~\ref{thm:main} for $d=3$.
Here is a subtle detail of the argument. Let us say that a point $p$
in the closure of a subset $C$ of the sphere $S^2$ is a \emph{local
optimum} of $C$ if $p$ is the western-most point for the
intersection of $C$ with some open set containing $p$. The reason we
use the direction west is that the level curves for west, the
meridians of geography, are arcs of great circles, i.e. geodesics;
they intersect only once any other geodesic inside a spherically
convex set. This guarantees that all local optima are also
western-most points. This would not be the case had we used the
direction south, because the level curves for south, the parallels,
are not geodesics.
\section{Extension to higher dimensions}\label{sec:wit}
We extend in this section the argument of Section~\ref{sec:3d} to
higher dimensions. First, we extend the definition of west and
western-most corners. We prove that the extension has the same
property that (about) every cell of a Gaussian map has a single
western-most corner.
We also prove that the western-most corner of some cell in the
Gaussian map of a Minkowski sum is also a western-most corner of some
cell in any Gaussian map of a partial sum where its western-most point
is a node of the map. Finally, we present the formula that allows us
to count the number of western-most corners in the Gaussian map of the
Minkowski sum.
Here is a brief summary of the proof. Let $P=P_1+\cdots+P_r$ be a
Minkowski sum of $d$-dimensional polytopes in general
orientations. Because the summands are in general orientations, a node
of $\mathcal{G}(P)$ is also a node of $\mathcal{G}(P_S)$ if and only
if the support of its underlying facet is contained in $S$. This
implies that in all $\mathcal{G}(P_S)$ where a node exists, the local
geometry of the map around the node is the same as in
$\mathcal{G}(P)$. Therefore, if the node is a local optimum in
$\mathcal{G}(P)$, it is also a local optimum in any $\mathcal{G}(P_S)$
of which it is a node. This implies that a western-most corner in
$\mathcal{G}(P)$ is also a western-most corner in $\mathcal{G}(P_S)$
if and only if its western-most point in $\mathcal{G}(P)$ exists in
$\mathcal{G}(P_S)$, i.e. if and only if $S$ contains the support of
the underlying facet of the node. This is what allows us to use a counting
argument for deducing the number of western-most corners in
$\mathcal{G}(P)$ from the number of western-most corners in the
Gaussian map of partial sums. Since there is one western-most corner
per cell of the Gaussian map and face of the underlying polytope, this
allows us to find the number of faces of the sum $P$.
We now extend the definition of west. The Gaussian map
$\mathcal{G}(P)$ subdivides the sphere $S^{d-1}$ into a spherical cell
complex. The normal regions of ridges of $P$ are arcs of great circles
on $S^{d-1}$, edges of the Gaussian map $\mathcal{G}(P)$. Each of
these arcs of great circle is contained into the $2$-dimensional
subspace that is orthogonal to the underlying ridge. In the
$d$-dimensional space containing $S^{d-1}$, we choose a linear
subspace $U$ of dimension $d-2$, so that its intersection with every
$2$-dimensional subspace containing an edge of $\mathcal{G}(P)$ is
just the origin. The next lemma shows that this is always possible.
\begin{lemma}
In a $d$-dimensional space, for any finite family
$\{U_1,\ldots,U_n\}$ of $2$-dimensional linear subspaces, it is
possible to find a linear subspace $U$ of dimension $d-2$ such that
$U\cap U_i=\{\bvec{0}\}$ for any $i$ in $1,\ldots,n$.
\end{lemma}
\begin{proof}
The orthogonal complements $U_i^\perp$ of the subspaces $U_i$ in the
family are of dimension $d-2$. If we choose a vector $\bvec{u}$ that is
not in these orthogonal complements, then for any $i$,
$span(\{\bvec{u}\}\cup U_i^\perp)$ is of dimension $d-1$. If we choose now a
vector $\bvec{v}$ that is not any of these subspaces of dimension
$d-1$, then for any $i$, $span(\{\bvec{u},\bvec{v}\}\cup U_i^\perp)$
is the $d$-dimensional space. Define
$U=span(\{\bvec{u},\bvec{v}\})^\perp$. If a vector is in $U\cap
U_i$, it is orthogonal to $span(\{\bvec{u},\bvec{v}\}\cup
U_i^\perp)$, which is the whole space, and so it is the origin
$\bvec{0}$.
\end{proof}
This is a simple extension of the fact that in three dimensions, for
any number of planes going through the origin, we can choose
a vector that is in none of the planes.
Note that the intersection of $U$ with $S^{d-1}$ is equivalent to the
sphere $S^{d-3}$. If $d=3$, the intersection is the two antipodal
points on $S^2$ that we named north and south poles in
Section~\ref{sec:3d}, and if $d=2$, it is the empty set because
$U=\{\bvec{0}\}$.
We choose an orthonormal basis $\bvec{e_1},\ldots,\bvec{e_d}$ of the
$d$-dimensional space such that
$U=span(\{\bvec{e_3},\ldots,\bvec{e_d}\})$.
We then define successive parametrizations of the spheres
$S^n$, $1\leq n\leq d-1$ as follows:
$$
S^1 = \{\sin(\theta_1)\bvec{e_1}+\cos(\theta_1)\bvec{e_2}\;:\;\theta_1 \in [0,2\pi)\},
$$
$$
S^n = \{\sin(\theta_n)S^{n-1}+\cos(\theta_n)\bvec{e_{n+1}}\;:\;\theta_n \in [0,\pi]\},\quad n=2,\ldots,d-1.
$$
Note that a point of $S^{d-1}$ is in $U$ if and only if
$\sin(\theta_j)=0$ for some $j=2,\ldots,{d-1}$. For any point of
$S^{d-1}$ not in $U$, we define the direction west as
$\dot{\theta}_1$, the direction of augmentation of $\theta_1$. Note
that for $d=3$, it is equivalent to the definition of
Section~\ref{sec:3d}, and for $d=2$, it is a direction running around
$S^1$. In $S^{d-1}$, west is not defined on the subspace $U$ because
$\dot{\theta}_1=0$, so that the intersection of $U$ with $S^{d-1}$ is
a sphere of dimension $d-3$ that plays the same role as poles in three
dimensions.
Recall that in our parametrization of $S^{d-1}$, $\theta_1$ is in
$[0,2\pi)$. Formally, for any points $p$ and $q$ of $S^{d-1}$ that are
not in $U$, we say that $p$ is \emph{to the west} of $q$ if
$\theta_1(p)\in [\theta_1(q),\theta_1(q)+\pi]$ and $\theta_1(q)<\pi$,
or if $\theta_1(p)\in [\theta_1(q),2\pi)\cup[0,\theta_1(q)-\pi]$ and
$\theta_1(q)\geq\pi$.
For any spherically convex subset $C$ of $S^{d-1}$ that does not
intersect $U$, we define the \emph{western-most point} of $C$ as the
point in the closure of $C$ that is to the west of all points in
$C$. The next lemma, proved in Appendix~\ref{app:ls},
shows that the western-most point exists.
\begin{lemma}\label{lem:sense}
If a spherically convex subset $C$ of $S^{d-1}$ does
not intersect $U$, it is in a hemisphere defined by $\theta_1\in
[\alpha,\alpha+\pi]$ or $\theta_1\in [0,\alpha]\cup[\alpha+\pi,2\pi)$ for
some $\alpha\in[0,\pi)$.
\end{lemma}
We also define as \emph{western-most corner} of $C$ the subset of $C$
at distance less than $\varepsilon$ of the western-most point, where
$\varepsilon>0$ is smaller than the distance between any two
non-incident cells in $\mathcal{G}(P)$. Note that the western-most
point of $C$ is also the western-most point of the western-most corner
of $C$.
Recall that the Gaussian map of a polytope is a subdivision of
$S^{d-1}$ into a spherical cell complex. For any cell $C$ of
$\mathcal{G}(P)$ that does not intersect $U$, the western-most point
of $C$ is a node incident to $C$, which is unique because otherwise
there would be a great circle containing an edge of $\mathcal{G}(P)$
and intersecting $U$, which contradicts the way we chose $U$. As a
consequence, there is one unique western-most corner for each cell of
a Gaussian map that does not intersect $U$.
We now prove that a western-most corner of some cell of
$\mathcal{G}(P)$ is also a western-most corner of a cell of the
Gaussian map of any partial sum $\mathcal{G}(P_S)$ if and only if its
western-most point is a node of $\mathcal{G}(P_S)$. We call a point
$p$ in the closure of a subset $C$ of $S^{d-1}$ a \emph{local optimum}
of $C$ if $p$ is the western-most point of the intersection of $C$
with some open subset of $S^{d-1}$ containing $p$.
\begin{lemma}\label{lem:local}
A point $p$ is a local optimum of a cell $C$ of a Gaussian map $G$
if and only if it is a western-most point of $C$.
\end{lemma}
The proof is in Appendix~\ref{app:ls}. For the next lemma, recall that
$\varepsilon>0$ is smaller than the distance between any two
non-incident cells in $\mathcal{G}(P)$.
\begin{lemma}\label{lem:var}
Let $F$ be a facet of $P$, with its normal region $\mathcal{N}(F;P)$
a node of $\mathcal{G}(P)$. Let $p$ be a point of $S^{d-1}$ at
distance less than $\varepsilon$ of $\mathcal{N}(F;P)$. For any
partial sum $P_S$ with $I_F\subseteq S$, the dimensions of the cells
containing $p$ in $\mathcal{G}(P)$ and $\mathcal{G}(P_S)$ are the
same.
\begin{proof}
In the Gaussian map $\mathcal{G}(P)$, the subset of $S^{d-1}$ at a
distance less than $\varepsilon$ of $\mathcal{N}(F;P)$ intersects
only the normal regions of subfaces of $F$. Therefore, for any point
$p$ in that subset, $\mathcal{S}(P;p)$ is a subface $G$ of
$F$. Recall that for any subface $G$ of a facet $F$, $I_G\subseteq
I_F$. So for any partial sum $P_S$ such that $I_F\subseteq S$,
$I_G\subseteq S$, which means that not only $P_S$ has a facet with
the same normal region as $F$, but $\mathcal{S}(P_S;p)$ is a subface
of that facet with the same dimension as $G$, and $p$ is in a cell
of the same dimension in $\mathcal{G}(P_S)$ as in $\mathcal{G}(P)$.
\end{proof}
\end{lemma}
We finally have the tools to prove:
\begin{lemma}
Let $W$ be a western-most corner of a cell $C$ in $\mathcal{G}(P)$,
with $\mathcal{N}(F;P)$ the western-most point of $C$. Then
$W$ is a western-most corner of some cell of the same dimension
in the Gaussian map of a partial sum $\mathcal{G}(P_S)$ if and only
if $I_F\subseteq S$.
\begin{proof}
First, if $I_F\not\subseteq S$, then $\mathcal{N}(F;P)$ is not a
node of $\mathcal{G}(P_S)$, and so $W$ cannot be a western-most
corner. Suppose $I_F\subseteq S$; then $\mathcal{N}(F;P)$ is a node
of $\mathcal{G}(P_S)$. Furthermore, by Lemma~\ref{lem:var}, the
points in $W$ are in a cell of the same dimension in
$\mathcal{G}(P_S)$ as in $\mathcal{G}(P)$, and the points in the
closure of $W$ are in a cell of the same dimension in
$\mathcal{G}(P_S)$ as in $\mathcal{G}(P)$. As a consequence, since
$\mathcal{N}(F;P)$ is the western-most point of $W$ in
$\mathcal{G}(P)$, it is also the western-most point of $W$ in
$\mathcal{G}(P_S)$. But $W$ is the intersection, of the cell it is in,
with an open subset, and so $\mathcal{N}(F;P)$ is a local optimum of
the cell that contains $W$ in $\mathcal{G}(P_S)$. By
Lemma~\ref{lem:local}, it is also the western-most point of the cell
that contains $W$ in $\mathcal{G}(P_S)$, and so $W$ is the
western-most corner of that cell in $\mathcal{G}(P_S)$.
\end{proof}
\end{lemma}
This is the most important lemma. It is the ultimate goal of the
definitions in this section, which is to have a witness of the
existence of a cell, a witness whose presence in the Gaussian maps of
partial sums depends on a simple rule.
However, according to the definitions so far, cells intersecting $U$
do not have a western-most corner. In any cell that intersects $U$, it
is possible to turn around $U$, always going west, much like the way
it is possible to turn around a pole on $S^2$. To deal with this
problem, we consider the restriction of the Gaussian map to $U$. Let
us denote as $S_U$ the intersection of $S^{d-1}$ with $U$. $S_U$ is a
sphere equivalent to $S^{d-3}$, and the restriction of a spherical
cell complex on $S^{d-1}$ to $S_U$ also defines a spherical cell
complex on $S_U$. In fact, the restriction to $S_U$ of the Gaussian
map on $S^{d-1}$ of a $d$-dimensional polytope is the Gaussian map on
$S_U$ of the orthogonal projection of the polytope onto $U$
Since $S_U$ is a sphere equivalent to $S^{d-3}$, we can define west on
$S_U$ as we have done for $S^{d-1}$ (See Figure~\ref{fig:s3}). For
any cell of $\mathcal{G}(P)$ that intersects $S_U$, we define its
western-most corner as the western-most corner of its intersection
with $S_U$ in the restriction of $\mathcal{G}(P)$ to $S_U$. If $d>5$,
this again does not define a westernmost point for every cell, because
west is not defined on the intersection of $S_U$ with a subspace of
dimension $d-4$; so we restrict the Gaussian map to that subspace, and
start again recursively.
\begin{figure}
\begin{center}
\psset{unit=0.9cm,shortput=nab,linewidth=0.5pt,arrowsize=2pt 3}
\pspicture(-5,-2.5)(5,2.5)%
\definecolor{verylightblue}{rgb}{0.8,0.8,1}
\definecolor{lightblue}{rgb}{0.6,0.6,1}
\psellipticarc(2,1.2)(0.6,1.6){-30}{50}
\psellipticarc(-2.4,1.5)(0.8,1.2){140}{180}
\psset{fillstyle=solid,fillcolor=verylightblue,linecolor=blue}
\psline(0.5,2.5)(1.4,2.4)(2.3,2.2)(4,1.3)(4.5,0.4)(4.4,-0.3)(4.2,-0.7)(3.5,-1.4)(2.7,-2.0)(2.1,-2.3)(1.6,-2.4)(0.6,-2.5)(-0.5,-2.5)(-1.4,-2.4)(-2.3,-2.2)(-4,-1.3)(-4.5,-0.4)(-4.4,0.3)(-4.2,0.7)(-3.5,1.4)(-2.7,2.0)(-2.1,2.3)(-1.6,2.4)(-0.6,2.5)(0.5,2.5)(1.0,1.5)(-0.1,1.6)(-1,1.5)(-2.2,1.3)(-3.3,0.4)(-3.5,0)(-3.0,-1)(-1.9,-1.4)(-1.0,-1.5)(0.1,-1.6)(1,-1.5)(2.2,-1.3)(3.3,-0.4)(3.5,0)(3.0,1)(1.9,1.4)(-0.1,1.6)
\psset{fillstyle=solid,fillcolor=lightblue}
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){61}{80}
\pspolygon(0.5,2.5)(1.0,1.5)(0.6,1.8)
\pspolygon(1.9,1.4)(2.3,2.2)(1.8,2.0)
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){35}{61}
\psellipticarc[fillstyle=none,linecolor=black](0,0)(4.2,2.1){35}{50}
\pspolygon(3.0,1)(3.5,1.4)(4,1.3)
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){-2}{35}
\psellipticarc[fillstyle=none,linecolor=black]{<-}(0,0)(4.2,2.1){10}{35}
\pspolygon(4.4,-0.3)(3.5,0)(4.0,0.1)
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){-35}{-2}
\pspolygon(3.3,-0.4)(3.5,-1.4)(2.9,-1.2)
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){-59}{-35}
\pspolygon(2.1,-2.3)(2.2,-1.3)(1.9,-1.8)
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){-79}{-59}
\pspolygon(1,-1.5)(0.6,-2.5)(0.7,-1.9)
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){-102}{-79}
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){80}{101}
\pspolygon(-1,1.5)(-0.6,2.5)(-0.7,1.9)
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){101}{122}
\pspolygon(-2.1,2.3)(-2.2,1.3)(-2.0,1.8)
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){122}{145}
\pspolygon(-3.3,0.4)(-3.5,1.4)(-2.9,1.2)
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){145}{178}
\pspolygon(-4.4,0.3)(-3.5,0)(-4.0,-0.1)
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){178}{215}
\psellipticarc[fillstyle=none,linecolor=black]{<-}(0,0)(4.2,2.1){190}{215}
\pspolygon(-3.0,-1)(-3.5,-1.4)(-4,-1.3)
\psellipticarc[fillstyle=none,linecolor=black](0,0)(4.2,2.1){215}{230}
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){-145}{-122}
\pspolygon(-0.5,-2.5)(-1.0,-1.5)(-0.9,-2.2)
\pspolygon(-1.9,-1.4)(-2.3,-2.2)(-2.2,-1.7)
\psellipticarc[linewidth=1.5pt,fillstyle=none,linecolor=black](0,0)(4,2){-122}{-102.5}
\psset{fillstyle=none,fillcolor=white,linecolor=black}
\psellipticarc{->}(2,1.2)(0.6,1.6){50}{150}
\psellipticarc{->}(2.3,-1.7)(0.5,1.2){-50}{140}
\psellipticarc{<-}(-2.4,1.5)(0.8,1.2){-30}{140}
\psline(-1,-1)(-1.5,-1.75)
\rput*(-1,-1){$S^3\cap U$}
\psline(4.4,2)(3.1,1.5)
\rput*[l](4.4,2){west in $S^3\cap U$}
\psline(0,0)(-1.5,0.95)
\psline(0,0)(1.85,-0.7)
\psline(0,0)(2.35,0.6)
\rput*(0,0){west in $S^3\setminus U$}
\endpspicture%
\end{center}
\caption{Representation of a map in $S^3$ by stereographic projection
in Euclidean space. West defined in $S^3\setminus U$ is turning
around the intersection of $S^3$ with a $2$-dimensional subspace
$U$. Cells which intersect the subspace have their western-most
corner defined by a different direction west defined on the
intersection, which is equivalent to $S^1$.}\label{fig:s3}
\end{figure}
We present now the complete construction. We have chosen a subspace
$U$ of dimension $d-2$ such that its intersection with any
two-dimensional plane containing an edge of $\mathcal{G}(P)$ is just
the origin. Let us write $U^{d-2}=U$, and denote as $G^{d-2}$ the
restriction of the Gaussian map $\mathcal{G}(P)$ to
$U^{d-2}$. $G^{d-2}$ is a spherical cell complex on $S^{d-3}$. Then,
for any $i$ larger than $2$, we define from $U^i$ and $G^i$ a subspace
$U^{i-2}$, which is a subspace of $U^i$, such that the intersection of
$U^{i-2}$ with any two-dimensional plane containing an edge of $G^i$
is just the origin. We then define $G^{i-2}$ as the restriction of
$G^i$ to $U^{i-2}$, which is a spherical cell complex on
$S^{i-3}$. This defines a sequence of subspaces $U^{d-2}\supset
U^{d-4}\supset\cdots$ and a sequence of spherical cell complexes
$G^{d-2}\supset G^{d-4}\supset\cdots$. If $d$ is even, the sequences
end with $G^2$, which is a spherical cell complex on $S^1$, and $U^0$
is just the origin and does not intersect $S^1$. If $d$ is odd, they
end with $G^3$, a spherical cell complex on $S^2$, and $U^1$ is a
one-dimensional subspace whose intersection with $S^2$ defines two
antipodal points that we called north and south pole in
Section~\ref{sec:3d}.
For each $G^i$, spherical cell complex of $S^{i-1}$, we can define a
direction west for every point of $S^{i-1}$ that is not on $U^{i-2}$,
as we have done for $\mathcal{G}(P)$ on $S^{d-1}$. Then for any cell
$C$ of $\mathcal{G}(P)$, let $i$ be the smallest number such that the
intersection of $C$ with $U^i$ is nonempty. We then define the
western-most point of $C$ to be the western-most point of $C\cap U^i$,
cell of $G^i$ on the sphere $S^{i-1}$. The western-most corner of $C$
is also the western-most corner of $C\cap U^i$.
Note that if a cell $C$ is of dimension $d-k-1$, i.e. it is the normal
region in $\mathcal{G}(P)$ of a $k$-dimensional face, it does not
intersect $U^i$ for any $i\geq k$, because $U^i$ was chosen so as not
to intersect edges and nodes of $G^{i+2}$, which are restrictions to
$G^{i+2}$ of cells of dimension $d-i-1$ and $d-i-2$ in
$\mathcal{G}(P)$. For instance, if $d$ is odd, only normal region of
vertices may intersect with $U^1$. Since $U^1$ only intersects
$S^{d-1}$ in two antipodal points, there are exactly two cells of
dimension $d-1$ in any Gaussian map that intersect $U^1$. These are
the only two cells that do not have a western-most corner. When $d$ is
even, west is defined on every point of $G^2$, spherical cell complex
on $S^1$, and so every cell of a Gaussian map has a western-most
corner.
We have now defined a western-most corner for every cell of Gaussian
maps, with the exception, if $d$ is odd, of the two cells that contain
a pole. As before for cells that do not intersect $U$, the
western-most corner of a cell of $\mathcal{G}(P)$ is also a
western-most corner of a cell of the same dimension in the Gaussian
map of a partial sum $\mathcal{G}(P_S)$ if and only if $S$ contains
the support of its western-most point, or rather the support of the
cell whose restriction is its western-most point. The cardinality of
the support is always less than $d$.
Now that we have a complete definition of western-most corners, all
that remains is to count them. The support of any face of $P$ has
cardinality less than $d$, so all western-most corners of cells of
$\mathcal{G}(P)$ can be found in the Gaussian map of partial sums of
at most $d-1$ summands. It is not difficult to see that for any $j\geq
|I_F|$, there are ${{r-|I_F|} \choose {j-|I_F|}}$ subsets of
$\{1,\ldots,r\}$ of cardinality $j$ that contain $I_F$.
The formula of the main theorem was found by observing low-dimensional
cases. It is based on the following combinatorial equivalence:
\begin{lemma}\label{lem:for}
For any $1\leq s<d\leq r$,
$$
\sum_{j=1}^{d-1}(-1)^{d-1-j} {{r-1-j} \choose {d-1-j}} {{r-s} \choose
{j-s}} = 1.
$$
\end{lemma}
The proof is in Appendix~\ref{app:for}.
By this Lemma, if we count all the western-most corners in partial
sums of $j$ polytopes, multiply by $(-1)^{d-1-j}{r-1-j \choose
d-1-j}$, and sum over $j$, we end up counting exactly once each
western-most corner, no matter what is the cardinality of the relevant
support. Therefore, if $w_k(.)$ is the number of western-most corners
of $k$-dimensional cells in a Gaussian map,
$$
w_k(\mathcal{G}(P)) = \sum_{j=1}^{d-1}(-1)^{d-1-j} {{r-1-j} \choose
{d-1-j}} \sum_{S\in \mathcal{C}_j^r}w_k(\mathcal{G}(P_S)),\quad k=0,\ldots,d-1
$$
where $\mathcal{C}_j^r$ is the family of subsets of $\{1,\ldots,r\}$
of cardinality $j$. Since there is one western-most corner of a
$k$-dimensional cell for each $d-1-k$ face of the underlying polytope,
this proves Theorem~\ref{thm:main} for any $d$ and $k$. The only
exception is that if $d$ is odd, any Gaussian map has two regions of
dimension $d-1$ that contain the poles, and that have no western-most
corner, and so in that case, $w_{d-1}(\mathcal{G}(\mathcal{P}))=
f_0(\mathcal{P})-2$. This gives the special case of the theorem for
$d$ odd and $k=0$.
In order to prove Corollary~\ref{cor}, it is enough to point out that
each western-most corner of cells of $\mathcal{G}(P)$ can be found at
least once (and often a lot more) in the Gaussian map of partial sums
of $d-1$ summands. And so, the last term of the sum in
Theorem~\ref{thm:main} is an upper bound on the number of faces.
\section{Maximum number of vertices}\label{sec:max}
Using Corollary~\ref{cor}, we show bounds on the number of vertices of
Minkowski sums. The trivial bound tells us that if $r<d$, then
$f_0(P_1+\cdots+P_r)\leq \prod_{i=1}^rf_0(P_i)$. Consequently, if
$r\geq d$, we get by Corollary~\ref{cor} that
$$
f_0(P_1+\cdots+P_r) \leq \sum_{S\in \mathcal{C}_{d-1}^r}\prod_{i\in S}f_0(P_i).
$$
This can be seen as enumerating all possible combinations of $d-1$
vertices chosen each from a different summand. This is necessarily
lower than all possible combinations of $d-1$ vertices from the
summands. If the summands have $n$ vertices in total, this upper bound
is ${n \choose d-1}$, which is in $O(n^{d-1})$. If each summands has
$n$ vertices, then we have:
$$
f_0(P_1+\cdots+P_r) \leq \sum_{S\in
\mathcal{C}_{d-1}^r}n^{d-1}={r \choose d-1}n^{d-1},
$$
which is in $O(r^{d-1}n^{d-1})$. The previous known bound was in
$O(r^{d-1}n^{2(d-1)})$~\cite{Gritzmann93}.
Note that a construction from~\cite{Fukuda07} allows us to choose
$d-1$ polytopes of $n$ vertices each such that the sum has $n^{d-1}$
vertices. It is easy to adapt this construction to choose $r$
polytopes such that any partial sum of $d-1$ summands has this many
vertices, which by Theorem~\ref{thm:main} means that the total sum has
exactly (for $d$ even) $\sum_{j=1}^{d-1}(-1)^{d-1-j}{r-1-j\choose
d-1-j}{r\choose j}n^j$ vertices.
Unfortunately, except in three dimensions, the maximum number of
facets of a Minkowski sum of polytopes remains open, even for two
summands in four dimensions, so we cannot write an upper bound for facets.
We can however tell that if the number of facets in the sum of $d-1$ polytopes
is in $O(p(n))$, their number in the sum of $r\geq d$ polytopes is in
$O(r^{d-1}p(n))$. Finding $p(n)$ should be the object of further research.
\section{Summary}\label{sec:con}
We have extended the intuitive concept of west from three dimensions
to higher dimensions. Thanks to the properties of the concept, we were
able to prove a relation on the number of vertices in sums of many
polytopes, and show that this number has a comparatively low order
of complexity. For faces of higher dimensions, the result also shows
that the complexity of Minkowski sums of many polytopes is not much
more complex than that of $d-1$ summands.
|
\section{Introduction}
For an integer $n \ge 1$, we denote the ring of integers modulo $n$ by $\Z_n$, and the group of multiplicative units modulo $n$ by $\Z_n\units$.
A well-studied family of graphs are the \emph{unitary Cayley graphs} on $\Z_n$, which are defined by $X_n = \Cay(\Z_n, \Z_n\units)$.
These form the basis of the subject of graph representations~\cite{EE89}, and are also studied as objects of independent interest: see for example~\cite{DG95, BG04, KS07, RV09}.
We consider a subgraph $G_n \le X_n$ of the unitary Cayley graphs, defined as follows.
Let $Q_n = \ens{u^2 \,\big|\, u \in \Z_n\units}$ be the group of \emph{quadratic units} modulo $n$ (quadratic residues which are also multiplicative units), and $T_n = \pm Q_n$.
We then define $G_n = \Cay(\Z_n, T_n)$, in which two vertices given by $a,b \in \Z_n$ are adjacent if and only if their difference is a quadratic unit in $\Z_n$, \ie\ if $a - b \in \ens{\pm u^2 \,\big|\, u \in \Z_n\units}$.
In the case where $n \equiv 1 \pmod{4}$ and is prime, $G_n$ coincides with the Paley graph on $n$ vertices: thus the graphs $G_n$ are a circulant generalization of these graphs for arbitrary $n$.
We refer to $G_n$ as the \emph{(undirected) quadratic unitary Cayley graph} on $\Z_n$.
We present some structural properties of quadratic unitary Cayley graphs $G_n$.
In particular, we characterize its decompositions into tensor products over relatively prime factors of $n$, and categorize the graphs $G_n$ in terms of their diameters.
From these results, we obtain a corollary regarding the decomposition of symplectic matrices $S \in \Sp_{2m}(\Z_n)$ in terms of symplectic row-operations, consisting of symplectic scalar multiplications, symplectic row-swaps, and symplectic row-additions/subtractions.
We also characterize the conditions under which quadratic unitary graphs are perfect, by examining special cases of quadratic unitary graphs which are self-complementary.
\begin{notation}
Throughout the following, $n = p_1^{m_1} p_2^{m_2} \cdots p_t^{m_t}$ is a decomposition of $n$ into powers of distinct primes, and $\sigma: \Z_n \to* \Z_{\smash{p_1}^{\!m_1}} \oplus \cdots \oplus \Z_{\smash{p_t}^{\!m_t}}$ is the isomorphism of rings which is induced by the Chinese Remainder theorem.
(We refer to similar isomorphisms $\rho: \Z_n \to \Z_M \oplus \Z_N$ for coprime $M$ and $N$ as \emph{natural} isomorphisms.)
We sometimes describe the properties of $G_n$ in terms of the directed Cayley graph $\Gamma_n = \Cay(\Z_n, Q_n)$, whose arcs $a \to* b$ correspond to addition (but not subtraction) of a quadratic unit to a modulus $a \in \Z_n$; we may refer to this as the \emph{directed quadratic unitary Cayley graph}.
\end{notation}
\section{Tensor product structure}
By the isomorphism $\Z_n\units \cong \Z\units_{\smash{p_1}^{\!m_1}} \oplus \cdots \oplus \Z_{\smash{p_t}^{\!m_t}}\units$ induced by $\sigma$, unitary Cayley graphs $X_n$ may be decomposed as \emph{tensor products} $X_n \cong X_{\smash{p_1}^{\!m_1}} \ox \cdots \ox X_{\smash{p_t}^{\!m_t}}$ of smaller unitary Cayley graphs (also called direct products~\cite{RV09} or Kronecker products~\cite{W62}, among other terms):
\begin{definition}
\label{def:tensorProduct}
The \emph{tensor product} $A \ox B$ of two \mbox{(di-)graphs} $A$ and $B$ is the \mbox{(di-)graph} with vertex-set $V(A) \x V(B)$, where $((u,u'), (v,v')) \in E(A \ox B)$ if and only if $((u,v), (u',v')) \in E(A) \x E(B)$.\footnote{%
We write $A_1 \ox (A_2 \ox A_3) = (A_1 \ox A_2) \ox A_3 = A_1 \ox A_2 \ox A_3$, and so on for higher-order tensor products,
similarly to the convention for Cartesian products of sets.
}
\end{definition}
\noindent
Corollary~3.3 of \cite{RV09} gives an explicit proof that $X_n \cong X_{\smash{p_1}^{\!m_1}} \ox \cdots \ox X_{\smash{p_t}^{\!m_t}}$; a similar approach may be used to decompose any \mbox{(di-)graph} $\Cay(R,M)$ for rings $R = R_1 \oplus \cdots \oplus R_t$ and multiplicative monoids $M_1 = M_1 \oplus \cdots \oplus M_t$ where $M_j \subset R_j$.
For instance, as $Q_n \cong Q_{\smash{p_1}^{\!m_1}} \oplus \cdots \oplus Q_{\smash{p_t}^{\!m_t}}$\,, it follows that $\Gamma_n \cong \Gamma_{\smash{p_1}^{\!m_1}} \ox \cdots \ox \Gamma_{\smash{p_t}^{\!m_t}}$ as well.
It is reasonable to suppose that the graphs $G_n$ will also exhibit tensor product structure; however, they do not always decompose over the prime power factors of $n$ as do $X_n$ and $\Gamma_n$.
This is because $T_n$ may fail to decompose as a direct product of groups over the prime-power factors $p_j^{m_j}$.
By definition, for each $j$, we either have $T_{\smash{p_j}^{\!m_j}} = Q_{\smash{p_j}^{\!m_j}}$ or $T_{\smash{p_j}^{\!m_j}} \cong Q_{\smash{p_j}^{\!m_j}} \oplus \gen{-1}$; when $Q_{\smash{p_j}^{\!m_j}} < T_{\smash{p_j}^{\!m_j}}$ for multiple $p_j$, one cannot decompose $T_n$ over the prime-power factors of $n$.
We may generalize this observation as follows:
\begin{theorem}
\label{thm:tensorProductFactorizable}
For coprime integers $M, N \ge 1$, we have $G_M \ox G_N \cong G_{MN}$ if and only if either $-1 \in Q_M$ or $-1 \in Q_N$.
\end{theorem}
\begin{proof}
We have $G_M \ox G_N \cong G_{MN}$ if and only if $T_M \oplus T_N \cong T_{MN}$.
Let $\rho: \Z_{MN} \to \Z_M \oplus \Z_N$ be the natural isomorphism: this induces an isomorphism $Q_{MN} \cong Q_M \oplus Q_N$, and will also induce an isomorphism $T_{MN} \cong T_M \oplus T_N$ if the two groups are indeed isomorphic.
Clearly, $\sigma(T_{MN}) \le T_M \oplus T_N$; we consider the opposite inclusion.
If $-1 \notin Q_M$ and $-1 \notin Q_N$, we have $(-1,\,1),(1,-1) \notin Q_M \oplus Q_N$; as both tuples are elements of $T_M \oplus T_N$, but neither of them are elements of $\pm(Q_M \oplus Q_N) = \sigma(\pm Q_{MN}) = \sigma(T_{MN})$, it follows that $T_{MN}$ and $T_M \oplus T_N$ are not isomorphic in this case.
Conversely, consider $u \in \Z_n\units$ arbitrary, and let $(u_M, u_N) = \rho(u)$.
If $-1 \in Q_M$, let $i \in \Z_M$ such that $i^2 = -1$: for any $s_M,s_N \in \ens{0,1}$, we then have
\begin{align}
\Big((-1)^{s_M} u_M^2,\, (-1)^{s_N} u_N^2\Big)
\;=&\;\,
(-1)^{s_N} \Big((-1)^{s_M - s_N} u_M^2,\, u_N^2\Big)
\notag\\=&\;\,
(-1)^{s_N} \Big( \big[i^{(s_M - s_N)} u_M\big]^2,\, u_N^2\Big)
\,.
\end{align}
Thus $T_M \oplus T_N \le \sigma(T_{MN})$; and similarly if $-1 \in Q_N$.
\end{proof}
\paragraph{Remark.} The above result is similar to~\cite[Theorem~8]{KS09}, which uses a ``partial transpose'' criterion to indicate when a graph may be regarded as a symmetric difference of tensor products of graphs on $M$ and $N$ vertices; the presence of $-1$ in either $Q_M$ or $Q_N$ is equivalent to $G_{MN}$ being invariant under partial transposes (w.{}r.{}t.{} to the tensor decomposition induced by $\rho$).
\begin{corollary}
\label{cor:decomposeGnTensors}
For $n \ge 1$, let $n = p_1^{m_1} \cdots p_\tau^{m_\tau} N$ be a factorization of $n$ such that $p_j \equiv 1 \pmod{4}$ for each $1 \le j \le \tau$, and $N$ has no such prime factors.
Then $G_n \cong G_{\smash{p_1}^{\!m_1}} \ox \cdots \ox G_{\smash{p_\tau}^{\!m_\tau}} \ox G_N$\,.
\end{corollary}
\begin{proof}
For $p_j$ odd, $\Z_{\smash{p_j}^{\!m_j}}\units$ is a cyclic group~\cite{Gauss} of order $(p_j - 1)p_j^{m_j - 1}$ in which $-1$ is the unique element of order two: then $-1$ is a quadratic residue modulo $p_j^{m_j}$ if and only if $p_j \equiv 1 \pmod{4}$.
As this holds for all $1 \le j \le \tau$, repeated application of Theorem~\ref{thm:tensorProductFactorizable} yields the decomposition above.
\end{proof}
\begin{corollary}
For $n \ge 1$, we have $G_n \cong G_{\smash{p_1}^{\!m_1}} \ox \cdots \ox G_{\smash{p_t}^{\!m_t}}$ if and only if either $n$ has at most one prime factor $p_j \not\equiv 1 \pmod{4}$, or $n$ has two such factors and $n \equiv 2 \pmod{4}$.
\end{corollary}
\begin{proof}
Suppose that $G_n$ decomposes as above.
Let $N$ be the largest factor of $n$ which does not have prime factors $p \equiv 1 \pmod{4}$: we continue from the proof of Corollary~\ref{cor:decomposeGnTensors}.
By Theorem~\ref{thm:tensorProductFactorizable}, $G_N$ itself decomposes as a tensor factor over its prime power factors $p_{\tau+1}^{m_{\tau+1}}, \ldots, p_t^{m_t}$ if and only if there is at most one such prime $p_j$ such that $-1 \notin Q_{\smash{p_j}^{\!m_j}}$.
However, by construction, all odd prime factors $p_j$ of $N$ satisfy $p_j \equiv 3 \pmod{4}$, in which case $-1 \notin Q_{\smash{p_j}^{\!m_j}}$ for any of them.
Furthermore, for $m \ge 2$, we have $r \in Q_{2^m}$ only if $r \equiv 1 \pmod{4}$; then $-1 \in Q_{2^m}$ if and only if $2^m = 2$.
Thus, if $G_n \cong G_{\smash{p_1}^{\!m_1}} \ox \cdots \ox G_{\smash{p_t}^{\!m_t}}$, it follows either that $N = p^m$ for some prime $p \equiv 3 \pmod{4}$, in which case the decomposition of Corollary~\ref{cor:decomposeGnTensors} is the desired decomposition, or $N = 2 p^m$ for some prime $p \equiv 3 \pmod{4}$, in which case $n \equiv 2 \pmod{4}$.
The converse follows easily from Corollary~\ref{cor:decomposeGnTensors} and Theorem~\ref{thm:tensorProductFactorizable}.
\end{proof}
We finish our discussion of tensor products with an observation for prime powers.
Let $\mathring{K}_M$ denote the complete pseudograph on $M$ vertices (\ie\ an $M$-clique with loops):
\begin{lemma}
\label{lemma:primePowerTensorDecomp}
For $m \ge 3$, we have $G_{2^m\!} \cong G_8 \ox \mathring K_{2^{m-3}\!\!}$ \,and\, $\Gamma_{2^m\!} \cong \Gamma_8 \ox \mathring K_{2^{m-3}}$; for $p$ an odd prime and $m \ge 1$, we have $G_{p^m\!} \cong G_p \ox \mathring K_{p^{m-1}\!\!}$ \,and\, $\Gamma_{p^m\!} \cong \Gamma_p \ox \mathring K_{p^{m-1}}$.
\end{lemma}
\begin{proof}
We prove the results for $\Gamma_{p^m}$; the results for $G_{p^m}$ are similar.
\begin{itemize}
\item
Let $n = 2^m$ for $m \ge 3$.
We have $q \in Q_n$ if and only if $q \equiv 1 \pmod{8}$.
Let $\tau: \Z_{2^m} \to* \Z_8 \x \Z_{2^{m-3}}$ (not a ring homomorphism) be defined by $\tau(r) = (r', k')$ such that $r = 8k' + r'$ for $r' \in \ens{0, \ldots, 7}$.
Then, we have $a - b \in Q_n$ if and only if $\tau(a - b) \in \ens{1} \x \Z_{2^{m-3}}$, so that $\tau$ induces a homomorphism $\Gamma_n \cong \Gamma_8 \ox \mathring K_{2^{m-3}}$.
\item
Similarly, for $n = p^m$ for $p$ an odd prime and $m \ge 1$, we have $q = pk' + q' \in Q_n$ (for $q' \in \ens{0, \ldots, p-1}$, which we we identify with $\Z_p$) if and only if $q' \in Q_p$.
If $\tau: \Z_{2^m} \to* \Z_p \x \Z_{p^{m-1}}$ is defined by $\tau(q) = (q', k')$, we then have $a - b \in Q_n$ if and only if $\tau(a - b) \in Q_p \x \Z_{p^{m-1}}$.
Thus, $\tau$ induces a homomorphism $\Gamma_n \cong \Gamma_p \ox \mathring K_{p^{m-1}}$. \qedhere
\end{itemize}
\end{proof}
\noindent
Together with Corollary~\ref{cor:decomposeGnTensors}, and the fact that $\mathring K_{p^m}$ itself may be decomposed for any prime $p$ as an $m$-fold tensor product $\mathring K_p \ox \cdots \ox \mathring K_p$, the graph $G_n$ may be decomposed very finely whenever $n$ is dominated by prime-power factors $p^m$ for $p \equiv 1 \pmod{4}$.
\section{Induced paths and cycles of $G_n$}
\label{sec:pathsCycles}
Even when the graph $G_n$ does not itself decompose as a tensor product, we may fruitfully describe such properties as walks in the graphs $G_n$ in terms of correlated transitions in tensor-factor ``subsystems''.
This intuition will guide the analysis of this section in our characterization both of the diameters of the graphs $G_n$, and of the factors of $n$ for $G_n$ a perfect graph.
As $T_n$ is a multiplicative subgroup of $\Z_n\units$, we may easily show that the graphs $G_n$ are arc-transitive.
For any pair of edges $vw, v'w' \in E(G_n)$, the affine function $f(x) = (w'-v')(w-v)^{\text{--}1}(x-v) + v'$ is an automorphism of $G_n$ which maps $v \mapsto* v'$ and $w \mapsto* w'$.
Consequently $G_n$ is vertex-transitive as well, so that we may bound the diameter by bounding the distance of vertices $v \in V(G)$ from $0 \in V(G)$, and also restrict our attention to odd induced cycles (or \emph{odd holes}) which include $0$ in our analysis of perfect graphs.
Let $A_n$, $B_n$ be the adjacency graphs of the graph $G_n$ and the digraph $\Gamma_n$ respectively.
We then have $A_n = B_n = B_n\trans$ if and only if $-1$ is a quadratic residue modulo $n$, and $A_n = B_n + B_n\trans$ otherwise; in either case, we have $A_n \,\propto\, B_n + B_n\trans$.
As $B_n$ may be decomposed as a Kronecker product (corresponding to the tensor decomposition of $\Gamma_n$), this suggests an analysis of walks in $G_n$ in terms of ``synchronized walks'' in the rings $\Z_{\smash{p_j}^{\!m_j}}$ by adding or subtracting quadratic units, where one must add a quadratic unit in all rings simultaneously or subtract a quadratic unit in all rings simultaneously.
This will inform the analysis of properties such as the diameters and perfectness of the graphs $G_n$.
\subsection{Characterizing paths of length two for $n$ odd}
To facilitate the analysis of this section, we will be interested in enumerating paths of length two in $G_n$ between distinct vertices.
Because $A_n \,\propto\, B_n + B_n\trans$ for all $n$, we have
\begin{align}
A_n^2
\;\;\propto&\;\;
B_n^2
\,+\,
2 B_n B_n\trans
\,+\,
\big( B_n \trans \big)^2
\notag\\[0.5ex]\cong&\;\;
\sqparen{\bigotimes_{j = 1}^t B_{\smash{p_j}^{\!m_j}}^2}
\;+\;
2 \sqparen{\bigotimes_{j = 1}^t B_{\smash{p_j}^{\!m_j}} B_{\smash{p_j}^{\!m_j}}\trans}
\;+\;
\sqparen{\bigotimes_{j = 1}^t \big( B_{\smash{p_j}^{\!m_j}} \trans\big)^2}\,,
\end{align}
where congruence is up to a permutation of the standard basis.
Thus, we may characterize the paths of length two in $G_n$ between distinct vertices $r, s \in \Z_n$ in terms of the number of ways that we may represent $s - r$ in the form $\alpha^2 + \beta^2$, $\alpha^2 - \beta^2$, and $-\alpha^2 - \beta^2$ for some units $\alpha,\beta \in \Z_n\units$; and these we may characterize in terms of products over the number of representations in the special case where $n$ is a prime power.
\begin{definition}
For $n > 0$ and $r \in \Z_n$, we let $S_n(r)$ denote the number of solutions $(x,y) \in Q_n \x Q_n$ to the equation $r \,=\, x + y$; similarly, $D_n(r)$ denotes the number of solutions $(x,y) \in Q_n \x Q_n$ to the equation $r \;=\; x - y$.
\end{definition}
\noindent
Thus, when $-1 \in Q_n$ and $A_n = B_n = \frac{1}{2}(B_n + B_n\trans)$, the number of paths of length two from $0$ to $r \ne 0$ is $S_n(r)$; otherwise, if $-1 \notin Q_n$, the number of such paths is $S_n(r) + 2D_n(r) + S_n(-r)$.
Thus, the number of paths of length two from $0$ to $r$ reduces to avaluation of the functions $S_n$ and $D_n$.
We may evaluate these functions for $n$ a prime power, through a straightforward generalization of standard results on patterns of quadratic residues and non-residues to prime power moduli:
\begin{lemma}
\label{lemma:consecutiveResidues}
For $p$ a prime and $m \ge 1$, let $C^{++}_{p^m}$ (respectively $C^{--}_{p^m}$) denote the number of consecutive pairs of quadratic units (resp. consecutive pairs of non-quadratic units) modulo $p^m$, and $C^{+-}_{p^m}$ (respectively $C^{-+}_{p^m}$) denote the number of sequences of a quadratic unit followed by a non-quadratic unit (resp. a non-quadratic unit followed by a quadratic unit) modulo $p^m$.
For primes $p \equiv 1 \pmod{4}$, we have
\begin{subequations}
\begin{align}
C^{++}_{p^m}
=&\;
\frac{(p-5)p^{m-1}}{4} \;,
&
C^{+-}_p =\, C^{-+}_p =\, C^{--}_p
=&\;
\frac{(p-1)p^{m-1}}{4} \;;
\end{align}
otherwise, if $p \equiv 3 \pmod{4}$, we have
\begin{align}
C^{+-}_p
=&\;
\frac{(p+1)p^{m-1}}{4} \;,
&
C^{++}_p =\, C^{-+}_p =\, C^{--}_p
=&\;
\frac{(p-3)p^{m-1}}{4} \;.
\end{align}
\end{subequations}
\end{lemma}
\begin{proof}
As $r \in \Z$ is a quadratic residue, quadratic non-residue, and/or unit modulo $p^m$ if and only the same properties hold modulo $p$, the distribution of quadratic and non-quadratic units modulo $p^m$ is simply that of the integers modulo $p$, repeated $p^{m-1}$ times.
It then suffices to multiply the formulae given for $C_p^{++}$, $C_p^{+-}$, $C_p^{-+}$, $C_p^{--}$ (obtained by Aladov~\cite{Aladov1896}) by $p^{m-1}$.
\end{proof}
\begin{lemma}
\label{lemma:countingSumSquares}
Let $p$ be an odd prime, $m > 0$, and $r \in \Z_{p^m}$.
If $p \equiv 1 \pmod{4}$, we have
\begin{subequations}
\label{eqn:sumOfSquares}
\begin{align}
\mspace{-5mu}
S_{p^m}(r)
\,=\,
D_{p^m}(r)
\,=&\;
\begin{cases}
\frac{1}{4} (p-5) p^{m-1} \;, & \text{for $r$ a quadratic unit},\! \\[1ex]
\frac{1}{4} (p-1) p^{m-1} \;, & \text{for $r$ a non-quadratic unit},\! \\[1ex]
\frac{1}{2}(p-1)p^{m-1} \;, & \text{for $r$ a zero divisor};\!
\end{cases}
\intertext{for $p \equiv 3 \!\!\!\pmod{4}$, we instead have}
\mspace{-5mu}
S_{p^m}(r)
\,=&\;
\begin{cases}
\frac{1}{4} (p-3) p^{m-1} \;, & \text{for $r$ a quadratic unit},\! \\[1ex]
\frac{1}{4} (p+1) p^{m-1} \;, & \text{for $r$ a non-quadratic unit},\! \\[1ex]
0 \;, & \text{for $r$ a zero divisor};\!
\end{cases}
\\[1em]
\mspace{-5mu}
D_{p^m}(r)
\,=&\;
\begin{cases}
\frac{1}{4} (p-3) p^{m-1} \;, & \text{for $r$ a unit},\! \\[1ex]
\frac{1}{2} (p-1) p^{m-1} \;, & \text{for $r$ a zero divisor}.\!
\end{cases}
\end{align}
\end{subequations}
\end{lemma}
\begin{proof}
We proceed by cases, according to whether $r$ is a quadratic unit, non-quadratic unit, or zero modulo $p$:
\begin{itemize}
\item
Suppose $r \in Q_n$.
Each consecutive pair $q,q+1 \in Q_{p^m}$ yields a solution $(x,y) = (r(q+1),rq) \in Q_{p^m} \x Q_{p^m}$ to $x - y = r$; then we have $D_{p^m}(r) = C_{p^m}^{++}$.
Similarly, each such pair yields a solution $(x,y) = (rq(q+1)^{\text{--}1}, r(q+1)^{\text{--}1}) \in Q_{p^m} \x Q_{p^m}$ to $x + y = r$; then $S_{p^m}(r) = C_{p^m}^{++}$ as well.
\item
Suppose $r \in \Z_{p^m}\units \setminus Q_{p^m}$.
Each consecutive pair $s,s+1 \in \Z_{p^m}\units \setminus Q_{p^m}$ represents a solution in non-quadratic units to $x - y = 1$; these may then be used to obtain solutions $(rx, ry) \in Q_{p^m} \x Q_{p^m}$ to $rx - ry = r$, so that $D_{p^m}(r) = C_{p^m}^{--}$.
In the case that $p \equiv 1 \pmod{4}$, the negation of a quadratic unit is also a quadratic unit; in this case, we have the same number of solutions $(rx, -ry) \in Q_{p^m} \x Q_{p^m}$ to $rx + (-ry) = r$, so that $S_{p^m}(r) = C_p^{--}$ as well.
If instead $p \equiv 3 \pmod{4}$, we instead consider quadratic units $s \in Q_{p^m}$ such that $s+1$ is a non-quadratic unit.
Each such pair yields a solution $(x,y) = (r(s+1), -rs) \in Q_{p^m} \x Q_{p^m}$ to $x + y = r$; then we have a solution for each such pair $s,s+1$, so that $S_{p^m}(r) = C_{p^m}^{+-}$.
\item
Finally, suppose $r$ is a multiple of $p$.
The congruence $x + y \equiv 0 \pmod{p}$ is satisfiable for $(x,y) \in Q_{p^m} \x Q_{p^m}$ only if $-x$ is a quadratic unit modulo $p$ for some $x \in Q_{p^m}$, \ie\ if $p \equiv 1 \pmod{4}$.
If this is the case, then every $x \in Q_{p^m}$ contributes a solution $(x,y) = (x, r - x) \in Q_{p^m} \x Q_{p^m}$ to $x + y = r$; otherwise, in the case $p \equiv 3 \pmod{4}$, there are no solutions.
Similarly, regardless of the value of $p$, each quadratic unit $x \in Q_{p^m}$ contributes a solution $(x,y) = (x, x - r) \in Q_{p^m} \x Q_{p^m}$ to $x - y = r$.
Thus $D_{p^m}(r) = \frac{1}{2}(p-1)$ for all $p$; $S_{p^m}(r) = \frac{1}{2}(p-1)$ for $p \equiv 1 \pmod{4}$; and $S_{p^m}(r) = 0$ for $p \equiv 3 \pmod{4}$.
\qedhere
\end{itemize}
\end{proof}
\begin{corollary}
\label{cor:diameterOddPrimePower}
$\diam(G_{p^m}) \le 2$ for $p$ an odd prime and $m > 0$; this inequality is strict if and only if $p \equiv 3 \pmod 4$ and $m = 1$.
\end{corollary}
\begin{proof}
Clearly for $p \equiv 1 \pmod{4}$ we have $\diam(G_{p^m}) = 2$; suppose then that $p \equiv 3 \pmod 4$.
We may form any zero divisor $s = pk$ as a difference of quadratic units $x \in Q_{p^m}$ and $x - pk \in Q_{p^m}$, so that $\diam(G_{p^m}) \le 2$.
We have $\diam(G_{p^m}) = 1$ only if $0$ is the only zero divisor of $\Z_{p^m}$; this implies that $m = 1$, in which case $T_{p^m} = \Z_p\units$, so that the converse also holds.
\end{proof}
In Lemma~\ref{lemma:countingSumSquares}, $n = 3^m$ and $n = 5^m$ are cases for which there do not exist paths of length two from zero to any quadratic unit.
This does not affect the diameters of the graphs $G_{3^m}$ or $G_{5^m}$ for $m > 0$; however, using the following Lemma, we shall see that this deficiency affects the diameters of $G_n$ for any other $n$ a multiple of either $3$ or $5$.
\begin{lemma}
\label{lemma:oddPathsLength2}
For $n > 0$ odd and $r \in \Z_n$, we have $S_n(r) = 0$ if and only if at least one of the following conditions hold:
\begin{romanum}
\item
$n$ is a multiple of $3$, and $r \not\equiv 2 \pmod{3}$;
\item
$n$ is a multiple of $5$, and $r \equiv \pm 1 \pmod{5}$; or
\item
$n$ has a prime factor $p_j \equiv 3 \pmod{4}$ such that $r \in p_j \Z_n$.
\end{romanum}
Similarly, we have $D_n(r) = 0$ if and only if at least one of the following conditions hold:
\begin{romanum}
\item
$n$ is a multiple of $3$, and $r \not\equiv 0 \pmod{3}$; or
\item
$n$ is a multiple of $5$, and $r \equiv \pm 1 \pmod{5}$.
\end{romanum}
\end{lemma}
\begin{proof}
For $r \in \Z_n$ arbitrary, let $(r_1, r_2, \ldots, r_t) = \sigma(r)$.
By the decompositions $B^2_n \,\cong\, B^2_{\smash{p_1}^{\!m_1}} \ox \cdots \ox B^2_{\smash{p_t}^{\!m_t}}$ and $B_n B_n\trans \,\cong\, B_{\smash{p_1}^{\!m_1}}B_{\smash{p_1}^{\!m_1}}\trans \ox \cdots \ox B_{\smash{p_t}^{\!m_t}}B_{\smash{p_t}^{\!m_t}}\trans$, we may express $S_n(r)$ and $D_n(r)$ as products over the prime-power factors of $n$,
\begin{align}
S_n(r)
\;=&\;
\prod_{j = 1}^t S_{\smash{p_j}^{\!m_j}}(r_j) \;,
&
D_n(r)
\;=&\;
\prod_{j = 1}^t D_{\smash{p_j}^{\!m_j}}(r_j) \;.
\end{align}
These are zero if and only if there exist $1 \le j \le t$ such that $S_{\smash{p_j}^{\!m_j}}(r_j) = 0$ or $D_{\smash{p_j}^{\!m_j}}(r_j) = 0$, respectively.
By Lemma~\ref{lemma:countingSumSquares}, $S_{\smash{p_j}^{\!m_j}}(r_j) = 0$ if and only if either $r_j$ is a zero divisor of $\Z_{\smash{p_j}^{\!m_j}}$ for a prime factor $p_j \equiv 3 \pmod{4}$, or if $p_j \in \ens{3,5}$ and $r_j$ is a quadratic unit modulo $\Z_{\smash{p_j}^{\!m_j}}$\,; similarly, $D_{\smash{p_j}^{\!m_j}}(r_j) = 0$ if and only if $p_j = 3$ and $r_j$ is a unit modulo $3$, or $p_j = 5$ and $r_j$ is a quadratic unit modulo $5$.
\end{proof}
\subsection{Diameter of $G_n$ for odd $n$}
For odd integers $n$, characterizing the diameters of $G_n$ involves accounting for ``problematic'' prime factors of $n$ (those described in Lemma~\ref{lemma:oddPathsLength2}), which present obstacles to the construction of short paths between distinct vertices:
\begin{theorem}
\label{thm:oddDiameter}
Let $n > 1$ odd.
Let $\gamma_3(n) = 1$ if $n$ is a multiple of $3$, and $\gamma_3(n) = 0$ otherwise; $\delta_3(n) = 1$ if $n$ has prime factors $p_j \equiv 3 \pmod{4}$ for $p_j > 3$, and $\delta_3(n) = 0$ otherwise; and $\gamma_5(n) = 1$ if $n$ is a multiple of $5$, and $\gamma_5(n) = 0$ otherwise.
Then, we have
\begin{align*}
\diam(G_n)
\,=&\;
\begin{cases}
1 , & \!\!\text{if $n$ is prime and $n \equiv 3 \!\!\!\!\pmod{4}$}; \\[0.5ex]
2 , & \!\!\text{if $n$ is prime and $n \equiv 1 \!\!\!\!\pmod{4}$}; \\[0.5ex]
2 , & \!\!\text{if $\omega(n) = 1$ and $n$ is composite}; \\[0.5ex]
2 + \gamma_3(n) \delta_3(n) + \gamma_5(n) , & \!\!\text{if $\omega(n) > 1$}.
\end{cases}
\end{align*}
In particular, $\diam(G_n) \le 4$.
\end{theorem}
\begin{proof}
The diameters for $\omega(n) = 1$ are characterized by Corollary~\ref{cor:diameterOddPrimePower}: we thus restrict ourselves to the case $\omega(n) > 1$.
We have $\diam(G_n) \le 2$ if and only if either $S_n(r)$, $S_n(-r)$, or $D_n(r)$ is positive for all $r \in \Z_n \setminus T_n$.
By Lemma~\ref{lemma:oddPathsLength2}, $D_n(r) > 0$ for all $r \in \Z_n$ if $n$ is relatively prime to $15$; then $\diam(G_n) = 2$, and $r = u - u'$ for some $u, u' \in Q_n$ for any $r \in \Z_n$ if $\gamma_3 = \gamma_5 = 0$.
If $n$ is a multiple of $5$, however, we have $S_n(r) = S_n(-r) = D_n(r) = 0$ for any non-quadratic unit $r \equiv \pm 1 \pmod{5}$, of which there is at least one (as $n$ is not a power of $5$): thus $\diam(G_n) \ge 3$ if $\gamma_5(n) = 1$.
Suppose that $n$ is relatively prime to $5$, and is a multiple of $3$.
Again by Lemma~\ref{lemma:oddPathsLength2}, there are walks of length two from $0$ to $r$ if $r \equiv 0 \pmod{3}$, as we have $D_n(r) > 0$ in this case.
However, if $n$ has prime factors $p_j > 3$ such that $p_j \equiv 3 \pmod{4}$, there exist $r \in p_j \Z_n$ such that $r \not\equiv 0 \pmod{3}$, in which case we have $S_n(r) = S_n(-r) = D_n(r) = 0$.
Thus, if $\gamma_3(n) = \delta_3(n) = 1$, we have $\diam(G_n) \ge 3$.
Otherwise, if $\delta_3(n) = 0$, we have either $S_n(r) > 0$ in the case that $r \equiv 2 \pmod{3}$, or $S_n(-r) > 0$ in the case that $r \equiv 1 \pmod{3}$.
In this case, every vertex $r \ne 0$ is reachable by a path of length two, so that $\diam(G_n) = 2$ if $\gamma_3(n) = 1$ and $\delta_3(n) = \gamma_5(n) = 0$.
Finally, suppose that either $\gamma_5(n) = 1$ or $\gamma_3(n) = \delta_3(n) = 1$: from the analysis above, we have $\diam(G_n) \ge 3$.
For $r \in \Z_n$, let $(r_1, \ldots, r_n) = \sigma(r)$, where we arbitrarily label $p_3 = 3$ if $n$ is a multiple of $3$, and $p_5 = 5$ if $n$ is a multiple of $5$.
We may then classify the distance of $r \in V(G_n)$ away from zero, as follows.
\begin{itemize}
\item
Suppose that $n$ is a multiple of $3$ and some other $p_j \equiv 3 \pmod{4}$, and that either $n$ is relatively prime to $5$ or $r \not\equiv \pm1 \pmod{5}$.
By Lemma~\ref{lemma:oddPathsLength2}, we have $D_n(r) > 0$ if $r \equiv 0 \pmod{3}$, in which case it is at a distance of two from $0$.
Otherwise, for $r \equiv \pm 1 \pmod{3}$, let $s = r \mp u$ for $u \in Q_n$: then $s \equiv 0 \pmod{3}$.
Then $D_n(s) > 0$, in which case $r = u'' - u' \pm u$ for some choice of units $u', u'' \in Q_n$, so that $r$ can be reached from $0$ by a walk of length three.
\item
Suppose that $n$ is a multiple of $5$ and that $r \not\equiv 0 \pmod{5}$.
We may select coefficients $u_j \in Q_{\smash{p_j}^{\!m_j}}$ such that $r_5 - u_5 \in \ens{2,3}$, and such that $u_j \ne r_j$ for any $p_j \ge 7$.
Let $u = \sigma^{\text{--}1}(u_1, \ldots, u_t)$: by construction, we then have $r - u \equiv \pm 2 \pmod{5}$ and $r - u \not\equiv 0 \pmod{p_j}$ for $p_j \ge 7$.
Then either $S_n(r - u) > 0$, $S_n(u - r) > 0$, or $D_n(r - u) > 0$ (according to whether or not $n$ is a multiple of $3$, and which residue $r$ has modulo $3$ if so): $r$ can then be reached from $0$ by a path of length three.
\item
Suppose that $n$ is a multiple of $5$, and that $r \equiv 0 \pmod{5}$.
If $n$ is not a multiple of $3$, or if $r \equiv 0 \pmod{3}$, then $D_n(r) > 0$; $r$ can then be reached from $0$ by a walk of length two.
We may then suppose that $n$ is a multiple of $3$ and $r \equiv \pm 1 \pmod{3}$.
If we also have $r \not\equiv 0 \pmod{p_j}$ for any $p_j \equiv 3 \pmod{4}$, one of $S_n(r)$ or $S_n(-r)$ is non-zero; again, $r$ is at a distance of two from $0$.
Otherwise, we have $r \equiv 0 \pmod{p_j}$ for any $p_j \equiv 3 \pmod{4}$, so that $S_n(r) = S_n(-r) = D_n(r) = 0$; then $r$ has a distance at least three from $0$.
As well, any neighbor $s = r \pm u$ (for $u \in Q_n$ arbitrary) satsifies $s \equiv \pm 1 \pmod{5}$.
Then each neighbor of $r$ is then at distance three from $0$ in $G_n$, from which it follows that $r$ is at a distance of four from $0$.
\end{itemize}
Thus, there exist vertices at distance four from $0$ if $\gamma_3(n)\delta_3(n) + \gamma_5(n) = 2$; and apart from these vertices, or in the case that $\gamma_3(n)\delta_3(n) + \gamma_5(n) = 1$, each vertex is at a distance of at most three from $0$.
Then $\diam(G_n) = 2 + \gamma_3(n)\delta_3(n) + \gamma_5(n)$ if $\omega(n) > 1$, as required.
\end{proof}
\subsection{Restricted reachability results for $n$ coprime to $6$}
We may prove some stronger results on the reachability of vertices from $0$ in $G_n$ for $n$ odd: this will facilitate the analysis of perfectness results and the diameters for $n$ even.
\begin{definition}
\label{def:uniformDiameter}
For a \mbox{(di-)graph} $G$, the \emph{uniform diameter} $\udiam(G)$ is the minimum integer $d$ such that, for any two vertices $v,w \in V(G)$, there exists a (directed) walk of length $d$ from $v$ to $w$ in $G$.
\end{definition}
\noindent
Our interest in ``uniform'' diameters is due to the fact that if every vertex $v \in V(\Gamma_n)$ can be reached from $0$ by a path of exactly $d$ in $\Gamma_n$, then $v$ can also be reached from $0$ by a path of any length $\ell \ge d$ as well, which will prove useful for describing walks in $\Gamma_n$ to arbitrary vertices in terms of simultaneous walks in the digraphs $\Gamma_{\smash{p_j}^{\!m_j}}$.
We may easily show that $\Gamma_n$ has no uniform diameter when $n$ is a multiple of $3$.
For any adjacent vertices $v$ and $w$ such that $w - v \in Q_n$, we have $w - v \equiv 1 \pmod{3}$ by that very fact.
Then, there is a walk of length $\ell$ from $v$ to $w$ only if $\ell \equiv 1 \pmod{3}$; similarly, there is a walk of length $\ell$ from $w$ to $v$ only if $\ell \equiv 2 \pmod{3}$.
For similar reasons, $\Gamma_n$ has no uniform diameter for $n$ even.
However, for $n$ relatively prime to $6$, $\Gamma_n$ has a uniform diameter which may be easily characterized:
\begin{theorem}
\label{thm:oddUniformDiameter}
Let $n = p_1^{m_1} \cdots p_t^{m_t}$ be relatively prime to $6$.
Then
\begin{align*}
\udiam(\Gamma_n)
\;=\;
\begin{cases}
2 \;, & \text{if $n$ is coprime to $5$ and $\forall j : p_j \equiv 1 \!\!\!\!\pmod{4}$}; \\[0.5ex]
3 \;, & \text{if $n$ is coprime to $5$ and $\exists j : p_j \equiv 3 \!\!\!\!\pmod{4}$}; \\[0.5ex]
4 \;, & \text{if $n$ is a multiple of $5$}.
\end{cases}
\end{align*}
\end{theorem}
\begin{proof}
We begin by characterizing $\udiam(\Gamma_n)$, where $n = p^m$ for $p \ge 5$ prime, using Lemma~\ref{lemma:countingSumSquares} throughout to characterize $S_n(r)$ for $r \in \Z_n$.
\begin{itemize}
\item
If $p \equiv 1 \pmod{4}$ and $p > 5$, we have $S_{p^m}(r) > 0$ for all $r \in \Z_n$; then $\udiam(\Gamma_n) = 2$.
\item
If $p \equiv 3 \pmod{4}$ and $p > 5$, we have $S_{p^m}(r) = 0$ if and only if $r \in \Z_n$ is a zero divisor.
In particular, $\udiam(\Gamma_n) \ge 3$.
Conversely, as $\card{\Z_{p^m}\units} > p^{m-1}$, there exists $z \in Q_{p^m}\units$ such that $r - z$ is a unit; then there are quadratic units $x,y \in Q_{p^m}$ such that $r - z = x + y$, so that $\udiam(\Gamma_n) = 3$.
\item
If $p = 5$, we have $u \in Q_{5^m}$ if and only if $u \equiv \pm 1 \pmod{5}$; then $r$ can be expressed as a sum of $k$ quadratic units $r = u_1 + \cdots + u_k$ if and only if $r$ can be expressed modulo $5$ as a sum or difference of $k$ ones; that is, if $r \in \ens{-k , -k+2, \ldots, k-2, k} + 5 \Z_{5^m}$ (which exhausts $\Z_{5^m}$ for $k \ge 4$).
\end{itemize}
For $n$ not a prime power, we decompose $\Gamma_n \cong \Gamma_{\smash{p_1}^{\!m_1}} \ox \cdots \ox \Gamma_{\smash{p_t}^{\!m_t}}$; then a vertex $r = \sigma^{\text{--}1}(r_1, \ldots, r_t)$ is reachable by a walk of length $\ell$ in $\Gamma_n$ if and only if each $r_j \in V(\Gamma_{\smash{p_j}^{\!m_j}})$ are reachable by such a walk in their respective digraphs.
Thus, the uniform diameter of the tensor product is the maximum of the uniform diameters of each factor.
\end{proof}
The uniform diameter $\Gamma_n$ happens also to provide an upper bound on distances between vertices in $G_n$, under the constraint that we may only traverse walks $w_0\, w_1 \, \ldots \, w_\ell$ where the ``type'' of each transition $w_j \to* w_{j+1}$ is fixed to be either a quadratic unit or the negation of a quadratic unit, independently for each $j$.
More precisely:
\begin{lemma}
\label{lemma:oddSignWalks}
Let $n = p_1^{m_1} \cdots p_t^{m_t}$ be relatively prime to $6$, and $\ell \ge \udiam(\Gamma_n)$.
For any sequence $s_1, \ldots, s_\ell \in \ens{0,1}$, these exists a sequence of quadratic units $u_1, \ldots, u_\ell \in Q_n$ such that $r = (-1)^{s_1} u_1 + (-1)^{s_2} u_2 + \cdots + (-1)^{s_\ell} u_\ell$.
\end{lemma}
\begin{proof}
We first show that there are solutions to $r = u_1 - u_2 \pm u_3 \pm \cdots \pm u_\ell$, where all but the first two signs may be arbitrary.
We prove the result for $\ell = \udiam(\Gamma_n)$; one may extend to $\ell > \udiam(\Gamma_n)$ by induction.
\begin{itemize}
\item
Suppose $n$ is coprime to $5$: then for any $r \in \Z_n$, we have $D_n(r) > 0$, so that there exist $u, u' \in Q_n$ such that $r = u - u'$.
In the case that $n$ also has prime factors $p_j \equiv 3 \pmod{4}$, consider $s = r \mp u$ for any $u \in Q_n$: as there are solutions to $s = u - u'$ for $u,u' \in Q_n$, there are also solutions to $r = u - u' \pm u''$.
\item
Suppose $n = 5^{m_1} p_2^{m_2} \cdots p_t^{m_t}$.
\begin{itemize}
\item
If $r \not\equiv \pm1 \pmod{5}$.
Let $s \in \Z_n$ be such that $s \equiv 0 \pmod{5}$, and $s \not\equiv 0 \pmod{p_j}$ for any $p_j \ge 7$.
Then $r - s \not\equiv \pm1 \pmod{5}$, so that $D_n(r) > 0$; by Lemma~\ref{lemma:oddPathsLength2}, there are then quadratic units $u_1, u_2 \in Q_n$ such that $r - s = u_1 - u_2$.
We also have $S_n(s), S_n(-s), D_n(s) > 0$ by construction, which can be used to obtain decompositions $s = \pm u_3 \pm u_4$ for $u_3, u_4 \in Q_n$ depending on the choices of signs; we then have $r = u_1 - u_2 \pm u_3 \pm u_4$.
\item
If $r \equiv \pm 1 \pmod{5}$, consider $(r_1, \ldots, r_t) = \sigma(r)$.
\begin{subequations}
We select coefficients $u_j, u'_j \in Q_{\smash{p_j}^{\!m_j}}$ as follows.
We set $u'_1 = -u_1 = r_1$, so that
\begin{align}
(r_1 - 2u_1) \,\equiv\, (r_2 + 2u'_2) \,\equiv\, (r_1 - u_1 + u'_1) \,=\, \pm3 \!\!\pmod{5}.
\end{align}
For each $p_j \ge 7$, we require $u_j \ne 2^{\text{--}1} r_j$ and $u'_j \notin \ens{-2^{\text{--}1} r_j,\, u_j - r_j}$, but may otherwise leave $u_j$ unconstrained; we then have
\begin{align}
(r_j - 2u_j),\, (r_j + 2u_j'),\, (r_j - u_j + u_j') \,&\ne\, 0 \qquad\text{for $p_j \ge 7$}.
\end{align}
Let $u = \sigma^{\text{--}1}(u_1, \ldots, u_t)$ and $u' = \sigma^{\text{--}1}(u'_1, \ldots, u'_t)$.
By construction, we then have $D_n(r-2u), D_n(r+2u'), D_n(r-u+u') > 0$ by Lemma~\ref{lemma:oddPathsLength2}.
\end{subequations}
\begin{subequations}
There then exist $u'', u''' \in Q_n$ such that
\begin{align}
r \,=&\, u''' - u'' + u + u \,, \quad\text{or}
\\
r \,=&\, u''' - u'' + u - u' \,=\, u''' - u'' - u' + u \,,\quad\text{or}
\\
r \,=&\, u''' - u'' - u' - u'\,,
\end{align}
selecting $u'', u'''$ according to the desired signs for the latter two terms.
\end{subequations}
\end{itemize}
\end{itemize}
Thus, there are solutions to $r = u_1 - u_2 \pm u_3 \pm \cdots \pm u_\ell$ for $u_j \in Q_n$ and $\ell = \udiam(\Gamma_n)$, for arbitrary choices of signs and $r \in \Z_n$.
It follows that we may decompose $r = \pm u_1 \pm \cdots \pm u_\ell$ for arbitrary choices of sign, provided not all signs are the same.
By considering walks in $\Gamma_n$ of length $\udiam(\Gamma_n)$ from $0$ to either $r$ or $-r$, we also have decompositions $r = u_1 + \cdots + u_\ell$ and $r = -u_1 - \cdots - u_\ell$ for suitable choices of $u_1, \ldots, u_\ell \in Q_n$.
\end{proof}
\noindent The principal motivation for Lemma~\ref{lemma:oddSignWalks} is to bound the diameters of graphs $G_n$ over tensor decompositions of the ring $\Z_n$:
\begin{lemma}
\label{lemma:diameterFactors}
Let $M, N \ge 1$ be relatively prime integers, and let $n = MN$.
Then we have $\diam(G_n) \ge \max\ens{ \diam(G_N), \diam(G_M) }$.
Furthermore, if $M$ is coprime to $6$, we have $\diam(G_n) \le \max\ens{ \diam(G_N), \udiam(\Gamma_M) + 1 }$ as well.
\end{lemma}
\begin{proof}
Let $\rho : \Z_n \to \Z_N \oplus \Z_M$ be the natural isomorphism.
Let $r \in \Z_n$ be arbitrary, and $(r', r'') = \rho(r)$.
If $r = (-1)^{s_1} u_1 + \cdots + (-1)^{s_\ell} u_\ell$ for some $\ell > 0$ and $u_1, \ldots, u_\ell \in Q_n$, we also have
\begin{subequations}
\begin{align}
r'
\;=&\;\;
(-1)^{s_1} \, u'_1 \,+\, \cdots \,+\, (-1)^{s_\ell} \, u'_\ell \;,
\\[1ex]
r''
\;=&\;\;
(-1)^{s_1} \, u''_1 \,+\, \cdots \,+\, (-1)^{s_\ell} \, u''_\ell \;,
\end{align}
\end{subequations}
where $(u'_j, u''_j) = \rho(u_j)$.
For $\ell = \diam(G_n)$, it follows that $\ell \ge \diam(G_M)$ and $\ell \ge \diam(G_N)$.
Suppose further that $M$ is relatively prime to $6$: then $\udiam(\Gamma_M)$ is well-defined by Lemma~\ref{lemma:oddSignWalks}.
For any $a \in \Z_N$, let $\ell > 0$ be the length of a walk in $G_N$ from $0$ to $\ell$: there are then $u_1, \ldots, u_\ell \in Q_{N}$ and $s_1, \ldots, s_\ell \in \ens{0,1}$ such that $a = (-1)^{s_1} u'_1 \,+ \cdots +\, (-1)^{s_\ell} u'_\ell$.
If $\ell \ge \udiam(\Gamma_M)$, then for any $b \in \Z_M$, there also exist quadratic units $u''_1, \ldots, u''_\ell \in Q_M$ such that $b = (-1)^{s_1} u'''_1 \,+ \cdots +\, (-1)^{s_\ell} u'''_\ell$\,.
We may always obtain such a walk of length $\ell \ge \udiam(\Gamma_M)$ in $G_N$ by taking the shortest walk from $0$ to $a$ in $G_N$, and repeatedly adding closed walks of length two to the end until we obtain a walk of length $\ell \ge \udiam(\Gamma_M)$.
For such a walk, we then have
\begin{align}
\label{eqn:decomposeArbitrary}
r
\;=\;
\rho^{\text{--}1}(a,b)
\;=&\;
\rho^{\text{--}1}\Bigg(\sum_{j = 1}^\ell (-1)^{s_j} u'_j \;\;,\;\; \sum_{j = 1}^\ell (-1)^{s_j} u''_j\Bigg)
\notag\\=&\;
\sum_{j = 1}^\ell
(-1)^{s_j} \rho^{\text{--}1}(u'_j, u''_j)
\;=\;
\sum_{j = 1}^\ell
(-1)^{s_j} u_j \;,
\end{align}
for some choice of quadratic units $u_j = \rho^{\text{--}1}(u'_j, u''_j) \in Q_n$ and $s_j \in \ens{0,1}$.
If $\diam(G_N) > \udiam(\Gamma_M)$, this construction yields path-lengths $\udiam(\Gamma_M) \le \ell \le \diam(G_N)$; if instead $\udiam(\Gamma_M) \ge \diam(G_N)$, we obtain paths of length at most $\udiam(\Gamma_M) + 1$, which is saturated if there exist vertices $a \in V(G_N)$ whose distance $d_a$ from $0$ is such that $\udiam(\Gamma_M) - d_a$ is odd.
In either case, we have $\diam(G_n) \le \max\ens{\diam(G_N), \udiam(\Gamma_M)+1}$.
\end{proof}
\subsection{Diameter of $G_n$ for $n$ even}
The notable differences between the cases of $n$ odd and $n$ even are due to the sparsity of the quadratic units in $\Z_{2^m}$ compared to that of powers of other primes, and also that the sum or difference of two units (quadratic or otherwise) is necessarily a zero divisor if $n$ is even.
This results in a significant increase of the maximum diameter in the case of $n$ even, compared to $n$ odd:
\begin{theorem}
\label{thm:evenDiameter}
Let $n > 0$ even.
Let $\delta_3(n) = 1$ if $n$ has prime factors $p_j \equiv 3 \pmod{4}$ for $p_j > 3$, and $\delta_3(n) = 0$ otherwise.
Then we have
\begin{align}
\label{eqn:evenDiameterFormula}
\diam(G_n)
=&\,
\begin{cases}
12, & \text{if $n$ is a multiple of $24$}; \\[0.5ex]
6 , & \text{if $n$ is an odd multiple of $12$}; \\[0.5ex]
5 , & \text{if $n$ is a multiple of $10$, but not of $12$}; \\[0.5ex]
4 , & \text{if $n = 8K$ for $K > 0$ coprime to $15$}; \\[0.5ex]
3 + \delta_3(n), & \text{if $n = 6K$ for $K > 0$ coprime to $10$}; \\[0.5ex]
3 + \delta_3(n), & \text{if $n = 4K$ for $K > 1$ coprime to $30$}; \\[0.5ex]
3 , & \text{if $n = 2K$ for $K > 1$ coprime to $30$}; \\[0.5ex]
2 , & \text{if $n = 4$}; \\[0.5ex]
1 , & \text{if $n = 2$}.
\end{cases}
\end{align}
In particular, with Theorem~\ref{thm:oddDiameter}, we have $\diam(G_n) \le 12$ for all $n$, and $\diam(G_{p^m}) \le 4$ for any prime $p$ and $m \ge 0$.
\end{theorem}
\begin{proof}
We use Lemma~\ref{lemma:diameterFactors} to reduce the task of characterizing $\diam(G_n)$ for $n$ even to a small collection of representative cases, by factoring $n = NM$ for suitable choices of coprime factors $N$ and $M$.
\begin{itemize}
\item
Suppose $n$ is a multiple of $12$.
We may let $M$ be the largest factor of $n$ which is coprime to $12$, and $N = n/M$.
\begin{itemize}
\item
If $N = 2^m 3^{m'}$ for $m \ge 3$, we then have $u \in Q_N$ if and only if $u \equiv 1 \pmod{8}$ and $u \equiv 1 \pmod{3}$, or equivalently if $u \equiv 1 \pmod{24}$.
Then $T_N$ consists of those $q \in \Z_N$ such that $r \equiv \pm 1 \pmod{24}$.
The distance of a vertex in $G_N$ from $0$ is then characterized by its residue modulo $24$, in which case we may show that $\diam(G_N) = 12$.
\item
Otherwise, $N = 4 \cdot 3^{m'}$, in which case $u \in Q_N$ if and only if $u \equiv 1 \pmod{4}$ and $u \equiv 1 \pmod{3}$, or equivalently if $u \equiv 1 \pmod{12}$.
Then $T_N$ consists of those $q \in \Z_N$ such that $r \equiv \pm 1 \pmod{12}$; similarly as in the case above, we then have $\diam(G_N) = 6$.
\end{itemize}
Because $\diam(G_M), \udiam(\Gamma_M) \le 4$, we then have $\diam(G_n) = \diam(G_N)$ by Lemma~\ref{lemma:diameterFactors}.
Thus $\diam(G_n) = 12$ if $N$ is a multiple of $24$; otherwise we have $\diam(G_n) = 6$.
\item
Suppose $n$ is a multiple of $10$, but not of $12$: specifically, $n$ is not a multiple of $60$.
Let $M$ be the largest factor of $n$ which is coprime to $30$, and $N = n/M$.
We may show that $T_N$ contains only residues which are equivalent to $\pm 1$ modulo $10$:
\begin{itemize}
\item
If $n$ is an odd multiple of $30$, we have $N = 2 \cdot 3^m \cdot 5^{m'}$.
Then $u \in Q_N$ if and only if $u$ is odd, $u \equiv 1 \pmod{3}$, and $u \equiv \pm 1 \pmod{5}$; equivalently, if $u \equiv 1 \pmod{30}$ or $u \equiv 19 \equiv -11 \pmod{30}$.
\item
If $n$ is a multiple of $10$ but not of $30$, then without loss of generality $N = 2^{m_1} 5^{m_2}$.
We may show that $r \in Q_N$ if and only if both $r \equiv \pm 1 \pmod{5}$, and
\begin{align*}
r
\,\equiv\,
\begin{cases}
1 \!\!\!\!\pmod{2} \! & \text{if $m_1 = 1$}; \\[1ex]
1 \!\!\!\!\pmod{4} \! & \text{if $m_1 = 2$}; \\[1ex]
1 \!\!\!\!\pmod{8} \! & \text{if $m_1 \ge 3$}.
\end{cases}
\end{align*}
In each case, we have $u \in Q_N$ if and only if $u \in \ens{1,9} \pmod{\bar N}$ for $\bar N = 10$, $\bar N = 20$, or $\bar N = 40$ respectively.
\end{itemize}
As $N$ is a multiple of $10$ in either case, vertices $r \in \Z_N$ such that $r \equiv 5 \pmod{10}$ can only be reached by a path from $0$ with length at least five, so that $\diam(G_N) \ge 5$.
We may show that this bound is tight by showing that every even residue can be formed as a sum of four elements of $T_n$.
\begin{subequations}
Let $x \equiv_m y$ denote equivalence of two integers (or sets of integers) modulo $m$.
Then, we may easily verify that
\begin{align}
\begin{split}
\ens{\pm 1 \pm 1 \pm 1 \pm 1}
\;\equiv_{\scriptscriptstyle 30}&\;
\ens{26, 28, 0, 2, 4} ,
\\
\ens{\pm 1 \pm 1 \pm 1 \pm 11}
\;\equiv_{\scriptscriptstyle 30}&\;
\ens{8,10,12,14,16,18,20,22} ,
\\
-11 -11 -1 -1
\;\equiv_{\scriptscriptstyle 30}&\;\;
6 ,
\\
11 + 11 + 1 + 1
\;\equiv_{\scriptscriptstyle 30}&\;\;
24 ,
\end{split}
\end{align}
which proves the claim for $n$ an odd multiple of $30$.
For $n$ not a multiple of $30$, $T_N$ is the set of elements $q \in \Z_N$ such that $q = 5 \pm 4 \pmod{\bar N}$ or $q = -5 \pm 4 \pmod{\bar N}$.
It then suffices to show that all residues modulo $40$ are exhausted by sums or differences of four such residues: we have
\begin{align}
\begin{split}
\ens{\,\, (5\pm 4) + (5 \pm 4) + (5 \pm 4) + (5 \pm 4) \,\,}
\;\equiv_{\scriptscriptstyle 40}&\;
\ens{4, 12, 20, 28, 36} ,
\\
\ens{\,\, (5\pm 4) + (5 \pm 4) + (5 \pm 4) - (5 \pm 4) \,\,}
\;\equiv_{\scriptscriptstyle 40}&\;
\ens{34, \,2, 10, 18, 26} ,
\\
\ens{\,\, (5\pm 4) + (5 \pm 4) - (5 \pm 4) - (5 \pm 4) \,\,}
\;\equiv_{\scriptscriptstyle 40}&\;
\ens{24, 32, \,0, \,8, 16} ,
\\
\ens{\,\, (5\pm 4) - (5 \pm 4) - (5 \pm 4) - (5 \pm 4) \,\,}
\;\equiv_{\scriptscriptstyle 40}&\;
\ens{14, 22, 30, 38, \,6} .
\end{split}
\end{align}
\end{subequations}
As every odd residue modulo $N$ is adjacent to an even residue, it follows that every vertex in $G_N$ can be reached by a path of length at most five; then $\diam(G_N) = 5$.
As $M$ is coprime to both $3$ and $5$, we have $\diam(G_M) = 2$ and $\udiam(\Gamma_M) \le 3$; thus $\diam(G_n) = 5$ by Lemma~\ref{lemma:diameterFactors}.
\item
Suppose that $n = 8 K$ for $K$ coprime to $15$.
Let $M$ be the largest odd factor of $N$, and $N = n/M = 2^k$ for $k \ge 3$.
By construction, $M$ is coprime to $6$, so that $\udiam(\Gamma_M) \le 3$.
We have $u \in Q_N$ if and only if $u \equiv 1 \pmod{8}$: as every odd residue modulo $8$ can be expressed as a sum of three terms $\pm 1$, and every even residue modulo $8$ can be expressed as a sum of four terms $\pm 1$ (with $4$ requiring at least this many), it follows that $\diam(G_N) = 4$.
By Lemma~\ref{lemma:diameterFactors}, it follows that $\diam(G_n) = 4$ as well.
\item
In the remaining cases, we either have $n = 2K$ for $K$ coprime to $15$ and not a multiple of $4$, or $n = 6K$ for $K$ coprime to $10$.
We trivially have $\diam(G_n) = \frac{n}{2}$ for $n \in \ens{2,4}$; otherwise, $n$ has odd zero divisors.
As all walks of length one from $0$ in $G_n$ end at quadratic units, and all walks of length two from $0$ end at even elements of $\Z_n$, we require walks of length at least three from $0$ to reach odd zero divisors in $\Z_n$.
Thus, $\diam(G_n) \ge 3$.
Let $M$ be the largest factor of $n$ which is coprime to $30$.
By construction, $M$ is coprime to $6$, so that $\udiam(\Gamma_M) = 2 + \delta_3(M) = 2 + \delta_3(n)$.
\begin{itemize}
\item
If $n = 2K$ for $K$ coprime to $15$ and not a multiple of $4$, $M$ is simply the largest odd factor of $n$, in which case $n = 2^k$ for $k \in \ens{1,2}$.
We then have $N \in \ens{2,4}$, so that $\diam(G_N) = \frac{1}{2}N \le 2$.
\item
If $n = 6K$ for $K$ coprime to $10$, we have $N = 2 \cdot 3^k$ for some $k \ge 1$.
Then $u \in Q_N$ if and only if $u \equiv 1 \pmod{3}$ and is odd; that is, if $u \equiv 1 \pmod{6}$.
In particular, $T_N$ contains only elements which are equivalent to $\pm 1 \pmod{6}$; from this we may easily show $\diam(G_N) = 3$.
\end{itemize}
In either case, it follows by Lemma~\ref{lemma:diameterFactors} that
\begin{align}
3 \,\le\, \diam(G_n) \,\le\, \udiam(\Gamma_M) + 1 \,=\, 3 + \delta_3(n).
\end{align}
If $\delta_3(n) = 0$, we then have $\diam(G_n) = 3$; we may then restrict our attention to the case $\delta_3(n) = 1$.
Let $\rho: \Z_n \to \Z_M \oplus \Z_N$ be the natural isomorphism.
As $M$ is coprime to $15$, we have $D_M(a) > 0$ for every $a \in \Z_M$ by Lemma~\ref{lemma:oddPathsLength2}: then we may express any $a \in \Z_M$ as a difference $a = u'_1 - u'_2$ for $u'_1, u'_2 \in Q_M$.
\begin{itemize}
\item
If $N = 2$, any even residue $r$ may be expressed as $r = \rho^{\text{--}1}(a,0) = \rho^{\text{--}1}(u'_1,\,1) - \rho^{\text{--}1}(u'_2,\,1)$, which is a difference of the two quadratic units $u_j = \rho^{\text{--}1}(u'_j,\,1)$.
Thus every even residue can be reached in $G_n$ by a path of length two from $0$.
As every odd residue is adjacent to an even residue, we may reach any vertex by a path of length at most three; then $\diam(G_n) = 3$.
\item
If $N = 4$ or $N$ is a multiple of $6$, we have $u \in Q_N$ if and only if $u \equiv 1 \pmod{\bar N}$, where $\bar N = 4$ if $N = 4$, and $\bar N = 6$ otherwise.
We may easily show that the only residues $r \in \Z_n$ which may be expressed as a difference of two quadratic units are those such that $r \equiv 0 \pmod{\bar N}$; and for any residue $a \equiv 0 \pmod{p_j}$, we have $S_M(a) = S_M(-a) = 0$ by Lemma~\ref{lemma:oddPathsLength2}.
Therefore, no residue $r = \rho^{\text{--}1}(a,\pm 2) \in \Z_n$ can be reached by a path of length two from $0$ in $G_n$.
As any sum of the form $\pm u_1 \pm u_2 \pm u_3$ will be odd for $u_1, u_2, u_3 \in Q_n$, such residues $r$ are in fact at a distance at least four from zero.
As $\diam(G_n) \le 3 + \delta_3(n) = 4$, it follows that $\diam(G_n) = 4 = 3 + \delta_3(n)$ in this case.
\end{itemize}
\end{itemize}
In each case, the diameters agree with the formula in \eqref{eqn:evenDiameterFormula}.
\end{proof}
\subsection{Perfectness}
\label{sec:perfect}
A graph $G$ is \emph{perfect}~\cite{AR01} if, for every induced subgraph $H \subset G$, the size $\omega(H)$ of the maximum clique in $H$ is equal to the chromatic number $\chi(H)$.
We may easily identify two classes of quadratic unitary graphs which are perfect:
\begin{lemma}
\label{lemma:perfectSubcase}
For $n$ even or $n = p^m$ for $p \equiv 3 \pmod{4}$ prime, $G_n$ is perfect.
\end{lemma}
\begin{proof}
If $n$ is even, $G_n$ is bipartite, in which case $\omega(G_n) = \chi(G_n) = 2$ for any non-empty subgraph of $G_n$\,.
Otherwise, suppose that $n = p^m$ for a prime $p \equiv 3 \pmod{4}$.
Two vertices are adjacent in $G_{p^m}$ if and only if their residues modulo $p$ differ.
For any $H \subset G_{p^m}$, a maximum-size clique in $H$ is then any set of vertices having different residues modulo $p$, where the number of different residues represented is chosen to be maximal; at the same time, any minimum colouring of $G_{p^m}$ maps each residue class modulo $p$ to a common colour, with different residue classes having different colours.
Thus, $\omega(H) = \chi(H)$ for all such $H$, so that $G_{p^m}$ is perfect for $p \equiv 3 \pmod{4}$.
\end{proof}
A perfect graph $G$ contains no \emph{odd holes} (induced cycles of length $2k+1$ for $k > 1$, which have no cliques larger than two but are not bipartite).
Chudnovsky, Robertson, Seymour, and Thomas~\cite{CRST02} characterized perfect graphs in terms of odd holes, proving a conjecture of Berge~\cite{Berge}:
\medskip
\begin{named theorem}[Strong Perfect Graph Theorem]
A graph $G$ is perfect if and only if neither $G$ nor its complement $\bar G$ contain odd holes.
\end{named theorem}
\medskip
There exist imperfect quadratic unitary Cayley graphs $G_n$\,.
In particular, quadratic unitary Cayley graphs $G_p$ for $p \equiv 1 \pmod{4}$ are also circulant Paley graphs, which Maistrelli and Penman~\cite{MP06} show are imperfect by exploiting the fact that they are self-complementarity.\footnote{%
This is in fact the simplest case of a more comprehensive theorem, which shows that the only perfect Paley graph is that on nine vertices.
}
The Strong Perfect Graph Theorem implies that these graphs contain odd holes (again by self-complementarity) .
We may consider how to ``lift'' odd holes in such graphs $G_p$ to obtain odd holes in graphs $G_n$ having such factors $p$; and we may use a similar strategy for $G_n$ having distinct prime factors $p_1, p_2 \equiv 3 \pmod{4}$.
That is:
\begin{theorem}
\label{thm:perfectCharn}
$G_n$ is perfect if and only if $n$ is even, or $n = p^m$ for $p \equiv 3 \pmod{4}$ prime.
\end{theorem}
\begin{proof}
Building on Lemma~\ref{lemma:perfectSubcase}, we demonstrate that $G_n$ has an odd hole if $n$ is odd and is not a power of a prime $p \equiv 3 \pmod{4}$.
We proceed by constructing a simpler graph $G_\nu$, for $\nu$ a factor of $n$, which has an odd hole.
Let $n = p_1^{m_1} p_2^{m_2} \cdots p_t^{m_t}$\,:
\begin{enumerate}
\item
Suppose that $n$ has a prime factor which is equivalent to $1 \pmod{4}$.
Without loss of generality, $p_1 \equiv 1 \pmod{4}$; we then let $\nu = p_1$.
By~\cite{MP06}, $G_\nu$ is then not perfect, and so $G_\nu$ or its complement has an odd hole.
Note that $G_\nu$ is self-complementary, as multiplication of any pair of adjacent vertices by a non-quadratic unit $r$ yields a non-adjacent pair, and vice-versa.
Thus, $G_\nu$ contains an odd hole $u_0 ~ u_1 ~ u_2 ~ \cdots ~ u_{\ell-1} ~ u_0$ of some odd length $\ell$.
\item
Suppose instead that $n$ has two distinct prime factors equivalent to $3 \pmod{4}$.
Without loss of generality, $p_1,p_2 \equiv 3 \pmod{4}$; we then define $\nu = p_1 p_2$\,, and let $\rho: \Z_\nu \to \Z_{p_1} \oplus \Z_{p_2}$ be the natural isomorphism.
We demonstrate the existence of an induced five-cycle $u_0 ~ u_1 ~ u_2 ~ u_3 ~ u_4 ~ u_0$ in $G_\nu$ by constructing appropriate structures in the digraphs $\Gamma_{p_1} \ox \Gamma_{p_2} = \rho(\Gamma_\nu)$\,.
Let
\begin{align}
u_0 \;=\; 0 \;=&\; \rho^{\text{--}1}(0,0) \,, &
u_1 \;=\; 1 \;=&\; \rho^{\text{--}1}(1,1) \,.
\end{align}
We proceed by cases:
\begin{itemize}
\item
Suppose $2$ is not a quadratic residue modulo either $p_1$ or $p_2$\,.
Without loss of generality, we may suppose $3 \le p_1 < p_2$; in particular, $p_2 = 11$ or $p_2 \ge 19$.
By~\cite{B1925}, there are then consecutive triples $q, q+1, q+2 \in Q_{p_2}$.
Let $q \in Q_{p_2}$ be a minimal such residue: as $2 \notin Q_{p_2}$\,, it follows that $q \ne 1$, so that $q-1 \in -Q_{p_2}$.
In particular, there is an arc from $q+1$ to $2$.
Define the vertices
\begin{align}
\rho(u_2) \;=&\; (2,q+1) \,, &
\rho(u_3) \;=&\; (0,2) \,, &
\rho(u_4) \;=&\; (2,-q) \,.
\end{align}
The vertices are distinct; in particular, $u_2$ differs from $u_4$, because our choice that $q+1 \in Q_{p_2}$ implies that $q+1 \ne -q$.
By agreement of arc-directions $x_j \arc x_k \in E(\Gamma_{p_1})$ and $y_j \arc y_k \in E(\Gamma_{p_2})$ for pairs of coefficients $(x_k, y_k) = \rho(u_k)$, we have the arc-relations
\begin{align}
u_0 \arc u_1 \arc u_2 \arc u_3 \cra u_4 \cra u_0
\end{align}
in $\Gamma_\nu$\,.
There are no other arc relations among the vertices $u_k$\,.
In particular, $u_0$ and $u_3$ are non-adjacent as they have the same residue modulo $p_1$\,, and similarly for $u_2$ and $u_4$\,; and there are no arcs between the other pairs of vertices $u_j$ and $u_k$ because we have $x_j - x_k \in \pm Q_{p_1}$ if and only if $y_j - y_k \in \mp Q_{p_2}$\,.
\item
Suppose that $2$ is a quadratic residue modulo only one of $p_1$ or $p_2$\,.
Without loss of generality, we may suppose $2 \in Q_{p_2}$\,, in which case $p_2 \ge 7$\,.
Define the vertices
\begin{align}
\rho(u_2) \;=&\; (2,2) \,, &
\rho(u_3) \;=&\; (0,3) \,, &
\rho(u_4) \;=&\; (1,4) \,.
\end{align}
Clearly the vertices $u_j$ are distinct.
By agreement of arc-directions $x_j \arc x_k \in E(\Gamma_{p_1})$ and $y_j \arc y_k \in E(\Gamma_{p_2})$ for pairs of coefficients $(x_k, y_k) = \rho(u_k)$, we have the arc-relations
\begin{align}
u_0 \arc u_1 \arc u_2 \arc u_3 \arc u_4 \cra u_0\
\end{align}
in $\Gamma_\nu$\,.
There are no other arc relations among the vertices $u_k$\,.
In particular, $u_0$ and $u_3$ are non-adjacent as they have the same residue modulo $p_1$\,, and similarly for $u_1$ and $u_4$\,; and there are no arcs between the other pairs of vertices $u_j$ and $u_k$ because we have $x_j - x_k \in \pm Q_{p_1}$ if and only if $y_j - y_k \in \mp Q_{p_2}$\,.
\item
Suppose that $2$ is a quadratic residue modulo both $p_1$ and $p_2$\,.
In particular, $p_1, p_2 \ge 7$, so that there are pairs of consecutive quadratic residues modulo $p_1$ and $p_2$ (see Lemma~\ref{lemma:consecutiveResidues}).
Let $q_1-1, q_1 \in Q_{p_1}$ and $q_2-1, q_2 \in Q_{p_2}$ be the largest such pairs in each case; in particular it follows that $q_2 \ne -1$, so that $-q_2 -1 \in Q_{p_2}$\,.
Define the vertices
\begin{align}
\rho(u_2) \;=&\; (q_1,-q_2) \,, &
\rho(u_3) \;=&\; (0,1) \,, &
\rho(u_4) \;=&\; (1,q_2) \,.
\end{align}
The vertices are again distinct; and by arc agreement $x_j \arc x_k \in E(\Gamma_{p_1})$ and $y_j \arc y_k \in E(\Gamma_{p_2})$ for pairs of coefficients $(x_k, y_k) = \rho(u_k)$, we have the arc-relations
\begin{align}
u_0 \arc u_1 \arc u_2 \cra u_3 \arc u_4 \cra u_0
\end{align}
in $\Gamma_\nu$\,.
There are no other arc relations among the vertices $u_k$\,.
In particular, $u_0$ and $u_3$ are non-adjacent as they have the same residue modulo $p_1$\,, and similarly for $u_1$ and $u_4$\,; the vertices $u_1$ and $u_3$ have the same residue modulo $p_2$\,.
By construction, we have $2q_2 \in Q_{p_2}$ and $1 - q_1 \in -Q_{p_1}$\,, so there is no arc between $u_2$ and $u_4$\,; and similarly there is no arc between $u_0$ and $u_2$\,.
\end{itemize}
\end{enumerate}
In each case, $G_\nu$ contains an induced cycle $u_0 ~ u_1 ~ u_2 ~ \cdots ~ u_{\ell-1} ~ u_0$ for some odd $\ell \ge 5$.
Let
\begin{align}
\mu
\;=\;
\begin{cases}
p_1^{m_1} , & \text{if $\nu = p_1$}, \\[1ex]
p_1^{m_1} p_2^{m_2} , & \text{if $\nu = p_1 p_2$}:
\end{cases}
\end{align}
we may then obtain a similar odd hole in $G_\mu$ by identifying each $u_j$ with a corresponding $x_j \in \ens{0, \ldots, \nu-1} \subset \Z_\mu$.
Then $x_j - x_k \in Q_\mu$ if and only if $u_j - u_k \in Q_\nu$ for any $1 \le j,k \le \ell$, which implies that $x_0 ~ x_1 ~ \cdots ~ x_{\ell-1} ~ x_0$ is a cycle without chords in $G_\mu$.
Let $N$ be the largest factor of $n$ which is coprime to $\mu$ (\ie\ $N = n / \mu$), and let $\tau: \Z_n \to \Z_\mu \oplus \Z_N$ be the natural isomorphism.
\begin{itemize}
\item
Suppose $\nu = p_1 \equiv 1 \pmod{4}$.
If $n$ is a multiple of $5$, we may suppose that $p_1 = 5$ without loss of generality; then $N$ is coprime to $5$.
By Lemma~\ref{lemma:oddPathsLength2}, we then have $S_N(r) > 0$ for any $-r \in Q_N$: then $G_N$ contains a closed walk $0 \; u \; r \; 0$ for some $u \in Q_N$.
By concatenating this walk with $\frac{1}{2}(\ell - 3)$ copies of a closed walk of length two at $0$, we obtain a closed walk $y_0 ~ y_1 ~ \cdots ~ y_{\ell-1} \, y_0$ in $G_N$\,.
We may then construct a walk
\begin{align}
\label{eqn:constructedOddHole}
C = (x_0, y_0) ~ (x_1, y_1) ~ \cdots ~ (x_{\ell-1}, y_{\ell-1}) ~ (x_0, y_0)
\end{align}
in $G_\mu \ox G_N$: because $x_0 ~ x_1 ~ \cdots ~ x_{\ell-1} ~ x_0$ is an induced cycle, so is $C$.
\item
Otherwise, suppose $\mu = p_1^{m_1} p_2^{m_2}$.
If $n$ is a multiple of $3$, we may suppose that $p_1 = 3$ without loss of generality; then $N$ is coprime to $3$.
By Theorem~\ref{thm:oddUniformDiameter}, we then have $\ell > \udiam(\Gamma_N)$, in which case by Lemma~\ref{lemma:oddSignWalks} we may construct a closed walk $y_0 ~ y_1 ~ \cdots ~ y_{\ell-1} ~ y_0$ in $G_N$ by setting $y_0 = 0$, and letting $y_{j+1} - y_j \in \pm Q_N$ whenever $x_{j+1} - x_j \in \pm Q_\mu$ (\ie~with agreement in the signs), and similarly for $y_0 - y_{\ell-1}$.
Define the walk $C$ in $\Z_\mu \oplus \Z_N$ as given in \eqref{eqn:constructedOddHole}: we then have $(x_{j+1}, y_{j+1}) - (x_j, y_j) \in \pm (Q_\mu \oplus Q_\nu)$ for each $j$, and similarly $(x_0, y_0) - (x_{\ell-1}, y_{\ell-1}) \in \pm (Q_\mu \oplus Q_N)$.
\end{itemize}
In either case, if we define vertices $v_j = \tau^{\text{--}1}(x_j, y_j) \in V(G_n)$, the walk $v_0 ~ v_1 ~ \cdots ~ v_{\ell-1} \, v_0$ is an induced cycle of odd length in $G_n$.
Thus $G_n$ is not perfect, unless $n$ is even or a power of a prime $p \equiv 3 \pmod{4}$.
\end{proof}
\section{Decomposing symplectic operators mod $n$}
Our final result is a bound on the complexity of decompositions of symplectic operastors modulo $n$, which follows from the bound on the diameter of $G_n$.
We may define the \emph{symplectic form} (modulo $n$) as the $2m \x 2m$ matrix
\begin{align}
\sigma_{2m}
\;=\;
\left[\,\begin{matrix}
0_m & \!-I_m \\[0.5ex]
\;I_m\; & 0_m
\end{matrix}\,\right]
\,;
\end{align}
the \emph{symplectic group modulo $n$} $\Sp_{2m}(\Z_n)$ is the set of $2m \x 2m$ linear operators $S$ (\emph{symplectic operators}) with coefficients in $\Z_n$ such that $S\trans \sigma_{2m} S = \sigma_{2m}$.
\begin{convention}
For operators $S \in \Sp_{2m}(\Z_n)$ for a fixed $m$, we will adopt the convention of indexing the rows and columns by integers modulo $2m$, starting with $1$.
Thus, for a row $k \in \ens{m+1, \ldots, 2m}$ in the ``bottom'' half of a matrix $S$, the row $k+m \in \ens{1, \ldots, m}$ will be in the ``top'' half, and vice-versa.
\end{convention}
Symplectic operators are clearly invertible operations, and therefore may be reduced to $I_{2m}$ by Gaussian elimination.
We also consider a variant procedure, in which row-operations are constrained to themselves be symplectic.
For the operator definitions below, actions of operators are defined via the action of left-multiplication on a square matrix over $\Z_n$.
\begin{definition}
For row-indices $j,k \in \ens{1, \ldots, 2m}$, a \emph{symplectic row operation acting on $\Sp_{2m}(\Z_n)$} is one of the operators $M\supp{\alpha}_j$, $E_{j,k}$\,, $C_{j,k}$\,, or $C_{j,k}^{-1}$ defined as follows:
\begin{itemize}
\item
For any $\alpha \in \Z_n\units$, let $\mu_j\supp{\alpha} \in \GL_{2n}(\Z_n)$ be the linear operator which multiplies the $j\textsuperscript{th}$ row of its operand by $\alpha$.
Then, we define $M_j\supp{\alpha} \,=\, \mu_j\supp{\alpha}\, \mu_{j+m}\supp{\smash{\alpha^{\text{--}1}}}$.
\item
Let $\epsilon_{j,k} \in \GL_{2m}(\Z_n)$ be the linear operator which exchanges rows $j$ and $k$ of its operand.
Then we define
\begin{align}
E_{j,k}
\;=\;
\begin{cases}
\epsilon_{j,k} \, \mu_k\supp{-1} \,, & \text{if $j - k \equiv m \pmod{2m}$}; \\[1ex]
\epsilon_{j,k} \, \epsilon_{j+m,k+m} & \text{otherwise},
\end{cases}
\end{align}
for $\mu_j\supp{-1}$ as defined above.
\item
Let $\chi_{j,k} \in \GL_{2m}(\Z_n)$ be the linear operator which adds row $j$ of its operand to row $k$.
Then we define
\begin{align}
C_{j,k}
\;=\;
\begin{cases}
\chi_{j,k} \,, & \text{if $j - k \equiv m \pmod{2m}$}; \\[1ex]
\chi_{j,k} \, \chi_{k+m,j+m}^{-1} \,, & \text{if $1 \le j,k \le m$ or $m+1 \le j,k \le 2m$}; \\[1ex]
\chi_{j,k} \, \chi_{k+m,j+m} \,, & \text{otherwise}.
\end{cases}
\end{align}
\end{itemize}
\end{definition}
These operations are defined so as to be symplectic themselves; we wish to demonstrate an upper bound to the number of such symplectic row operations required to transform an arbitrary symplectic operator to the identity.
Hostens \emph{et al.}~\cite{HDM05} provide a decomposition of symplectic operators into $O(m^2 \log(n))$ symplectic row operations, in an application to the the decomposition of an important family of unitary operators for quantum computation (specifically, the Clifford group over qudits of dimension $n$).
We refine this decomposition to obtain an upper bound to $O(m^2)$, giving an upper bound which is independent of the modulus $n$.
\subsection{Reduction to greatest common divisors modulo $n$}
We first describe the decomposition of~\cite{HDM05} in detail.
The main concept is to reduce $S \in \Sp_{2m}(\Z_n)$ to another operator $S'$ which acts trivially on, \eg, the standard basis vectors $\unit_m, \unit_{2m}$.
This reduces the problem to decomposing an operator $\tilde S \in \Sp_{2m-2}(\Z_n)$,
\begin{align}
\label{eqn:symplecticGroupReduction}
\tilde S
\;=&\;
\left[\,\begin{matrix}
A'_{11} & A'_{12} \\[0.5ex]
A'_{21} & A'_{22}
\end{matrix}\,\right]
&
\text{for}\;\;
S'
\,=&\;
\left[\;\begin{array}{c@{\,}c@{\,}c|c|c@{\,}c@{\,}c|c}
& & & 0 & & & & 0 \\[-0.8ex]
& A'_{11} & & \vdots & & A'_{12} & & \vdots \\
& & & 0 & & & & 0 \\
\hline
0 & \cdots & 0 & \;1\; & 0 & \cdots & 0 & 0 \\
\hline
& & & 0 & & & & 0 \\[-0.8ex]
& A'_{21} & & \vdots & & A'_{22} & & \vdots \\
& & & 0 & & & & 0 \\
\hline
0 & \cdots & 0 & 0 & 0 & \cdots & 0 & \;1\;
\end{array}\;\right].
\end{align}
Embedding the matrix groups $\Sp_2(\Z_n) \subset \cdots \subset \Sp_{2m-2}(\Z_n) \subset \Sp_{2m}(\Z_n)$ in the manner described above, one may recursively apply this process to obtain a sequence of symplectic row operations which multiply to transform $S$ to $I_{2m}$.
As the inverse of each symplectic row operation is also a symplectic row operation, this yields a decomposition of $S$.
The reduction from $S$ to $S'$ as above is performed by a hybrid of Gaussian elimination and Euclid's algorithm for computing greatest common divisors.
\begin{subequations}
We illustrate this on a $2m \x 2$ matrix $[\; \vec v \; \vec w \;]$, for a pair of column vectors $\vec v = [\; v_1 \;\; v_2 \;\; \cdots \;\; v_{2m} \;]\trans$ and $\vec w = [\; w_1 \;\; w_2 \;\; \cdots \;\; w_{2m} \;]\trans$ subject to the constraint $\vec w\trans \sigma_{2m} \vec v = 1$.
By performing suitable symplectic row-additions, we may simulate the Euclidean algorithm in the second column, for each pair of rows $(j,\, j+m)$ for $j \in \ens{1, \ldots, m}$, to obtain
\begin{align}
\label{eqn:localGCDcompute}
\left[\;\begin{matrix}
v_1 & w_1 \\[0.2ex] v_2 & w_2 \\[0.2ex] \vdots & \vdots \\[0.2ex] v_m & w_m
\\[0.5ex]
\hline
\\[-2ex]
v_{m+1} & w_{m+1} \\[0.2ex] v_{m+2} & w_{m+2} \\[0.2ex] \vdots & \vdots \\[0.2ex] v_{2m} & w_{2m}
\end{matrix}\;\right]
\,\mapsto\,
\left[\;\begin{matrix}
\tilde v_1 & 0 \\[0.2ex] \tilde v_2 & 0 \\[0.2ex] \vdots & \vdots \\[0.2ex] \tilde v_m & 0
\\[0.5ex]
\hline
\\[-2ex]
\tilde v_{m+1} & \gcd(w_1,w_{m+1},n) \\[0.2ex] \tilde v_{m+2} & \gcd(w_2,w_{m+2},n) \\[0.2ex] \vdots & \vdots \\[0.2ex] \tilde v_{2m} & \gcd(w_m,w_{2m},n)
\end{matrix}\;\right]
\,=:\,
\left[\;\begin{matrix}
\tilde v_1 & 0 \\[0.2ex] \tilde v_2 & 0 \\[0.2ex] \vdots & \vdots \\[0.2ex] \tilde v_m & 0
\\[0.5ex]
\hline
\\[-2ex]
\tilde v_{m+1} & \gamma_1 \\[0.2ex] \tilde v_{m+2} & \gamma_2 \\[0.2ex] \vdots & \vdots \\[0.2ex] \tilde v_{2m} & \gamma_m
\end{matrix}\;\right],
\end{align}
computing ``greatest common divisors'' (modulo $n$) in the lower block in the second column, and using these to clear the upper block.
We then perform further row-additions to compute further greatest common divisors in the second column, in pairs of rows $(j, j+1)$ for $j \in \ens{m+1, \ldots, 2m-1}$, in to perform the following transformation of the the second column:
\begin{align}
\label{eqn:staircaseGCDcompute}
\left[\;\begin{matrix}
0 \\ 0 \\ \vdots \\ 0 \\
\hline
\gamma_1 \\ \gamma_2 \\ \vdots \\ \gamma_m
\end{matrix}\;\right]
\,\mapsto\,
\left[\;\begin{matrix}
0 \\ 0 \\ \vdots \\ 0 \\
\hline
0 \\ \gcd(\gamma_1,\gamma_2) \\ \vdots \\ \gamma_m
\end{matrix}\;\right]
\,\mapsto\,
\cdots
\,\mapsto\,
\left[\;\begin{matrix}
0 \\ 0 \\ \vdots \\ 0 \\
\hline
0 \\ 0 \\ \vdots \\ \gcd(\gamma_1, \gamma_2, \ldots, \gamma_{2m})
\end{matrix}\;\right].
\end{align}
\end{subequations}
Note that as $\vec w\trans \sigma_{2m} \vec v = 1$, there is an integer combination of the coefficients of $\vec w$ which is equivalent to $1$ modulo $n$; then
\begin{align}
\gcd(\gamma_1, \ldots, \gamma_m) \;=\; \gcd(w_1, \ldots, w_{2m}, n) \;=\; 1\,.
\end{align}
The above row-transformations then transform the two-column matrix $[\; \vec v \;\, \vec w\;]$ as follows:
\begin{align}
\left[\;\begin{matrix}
v_1 & w_1 \\[0.2ex] v_2 & w_2 \\[0.2ex] \vdots & \vdots \\[0.2ex] v_m & w_m
\\[0.5ex]
\hline
\\[-2ex]
v_{m+1} & w_{m+1} \\[0.2ex] v_{m+2} & w_{m+2} \\[0.2ex] \vdots & \vdots \\[0.2ex] v_{2m} & w_{2m}
\end{matrix}\;\right]
\,\mapsto*\,
\left[\;\begin{matrix}
v'_1 & 0 \\[0.2ex] v'_2 & 0 \\[0.2ex] \vdots & \vdots \\[0.2ex] 1 & \;\;0\;\;
\\[0.5ex]
\hline
\\[-2ex]
v'_{m+1} & 0 \\ v'_{m+2} & 0 \\ \vdots & \vdots \\ v'_{2m} & 1
\end{matrix}\;\right]
\,=:\,
\Big[\; \vec v' \;\; \unit_{2m} \;\Big],
\end{align}
where $v'_m = 1$ follows from $\unit_{2m}\trans \sigma_{2m} \vec v' = \vec w\trans \sigma_{2m} \vec v = 1$.
We may repeat the sequence of transformations to compute greatest common divisors in the first column, for each pair of rows $(j,j+m)$ for $j \in \ens{1, \ldots, m}$, and subsequently in row-pairs $(j,j+1)$ in the upper block:
\begin{align}
\label{eqn:symplecticReduceFirstColumn}
\left[\;\begin{matrix}
v'_1 & 0 \\[0.2ex] v'_2 & 0 \\[0.2ex] \vdots & \vdots \\[0.2ex] 1 & \;0\;
\\[0.5ex]
\hline
\\[-2ex]
v'_{m+1} & 0 \\ v'_{m+2} & 0 \\ \vdots & \vdots \\ v'_{2m} & 1
\end{matrix}\;\right]
\,\mapsto*\,
\left[\;\begin{matrix}
\varphi_1 & 0 \\[0.2ex] \varphi_2 & 0 \\[0.2ex] \vdots & \vdots \\[0.2ex] 1 & \;0\;
\\[0.5ex]
\hline
\\[-2ex]
0 & 0 \\ 0 & 0 \\ \vdots & \vdots \\ 0 & 1
\end{matrix}\;\right]
\,\mapsto*\,
\left[\;\begin{matrix}
0 & 0 \\[0.2ex] \gcd(\varphi_1,\varphi_2) & 0 \\[0.2ex] \vdots & \vdots \\[0.2ex] 1 & \:\;0\:\;
\\[0.5ex]
\hline
\\[-2ex]
0 & 0 \\ 0 & 0 \\ \vdots & \vdots \\ 0 & 1
\end{matrix}\;\right]
\,\mapsto*\,
\left[\;\begin{matrix}
0 & 0 \\[0.2ex] 0 & 0 \\[0.2ex] \vdots & \vdots \\[0.2ex] 1 & \;0\;
\\[0.5ex]
\hline
\\[-2ex]
0 & 0 \\ 0 & 0 \\ \vdots & \vdots \\ \;0\; & 1
\end{matrix}\;\right].
\end{align}
The reduction of~\cite{HDM05} applies this procedure for $\vec v = S\unit_m$, $\vec w = S\unit_{2m}$.
Applying these transformations to $S$ yields a matrix $S'$ as illustrated in \eqref{eqn:symplecticGroupReduction}, as the other columns $S' \unit_k$ for $k \notin \ens{m,2m}$ must satisfy
\begin{align}
\unit_k\trans S' \sigma_{2m} \unit_m \;=\; \unit_k S \sigma_{2m} S \unit_m \;=\; 0,
\end{align}
and similarly $\unit_k\trans S' \sigma_{2m} \unit_{2m} = 0$.
The complexity of a single iteration of this reduction is $O(m \log(n))$, which arises from the cost of repeating Euclid's algorithm (expressed in fixed-width integer addition steps) $O(m)$ times to reduce the $m\th$ and $2m\th$ columns to $\unit_m$ and $\unit_{2m}$ respectively.
Iterated $m$ times over all column-pairs $\unit_j, \unit_{j+m}$, we obtain the upper bound of $O(m^2 \log(n))$ reported by~\cite{HDM05}.
\subsection{Improved upper bounds via the diameter of $G_n$}
\bgroup
\def^{\text{--}1}{^{\text{--}1}}
The complexity of the above decomposition may be reduced to $O(m^2)$, by substituting an explicit simulation of Euclid's algorithm via symplectic row transformations with a product of constant size.
This is possible by using short paths in the graphs $G_n$ to reduce the number of addition steps in order to obtain coefficients $\gamma_j$ and $\varphi_j$ (or coefficients equivalent to them, up to a multiplicative unit) using a constant number of row-operations.
The primary obstacle to reducing the complexity of a single iteration of the reduction of~\cite{HDM05} is the computation of greatest common divisors in row-pairs $(j,j+m)$, arising from constraints on obtaining ``derived'' row-additions on these row-pairs.
\begin{subequations}
\label{eqn:derivedRowAdditions}
The iterated operator $C_{j,k}^{\,\alpha}$ (for $\alpha \in \ens{0, 1, \ldots, n-1}$, which we identify with $\alpha \in \Z_n$) can be easily obtained in constant depth for $j \not\equiv k + m \pmod{2m}$ and $\alpha \in \Z_n\units$, by the equality
\begin{align}
C_{j,k}^{\,\alpha}
\:=\;
M\supp{\smash{\alpha^{\text{--}1}}}_j \,
C_{j,k} \;
M\supp{\alpha}_j \;,
\end{align}
which one may verify by the action on standard basis vectors.
However, $C_{j,j+m}^{\,\alpha}$ cannot be decomposed in this manner: the closest we may come is in the case where $\alpha = u^2$ for some $u \in \Z_n\units$, in which case we have
\begin{align}
C_{j,k}^{\,\alpha}
\:=\;
C_{j,k}^{\,u^2}
\:=\;
M\supp{\smash{u^{\text{--}1}}}_j \,
C_{j,j+m} \;
M\supp{u}_j \;.
\end{align}
\end{subequations}
We may apply the result of Theorem~\ref{thm:evenDiameter} as follows:
\begin{lemma}
For distinct row-indices $j,k \in \ens{1, \ldots, 2m}$ and for any $\alpha \in \Z_n$, there exists a sequence of units $a_1, \ldots, a_\ell \in \Z_n\units$ and signs $s_1, \ldots, s_\ell \in \ens{-1, +1}$ for some $\ell \le 12$, such that
\begin{align}
C_{j,k}^{\,\alpha}
\;=&\:\:
M_j^{a_1^{\text{--}1}} C^{\,s_1}_{j,k} \,
M_j^{a_2^{\text{--}1} a_1} \, C^{\,s_2}_{j,k} \,
M_j^{a_3^{\text{--}1} a_2} \, \cdots \,
M_j^{a_\ell^{\text{--}1} a_{\ell-1}} C^{\,s_\ell}_{j,k} \, M_j^{a_\ell} \;.
\end{align}
\end{lemma}
\begin{proof}
It suffices to note that as $\diam(G_n) \le 12$, there exists such a sequence of signs and quadratic units $u_1, \ldots, u_\ell \in Q_n$ such that $\alpha = s_1 u_1 + s_2 u_2 + \cdots + s_\ell u_\ell$.
We may then take either $a_j = u_j$ (in the case that $k \ne j + m$) or a unit $a_j$ such that $u_j = a_j^2$ (in the case that $k = j + m$), and apply the decompositions of \eqref{eqn:derivedRowAdditions} to obtain the desired decomposition.\footnote{%
For $k \ne j + m$, we may in fact obtain the further bound of $\ell \le 3$, as the diameter of the unitary Cayley graph $X_n = \Cay(\Z_n, \Z_n\units)$ is at most three~\cite{KS07}.}
\end{proof}
We may apply this to reduce the complexity of decomposing symplectic operators as follows.
We use the following additional Lemma, whose proof is deferred to the appendix:
\begin{lemma}
\label{lemma:Bezout+n}
Let $\gamma = \gcd(x, y, n)$: then there exist $a,b,c \in \Z$ such that $ax + by + cn = \gamma$ and where both $a$ and $b$ are relatively prime to $n$.
\end{lemma}
For a vector $\vec x = [\; x_1 \;\; x_2 \;\; \cdots \;\; x_{2m} \;]\trans$, let $\gamma_j = \gcd(x_j, x_{j+m}, n)$ for each $j \in \ens{1, \ldots, m}$.
Let $a_j$ be coefficients such that $a_{j+m} x_{j+m} \,+\, a_j x_j \,\equiv\, \gamma_j \pmod{n}$ as guaranteed by Lemma~\ref{lemma:Bezout+n}, and define $r_j = a_{j+m}^{-1} a_j$: we then have
\begin{align}
C^{\,r_1}_{1,m+1}
\:\! \cdots
\:\! C^{\,r_m}_{m,2m}
\,\vec x
\:\!=\:\!
\left[\;\begin{matrix}
x_1 \\[0.5ex] x_2 \\[0.5ex] \vdots \\[0.5ex] x_m
\\[0.5ex]
\hline
\\[-2ex]
x_{m+1} + a_{m+1}^{-1} a_1 x_1 \\[1ex] x_{m+2} + a_{m+2}^{-1} a_2 x_2 \\[1ex] \vdots \\[1ex] x_{2m} + a_{2m}^{-1} a_m x_m
\end{matrix}\;\right]
\:\!=\:\!
\left[\;\begin{matrix}
x_1 \\[0.5ex] x_2 \\[0.5ex] \vdots \\[0.5ex] x_m
\\[0.5ex]
\hline
\\[-2ex]
a_{m+1}^{-1} \gamma_1 \\[1ex] a_{m+2}^{-1} \gamma_2 \\[1ex] \vdots \\[1ex] a_{2m}^{-1} \gamma_m
\end{matrix}\;\right]
\:\!=:\:\!
\left[\;\begin{matrix}
x_1 \\[0.5ex] x_2 \\[0.5ex] \vdots \\[0.5ex] x_m
\\[0.5ex]
\hline
\\[-2ex]
\tilde \gamma_1 \\[1ex] \tilde \gamma_2 \\[1ex] \vdots \\[1ex] \tilde \gamma_m
\end{matrix}\;\right].
\end{align}
Each coefficient $\tilde \gamma_j$ generates the same additive subgroup as $\gamma_j$ modulo $n$; if $d_1, \ldots, d_m$ are coefficients such that $x_j = d_m a_{j+m}^{-1} \gamma_j = d_m \tilde \gamma_j$, we then have
\begin{align}
C^{-d_1}_{m+1,1}
\:\! \cdots
\:\! C^{-d_m}_{2m,m}
\big[\; x_1 \; \cdots \; x_m \;\; \tilde\gamma_1 \; \cdots \; \tilde \gamma_m \;\big]\trans
=\;
\big[\; 0 \; \cdots \; 0 \;\; \tilde\gamma_1 \; \cdots \; \tilde \gamma_m \;\big]\trans.
\end{align}
The above performs the reduction of \eqref{eqn:localGCDcompute}, up to multiplicative units, in $O(m)$ symplectic row operations.
We may similarly emulate the reductions of \eqref{eqn:staircaseGCDcompute} and \eqref{eqn:symplecticReduceFirstColumn} in $O(m)$ symplectic row operations, using Lemma~\ref{lemma:Bezout+n} to reduce the computation of greatest common divisors (up to multiplicative unit factors) to performing powers of the operators $C_{j,k}$.
To summarize, using the bound on the diameter of the quadratic unitary graph $G_n$, we may refine the decomposition of symplectic operators in~\cite{HDM05} by substituting an explicit simulation of Euclid's algorithm by a constant-size sequence of symplectic operations.
This substitution provides an upper bound of $O(m^2)$ for a decomposition of an operator $S \in \Sp_{2m}(\Z_n)$, a bound independent of the modulus $n$.
\egroup
\section{Remarks and open problems}
It should be noted that quadratic unitary graphs, while easy to describe, are closely tied to unsolved problems in computational complexity theory.
In particular, testing adjacency in a graph $G_n$ is precisely the quadratic residuacity problem, which has no known efficient algorithms and is considered unlikely to be efficiently solvable (see \eg\ Chapter~3 of \cite{MVvO96}).\footnote{%
It should be noted that because quadratic residuacity can be reduced to integer factoring (by exploiting the Chinese Remainder theorem), and because factoring is solvable in a polynomial number of operations with bounded error with a \emph{quantum} computer~\cite{Shor}, testing adjacency in $G_n$ is also tractible for a quantum computer.}
Because of this, an efficient algorithm (deterministic or randomized) for discovering the shortest path between two vertices in $G_n$ should be considered unlikely.
We may then ask whether there are efficient algorithms for discovering ``short'' paths (having length bounded by a fixed constant) between vertices in $G_n$.
In Section~\ref{sec:perfect}, we provided a partially non-constructive proof that odd holes arise in quadratic unitary graphs $G_n$ which are odd but not a power of a prime $p \equiv 3 \pmod{4}$.
Numerical investigation suggests that, in particular, \emph{five-holes} (odd holes of size five) are very common in those $G_n$ which are not perfect graphs, even when restricting to five-holes involving the arc $0 \arc 1$.
It would be interesting to obtain a classification of all five-holes which occur in the imperfect graphs $G_n$\,.
As we noted in the introduction and in Section~\ref{sec:perfect}, the graphs $G_n$ for $n \equiv 1 \pmod{4}$ prime are also Paley graphs.
Shparlinski~\cite{Shpar06} shows that prime-order Paley graphs these graphs have high \emph{energy} (\ie\ the operator $1$-norm of the adjacency matrix), coming to within a factor of $(1 - \frac{1}{n})$ of the upper bound $\cE_{\max}(n) \,=\, \tfrac{1}{2}n(\sqrt n + 1)$ shown in~\cite{KM01} for graphs on $n$ vertices.
We may ask to what extent this and other properties of circulant Paley graphs generalize for quadratic unitary graphs.
The author would like to thank Charles Matthews, Robin Chapman, and Chris Godsil for helpful discussions.
This work was partly written during a visit to School of Computer Science at Reykjavik University.
This work was performed with financial support by MINOS EURONET and the EURYI scheme.
\bibliographystyle{plainnat}
|
\section{Introduction}
Transport phenomena can be safely enumerated amongst the most important physical processes occurring in both natural and artificial systems, and therefore the transport of charges or energy plays a central role in many scientific disciplines.
Particularly important examples are the life-enabling transport processes in the molecular machinery of biological systems, which take place at scales ranging from a few atoms to large macro-molecular structures such as photosynthetic complexes.
Pioneering studies of energy transfer, both theoretical and experimental, have already appeared in the early stages of the development of quantum mechanics~\cite{perrin32}, and this subject has been of interest ever since.
Recent experimental investigations reveal the presence of long-lived quantum coherence during charge or energy transport processes in photosynthetic complexes~\cite{Brixner05,Engel07,Lee07,Ishizaki09}, charge transport through DNA~\cite{Giese01}, and in polymers~\cite{Collini09}, even at physiological temperatures~\cite{Engel10,Collini10}.
These findings raise the following question: To what extent may a fully quantum-coherent nature of the transport processes (rather than a classically incoherent, diffuse hopping), and perhaps even the presence of stronger quantum signatures such as quantum entanglement~\cite{Horodecki09} be responsible for the remarkable efficiencies that we witness in these systems~\cite{Cai08,Sarovar09,Caruso09a,Scholak09}.
At room temperature, motion of the underlying atomic and molecular structures, along which transport happens, is a ubiquitous phenomenon.
Perhaps more important for living systems is the origin of molecular motion due to driving.
The relevant time-scales range from a few femtoseconds for a stretching mode oscillation of two covalently bound atoms to several picoseconds for the collective modes of larger molecules, and beyond for the conformational changes of proteins~\cite{Schotte03,Frauenfelder09}.
If the transport process and the underlying structure evolve on comparable time-scales, their dynamics do not separate, and we must expect an interplay between both that may affect the transport efficiency.
In the present work, we treat an abstract transport model in which the motion of the molecule backbone structure is classical, and focus on the quantum-coherent transport that takes place on this moving structure.
The quantum degrees of freedom that are carried by the molecular backbone will thus inherit time-dependent features from the classical motion.
We show how such a time dependence can actively drive the coherent excitation transfer and thereby increase its efficiency beyond any comparable static scenario.
The presence of this enhancement is a genuine quantum effect as it does not occur in a classical diffusive hopping transfer.
With this we identify and quantify the motion-induced quantum enhancement of the quantum-coherent transport.
It is known that dynamical structural changes in proteins affect transport processes in various ways.
Environmental noise induced by thermal motion, and specifically dephasing, may lead to an enhancement of the efficiency of transport through a network of sites~\cite{Balabin00,Plenio08,Mohensi08,Caruso09,Semiao09,Chin09}.
There, the generic idea is that dephasing can overcome the trapping of energy due to a disorder of the local energies, or due to destructive interference along different paths through the network (similar as destructive interference in a precisely aligned interferometer).
As an example, such a dephasing process can result from coupling the local energies to the collective vibrations of the network.
The transport process itself may, however, induce a rearrangement of the surrounding molecular structure that acts back on the transport process~\cite{Schulten91,Reddy93,Vos94,Goushcha99,Kriegl04,Wang06}, and may induce a directionality, for example.
In this case, protein motion and transport processes are coordinated as opposed to random changes due to thermal motion.
Indeed, such triggered, long-lived coherent motion of protein nuclei has been found to exist~\cite{Vos93,Vos00}, even at room temperature.
Here, we choose an approach consistent with the latter observation and treat the molecule motion as coordinated with the transfer process rather than as a noise source.
Therefore, in our case the enhancement of transport efficiency has nothing to do with dephasing, but arises from driving, i.e.\ a {\em constructive} interplay between the structural changes of the transport network and the quantum-coherent hopping between the sites.
Since the underling physical mechanisms of single-electron, hole, proton, or spin transport have an identical formal modeling, the results of this study can be straightforwardly applied to the corresponding parameter regimes of these systems as well.
Furthermore, we demonstrate simple and therefore robust control techniques to induce or enhance these effects in linear systems.
\paragraph{}%
In the present work, we shall deal with a rather abstract model to investigate the effects of motion on transport efficiencies.
Nevertheless, several biological applications are conceivable that largely resemble our chosen parameter regimes, and show features that agree with the assumptions underlying our model.
In proteins, it is suggested that energy transport plays a crucial role for conformational changes~\cite{CruzeiroReview,Xie02}.
For the protein to change its shape, and perform its function, it is often necessary to overcome an energetic barrier.
It is hypothesized that the necessary energy needs first to be available as vibrational energy in the protein, and that it is transported from a site where it is provided by ATP to the region within the protein where it is consumed by a conformational change.
A type of secondary structure in proteins, in which the chain of amino-acids forms an $\alpha$-helix, is suggested to provide such a transport pathway for vibrational energy~\cite{CruzeiroReview,Xie02}.
Although the sites that carry the excitation in form of a localized molecular vibration (the amide-I oscillator is predominantly a \mbox{$\mathsf{C=O}$} stretching mode oscillation) are arranged in a helix, the structure is essentially linear.
In this scenario, energy transport has been investigated under the name of the Davydov-Scott-model~\cite{CruzeiroReview,Scott92,Davydov}, to a greater degree of realism than we aim to achieve here, and with an emphasis on localization phenomena of excitations on the $\alpha$-helix due to motion~\cite{Cruz05,Tsivlin05,Tsivlin07}.
For photosynthesis in green sulfur bacteria, light is harvested by a large antenna complex, and the excitation is subsequently transferred via the well-studied Fenna-Matthews-Olson (FMO) protein, that thus functions as a wire, to the reaction center where the excitation energy is converted to chemical energy, for a review see~\cite{Cheng09} and references therein.
Within a single unit of the FMO trimer, the excitation has been shown to propagate quantum-coherently via seven bacteriochlorophyll pigments, entering and exiting at specific pigments~\cite{Brixner05,Engel07,Ishizaki09,Engel10}.
In the reaction center, an electron is then quantum-coherently transferred via several sites.
This electron transfer is accompanied by a structural changes in form of low frequency vibrations of the surrounding protein environment on the same time scale~\cite{Vos93,Vos00,Lee07}.
\paragraph*{}
The paper is organized as follows.
In section~\ref{basic setting}, we introduce the formal model, and start our investigation with the simplest possible case of energy transfer, namely the dimer model that consists of only two molecules.
This serves as a reference system to build intuition of how mechanical oscillations can enhance the energy transfer.
By comparing quantum excitation transfer with the classical F\"{o}rster theory of energy transfer, we explain how mechanical motion can constructively cooperate with the quantum dynamics of the excitation.
In section~\ref{standing oscillations}, we generalize this model to moving systems of many molecules, where we investigate motion modeled by their normal modes and the resulting effects on excitation transfer.
Section~\ref{guided transfer} extends the study of motion of a chain towards a scheme in which the interaction strength between sites is modulated by a pulse.
\section{Transport Model}
\label{basic setting}
Let us first set the formalism to describe excitation transport.
Here, we limit ourselves to a linear chain of $N$ sites, which represents an array of molecules, e.g.~an $\alpha$-helix, among which an excitation can be exchanged due to dipole-dipole couplings between the molecules.
When the probability of an excitation being present in the system is low, we can restrict our analysis to the single-excitation subspace.
Furthermore, we assume for simplicity that the molecules are sufficiently distant from each other such that their interaction can be reduced to the dominant interaction between nearest neighbors.
Generally, however, this is not the necessarily the case, as in the FMO complex, for example.
With these assumptions, the Hamiltonian of the system can be written as the tight-binding Hamiltonian of an interacting $N$-body system in the single-excitation manifold ($\hbar=1$),
\begin{equation}
\label{Hamiltonian}
H=\sum_{n=1}^{N}\varepsilon_{n}\proj{n}{n} + \sum_{n=1}^{N-1}J_n \Big( \proj{n}{n+1}+\proj{n+1}{n} \Big) ,
\end{equation}
where $\ket{n}$ denotes the state with the excitation at the $n$-th site having energy $\varepsilon_{n}$, and $J_n$ is the coupling strength between the $n$-th and $(n+1)$-th molecule, as depicted in Figure~\ref{chainmodel}~(left).
We assume the initial state of the system at time $t=0$ to be a single excitation localized at the first site (left end of the chain in figure~\ref{chainmodel}), i.e.~$\rho(0)=\proj{1}{1}$.
This is in accordance with the idea that the excitation enters the transport network at specific sites as verified in the FMO complex~\cite{Wen09}, or provided locally by ATP in the $\alpha$-helix scenario.
For spectroscopic investigations, fs-pulses of the right wave length are able to excite specific sites with in a complex of chromophores as done in the FMO kind~\cite{Brixner05}.
The motion of molecules will impose a time dependence onto the Hamiltonian~\eqref{Hamiltonian}.
We treat the spatial positions of the chain constituents as classical quantities that follow well-defined trajectories~\cite{Cai08}.
This semi-quantal approximation holds if the involved molecules are too large to show their quantum behavior, i.e.\ if the uncertainty with which the position of a molecule can be determined due to the classical thermal fluctuations is larger than the quantum width of the associated wave function in the position representation (see also~\cite{Cruz97}).
For a molecule of mass $M$ that is attached to a spring with spring constant $K$, and that is in contact with an environment at temperature $T$, the uncertainty of its position due to thermal fluctuations is $\Delta x_{th}^2 = 2 k_{B} T/K$, while the width of the (ground state) wave function is $\Delta x_{qu}^2 = 2 \hbar/\sqrt{M K}$, with $k_{B}$ being the Boltzmann constant, and $\hbar$ the Planck constant.
Therefore, $\Delta x_{th}>\Delta x_{qu}$ if the temperature is high enough, i.e.\ if $k_B T>\hbar\sqrt{K/M}$, and we can focus on classical motion of the excitation carriers.
For example, the typical critical temperature for an $\alpha$-helix structure is 60\,K (as obtained for the mass and spring constant in the effective one-dimensional model: $K\simeq39-58.5$\,N/m, $M\simeq5.7 \times 10^{-25}$\,kg~\cite{Scott92}) which is certainly much below room temperature.
\begin{figure}
\begin{minipage}[c]{0.7\textwidth}
\centering
\includegraphics{Figures/chainmodel.pdf}%
\end{minipage}%
\begin{minipage}[c]{0.3\textwidth}
\centering
\includegraphics[height=4.5cm]{Figures/energyDiagram.pdf}%
\end{minipage}
\caption{Left: Schematic transport model of next-neighbor coupled chain with local dissipation and the sink attached to the last site. Right:~Schematic energy diagram of two electronic states before and after excitation showing how excitation can induce motion of the nuclei that is coordinated with the launch of the excitation propagation in the complex~\cite{Vos94}.}
\label{chainmodel}
\end{figure}
Motion of the individual molecules changes their relative distance in time and thereby modulates the distance-dependent dipolar coupling.
At the same time, deformation of the molecules leads to a change in dipole moments and thus also induces a time-dependence of the coupling.
The latter will be dominant whenever the molecules are tightly embedded in a protein scaffold such as in the FMO complex.
For our purposes, and in order to demonstrate the effect, we only focus on a modulation of $J_n$ due to the change in distance, since we expect a change of dipoles in strength or direction to yield similar results.
For concreteness, we choose a simple form of motion where neighboring molecules change their distance periodically, which holds for small amplitudes in the harmonic regime.
We will later see that for typical times that the excitation spends in the complex only few molecule oscillation periods at the relevant frequencies suffice such that we essentially disregard damping and dephasing of the molecule motion.
For times short compared to the damping of the motion, the distance between site $n$ and $n+1$ is then given by
\begin{equation}
\label{distance}
d_n(t)=d_0 - \big[u_n(t)-u_{n+1}(t)\big]=d_{0} \left[ 1-2a_n\sin(\omega t+\phi_n) \right],
\end{equation}
where $u_n(t)$ is the displacement of the $n$-th molecule, $d_0$ is the equilibrium distance between two neighboring sites, and $a_n$ is the individual sites' relative amplitude of oscillation when they move with opposite phase around their equilibrium position.
This affects the dipolar coupling strengths $J_n$ according to
\begin{equation}
\label{eq:coupling}
J_n(t)=\frac{\tilde{J}_{0}}{[d_n(t)]^3}
=\frac{J_{0}}{[1-2a_n\sin(\omega t+\phi_n)]^3},
\end{equation}
where $\tilde{J}_0$ contains the dipole moments (here assumed to be constant) and physical constants, and we define $J_0=\tilde{J}_0/d_0^3$, which has the unit of an energy.
The extrema of the coupling are $J_{n,\text{min}}$ when the molecules are farthest apart from each other, that is, when $d_n=d_{0}(1+2a_n)$, and $J_{n,\text{max}}$ when their distance reaches the minimum value of $d_n=d_{0}(1-2a_n)$.
The deterministic motion of the molecules imposed by~\eqref{distance} is for a given amplitude, frequency, and phase synchronized with the propagation of the excitation.
This is in contrast to a stochastic motion of the molecules as caused by thermal fluctuations, for example.
It is, however, conceivable that upon the arrival on the excitation in the complex the electron configuration changes and nuclei begin to move in this new potential due to the Franck-Condon principle as shown in figure~\ref{chainmodel}~(right)~\cite{Vos94}.
The wave packet that describes the nuclear motion has been observed to exhibit surprisingly long coherence times in reaction center proteins~\cite{Vos93}, which is consistent with our ansatz to treat positions of the sites classically, and in concert with the initial excitation.
Other influences that motion might have on the dynamics of the transport include the detuning of the excitation energies of the individual sites.
Here, we will mostly ignore the such arising disorder and localization effects.
In a realistic biological context such as excitation transport in light harvesting complexes, these effect are usually not negligible~\cite{Damjanovic02}.
\paragraph{}
In addition to the Hamiltonian evolution, the excitation dynamics takes place in an open quantum system and therefore also suffers from dissipation.
We describe the loss of the energy excitation due to dissipation in Markov approximation by the following Lindblad super-operator:
\begin{equation}
L_\text{diss}\rho=\sum_{n=1}^{N}\gamma_{n}\big[2\sigma_{n}^{-}\rho\sigma_{n}^{+}-\{\sigma_{n}^{+}\sigma_{n}^{-},\rho\}\big].
\end{equation}
where $\sigma_{n}^{+}$ ($\sigma_{n}^{-}$) is the creation (annihilation) operator of the excitation at site $n$, and $\{A,B\}\equiv AB+BA$.
Since realistically each molecule feels a different environment, the local dissipation rates $\gamma_n$ are different in general.
For simplicity and due to so far unavailable experimental data, however, we will later assume them to be equal.
In order to measure how much of the excitation energy is transferred along the chain (and not lost due to dissipation), we introduce an additional site, the sink, representing the final ($N+1$)-th trapping site that resembles a reaction center, for example.
The sink is populated via irreversible decay of excitation from the last site, ~$N$.
This approach for quantifying the energy transfer efficiency follows Ref.~\cite{Plenio08}, and is formally implemented by adding the Lindblad operator
\begin{equation}
L_\text{sink}\rho=\gamma_{S}[2\sigma_{N+1}^{+}\sigma_{N}^{-}\rho\sigma_{N}^{+}\sigma_{N+1}^{-}-\{\sigma_{N}^{+}\sigma_{N+1}^{-}\sigma_{N+1}^{+}\sigma_{N}^{-},\rho\}]
\end{equation}
to the master equation, where $\gamma_{S}$ denotes absorption rate of the sink.
In order to calculate the efficiency of the energy transfer given by the asymptotic population of the sink, it is necessary to integrate the following master equation $(\hbar=1)$:
\begin{equation}
\label{mastereq}
\frac{\partial\rho}{\partial t}=L\rho=i[\rho,H(t)]+L_\text{diss}\rho+L_\text{sink}\rho.
\end{equation}
Since we are interested in the asymptotic population of the sink, which is our figure of merit for the transport efficiency, we search for long-time solutions of the above equation.
Asymptotically in time, the excitation will be either lost to the environment due to $L_\text{diss}$ or, ideally, trapped in the sink due to $L_\text{sink}$.
Note that it therefore does not suffice to merely search for the stationary state, which is characterized by the ground state of the $N$ sites: it thus contains no information about how likely the excitation is trapped in the sink instead of being lost to the environment.
In numerical simulations, we will usually consider uniform local energies $\varepsilon_{n}=\varepsilon$ unless otherwise noted.
Additional effects that arise from static disorder or detuning of the local energies due to the coupling to the nuclear motion are outside the focus of the present work, and are thereby mostly neglected.
Henceforth, all energies, time-scales, and rates will be expressed in units of $J_0$, and thereby we effectively set $J_0=1$.
\subsection*{Dimer Model}
In the remainder of this section, we study the case of a short chain composed of only $N=2$ interacting sites, the dimer, plus the sink in a local dissipative environment for $\varepsilon_1=\varepsilon_2\equiv\varepsilon$.
Although simple, this toy model exhibits all the essential features of larger systems, and gives us the possibility of guiding our intuition towards the mechanisms that underlie the transport processes.
For a time-\emph{independent} coupling $J$, the excitation coherently oscillates between both sites at a constant Rabi frequency until it is either lost into the environment or absorbed into the sink, when it is at site~2.
The asymptotic sink population is given, in this static situation, by~\cite{Plenio08}
\begin{equation}
P_\text{sink}^\text{st}(J)=\frac{\gamma_{S}J^{2}}{(2\gamma+\gamma_S)[\gamma(\gamma+\gamma_S)+J^2]}
\end{equation}
where $\gamma_1=\gamma_2\equiv\gamma$ is the rate of dissipation into the environment at each site, and the superscript ``$\text{st}$'' indicates the static, i.e.\ non-oscillatory, case with constant $J$.
\begin{figure}
\centering
\includegraphics[scale=0.8]{Figures/interaction.pdf}
\includegraphics[scale=0.8]{Figures/sitepopulation.pdf}
\caption{(a) Coupling strength. (b) Population dynamics for coherent transfer in a dimer molecule oscillating at the optimal frequency.
Curves represent the population of site~$2$ (solid), the sum of the populations in site $1$ and $2$ (dashed), and the population of the sink (dotted).
The time axis is rescaled using the oscillation frequency $\omega=4.54$, which corresponds to the maximum in figure~\ref{freqEstimate}(a).}
\label{2sites_optimalFreq}
\end{figure}
In the case of a time-\emph{dependent} coupling $J(t)$, the master equation can be straightforwardly integrated if the combined dissipation rates at both sites are equal, i.e.\ for $\gamma_1=\gamma_2+\gamma_S\equiv\Gamma$, and $\varepsilon_1=\varepsilon_2$ as above. With this choice, the populations of the respective sites read:
\begin{eqnarray}
\label{sitePop1}
P_1(t) =\cos^2\left(\int_0^t J(t^\prime) dt^\prime\right) e^{-2\Gamma t},
\qquad \text{and} \\
\label{sitePop2}
P_2(t) =\sin^2\left(\int_0^t J(t^\prime) dt^\prime\right) e^{-2\Gamma t}.
\end{eqnarray}
The sink population is finally obtained by integration: $P_\text{sink}(t)=2\gamma_S \int_0^t P_2(t) dt$.
Since the phase of the Rabi oscillation is obtained by integration over the coupling strength in \eqref{sitePop1} and \eqref{sitePop2}, we expect a faster propagation of the excitation between sites whenever the coupling strength is stronger, i.e.\ when the two sites are closer to one another.
Similarly, we witness a slower excitation propagation (or even an almost static situation) during times of weak coupling, when the sites are distant.
For a relative amplitude of $a\equiv a_1=1/4$, figure~\ref{2sites_optimalFreq}(a) shows the time-dependence of the interaction strength, which exhibits (for the chosen $a$) short times of strong interaction, separated by longer durations of weak coupling.
This asymmetry in the coupling strength is due to the disproportional dependence of $J(t)$ on the relative distance of the sites, cf.~\eqref{eq:coupling}.
For the right ratio of Rabi frequency and mechanical oscillation frequency $\omega$, the narrow peaks of strong interaction, when the sites are close, can be used to quickly transfer the excitation between the sites, whereas the longer ``valleys'' of low interaction strengths and slow Rabi oscillation are used to effectively lock the excitation on one site, cf.~figure~\ref{2sites_optimalFreq}(b).
Using this quick-transfer-and-locking strategy, we can estimate parameters that allow us to maximize the duration during which the excitation is located at the second site, and thereby most efficiently exposed to the decay into the sink.
Starting with the initial spatial position given by the phase $\phi\equiv\phi_1=\pi/2$ such that the sites are closest, and the interaction is strongest, we need about the first 1/4th of the mechanical oscillation cycle, when the interaction is still sufficiently strong, to transfer the excitation from site 1 to site 2 (in figure~\ref{2sites_optimalFreq}(b) $P_2$ approaches the total population of both sides).
During the next half cycle of the mechanical oscillation, when the interaction strength is low, the site population does not change significantly.
This effectively amounts to a locking of the excitation at site 2.
The following half-cycle, during which the interaction is strongest again, then needs to last long enough for the excitation to move again to site 1 and back.
Comparing with \eqref{sitePop1} and \eqref{sitePop2}, the above translates into the requirement that \mbox{$\int_0^{T/4}J(t^\prime) \rmd t^\prime\simeq\pi/2$} for the first transfer to site~2, where $T=2\pi/\omega$ is the period of the mechanical oscillation.
For a strong interaction modulation as in figure~\ref{2sites_optimalFreq}(a), we neglect the coupling strength when the two sites are distant, and the integral can be approximated by $J_\text{avg}\pi/\omega$, where \mbox{$J_\text{avg}=J_0(1+2a^2)/(1-4a^2)^{5/2}$} denotes the time-averaged coupling strength.
For the above situation, we thereby obtain an optimal frequency of $\omega \simeq 2 J_\text{avg}$.
Higher-order maxima are obtained by adding full Rabi cycles to a transfer step, i.e.\
\begin{equation}
\label{freqEstimateMax}
\int_0^{T/4} J(t^\prime) \rmd t^\prime \simeq \frac{\pi}{2}+m\pi
\quad \textrm{and} \quad
\omega \simeq \frac{2J_\text{avg}}{2m+1}, \quad m=0,1,2,\ldots.
\end{equation}
In contrast, for a locking of the excitation at site 1, i.e.\ for a minimum exposure to the sink decay, an analogous argument yields the minima of the transfer efficiency:
\begin{equation}
\label{freqEstimateMin}
\int_0^{T/4} J(t^\prime) \rmd t^\prime \simeq m\pi
\quad \textrm{and} \quad
\omega \simeq \frac{J_\text{avg}}{m}, \quad m=1,2,\ldots.
\end{equation}
Figure~\ref{freqEstimate}(a) summarizes the transfer efficiencies for a range of mechanical oscillation frequencies.
The estimated frequencies for extremal transfer agree well with the obtained numerical integration, even for the situation in the plot, in which the decay rates do not fulfill the assumption underlying the analytical solution in \eqref{sitePop1} and~\eqref{sitePop2}.
\begin{figure}
\centering
\includegraphics[width=0.6786\textwidth]{Figures/freqEstimate.pdf}
\hfill
\includegraphics[width=0.2714\textwidth]{Figures/amplPhaseDelta.pdf}
\caption{Asymptotic sink population dependent on the mechanical oscillation frequency $\omega$. (a) Initial phase of the mechanical oscillation is $\phi=\pi/2$, i.e.\ the first site of the chain is excited when the sites are closest. Estimated frequencies according to \eqref{freqEstimateMax} and \eqref{freqEstimateMin} for extremal sink population are indicated by vertical lines.
Horizontal lines indicate the asymptotic sink population for static couplings.
For asymptotically large $\omega$, the curve approaches the sink population $P_\text{sink}^\text{st}(J_\text{avg})$.
(b)~Amplitude dependence of the maximal enhancement with an initial phase $\phi=\pi/2$ (solid) and the optimal initial phase (dashed) shows an algebraic decrease for small amplitudes in a double logarithmic plot. The cross marks the amplitude used for the other plots.
}
\label{freqEstimate}
\end{figure}
In case the combined decay rates are not equal at all the sites, the symmetry of the problem is broken, and the eigenstates of the Hamiltonian, the symmetric and anti-symmetric superposition of the excitation being at sites 1 and~2, are coupled to one-another by the decay process.
Moreover, if the initial phase is not $\pi/2$, maxima of $P_\text{sink}(\omega)$ are no longer characterized by a perfect locking of the excitation at site 2 for all times, but only during the first few relevant oscillations until the excitation has essentially left the system.
\paragraph{}
In order to quantitatively demonstrate how the presence of mechanical oscillations genuinely enhances the energy transfer, we compare the asymptotic population of the sink $P_\text{sink}(\omega)$ for a chain oscillating with frequency $\omega$ with the obtained sink population for a fixed chain in the closest and hence strongest coupled site configuration, $P_\text{sink}^\text{st}(J_\text{max})$.
For this purpose we introduce the \emph{motion-induced quantum enhancement}
\begin{equation}
\Delta(\omega)=P_\text{sink}(\omega)-P_\text{sink}^\text{st}(J_\text{max}) .
\end{equation}
We choose a set of parameters for all the following simulations that allow us to conveniently identify the effects. The chosen time scales are approximately of the same order: $J_{0}=1$, $\gamma_n=1/10$, and set $\gamma_{S}=1/2$.
Figure~\ref{freqEstimate}(a) shows that for the moving molecules the sink population $P_\text{sink}(\omega)$ surpasses the maximal static case $P_\text{sink}^\text{st}(J_\text{max})$ by several percent for the chosen parameters, and exhibits an enhancement $\Delta>0$.
\paragraph{}
We emphasize that the motion-induced enhancement beyond the efficiency of the static case at maximal coupling strength is a true quantum effect. That is, the motion-induced efficiency enhancement in the quantum mechanical case is stronger than the motion-induced enhancement in the classical case.
The coherent excitation transfer (Rabi oscillation) can transfer \emph{all} of the excitation to the second site.
A modulation of the coupling strength on a comparable time-scale may then prolong the effective time that the excitation is localized close to the sink as compared to the static chain.
In the corresponding classical case, however, the excitation transfer takes place according to F\"orster theory~\cite{Forster}, and is governed by a rate equation, e.g.\
\begin{equation}
\frac{d P_n(t)}{dt} = \sum_{m=1}^N M_{nm}(t) P_m(t),
\end{equation}
where the population of the $n$-th site $P_n(t)$ depends on the transfer rates $M_{nm}$ for a stochastic hopping between sites $n$ and $m$.
For detailed balance $M_{nm}=M_{mn}$, this leads to a diffusive propagation of the excitation through the network of sites, and in absence of dissipation to an equilibration of the site populations.
The maximal population at the last site that is connected to the sink is therefore at best $P_N(t)=1/N$.
A modulation of hopping rates $M_{nm}(t)$ does only affect the speed with which this equilibration is reached but not the maximal population of the last site.
The mechanical oscillation of the sites which modulates the hopping rates can therefore not achieve a hopping rate above the closest site configuration, where the maximal population of the last site is reached fastest, and maintained highest in presence of dissipation and decay to the sink.
The nature of diffusive transport does therefore not allow for dynamic excitation locking strategies as pictured earlier.
The corresponding quantity $\Delta_\text{cl}(\omega)=P_\text{sink,cl}(\omega)-P_\text{sink,cl}^\text{st}(J_\text{max})$ then measures the efficiency of classical diffusive excitation transport through a moving molecule structure in comparison to classical transport through a static, maximally coupled structure.
We can therefore conclude that a value of $\Delta_\text{(cl)}(\omega)>0$ is classically forbidden.
A positive motion-induced quantum enhancement $\Delta(\omega)>0$ thus proves and quantifies a ``quantum advantage'' due to the quantum nature of the transport process.
\paragraph{}
The observed motion-induced efficiency enhancement is present for all values of amplitudes $a$ and magnitudes of the modulation of the coupling $J$.
However, the effect is less pronounced for smaller values than the ones used here.
The optimum oscillating frequency $\omega$ is then no longer characterized by a complete locking of the excitation at the second site, but rather by a modulation of the Rabi oscillation such that the excitation lingers slightly longer at the last site.
Our particular choice of $a=1/4$ amplifies the visibility of the effect, but it does not cause its presence.
Figure~\ref{freqEstimate}(b) collects the maximal enhancement $\Delta(\omega_\text{opt})$ at the optimal frequency for several values of the amplitude $a$.
Even for reasonably small amplitudes of the displacement, the effect is of the order of a percent.
The doubly logarithmic plot suggests that the effect diminishes only algebraically for amplitudes ranging from large displacements that are expected for a driven motion, down to small position fluctuations as caused by thermal effects.
In $\alpha$-helices the equilibrium distance between the amide-I oscillators is $d_0\simeq4.5$\,\AA{} whereas their relative displacement reaches values of $d_0a\sim0.1$\,\AA~\cite{Scott92,CruzeiroReview}, which amounts to an $a$ of one order of magnitude smaller.
Nevertheless, the coupling $J$ exhibits a variation of $\sim7\%$.
For smaller amplitudes of the mechanical oscillation than the ones chosen to obtain figure~\ref{freqEstimate}, the extrema persist but are less pronounced.
It is therefore conceivable that, at the right frequency of motion, even in biological or chemical systems where oscillation amplitudes are small, the quantum-coherent nature of this transport process on a moving structure yields efficiency improvements beyond what is classically achievable.
Please also note that the total efficiency gain with respect to the equilibrium situation is much larger.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{Figures/enhancementPhaseAvg50.pdf}
\caption{Asymptotic sink population $P_\text{sink}(\omega)$ for different initial phases $\phi$ of the mechanical oscillation.
Thin grey curves correspond to 50 initial phases uniformly spaced in $\phi\in[0, 2\pi]$. Thick black curves indicate the enveloping curves (outer curves), and the phase averaged asymptotic sink population (inner curve) of 200 uniformly sampled phases. Horizontal lines again show the sink population for static cases with $J_\text{max}$, $J_\text{avg}$, and $J_0$ (from top to bottom).
For asymptotically large $\omega$, all curves approach the sink population of the static case with $J_\text{avg}$.
}
\label{2sites_phases}
\end{figure}
So far, the discussion was limited to the motion of the molecules starting with the initial phase~$\phi=\pi/2$, i.e.\ with the molecule in the closest and hence maximally coupled configuration.
This is justified -- as explained earlier and shown in figure~\ref{chainmodel}~(right) -- when the motion is triggered by the excitation of the complex that changes the nuclear equilibrium position to greater distances than in the ground state.
For motion of mostly thermal origin, such a coordination is generally not the case, and the initial phase would be random.
In particular, one cannot expect an enhancement if the molecules are initially in a distant configuration and move away from each other.
We therefore expect that, in the absence of synchronization between excitation transfer and motion, the molecules will not be able to efficiently drive the transfer.
In order to test whether or not an enhancement of the excitation transfer above any comparable static case persists for an arbitrary initial phase of the mechanical oscillation, we statistically sample uniformly over initial phases.
In figure~\ref{2sites_phases}, each thin gray curve represents the sink population and enhancement for a given initial phase of the oscillation, of which the figure~\ref{freqEstimate} represents an instance.
The outer black curves are the envelopes of the data, and represent the maximal and minimal sink population that can be obtained for the chosen parameters at a given oscillation frequency~$\omega$.
The inner black curve shows the average sink population over all initial phases.
We can observe that the average sink population for a random initial phase lies below the static case with coupling $J_0$ for low frequencies, and approaches the efficiency of the static case with the time-averaged coupling strength $J_\text{avg}$ from below for large frequencies.
An enhancement of the transfer efficiency over the static case with $J_0$ can on average still exist simply due to the fact that the average coupling strength is stronger than $J_0$.
Furthermore, it becomes evident that an enhancement $\Delta>0$ can only be reached in a window of frequencies.
For very slow motion, i.e.\ for small~$\omega$, the excitation propagates on a quasi-static structure, and a certain threshold of the underlying motion has to be passed before $\Delta>0$ appears.
For very fast motion, $\omega\gg J_0$ the transport efficiencies converge towards the static configuration with the time-averaged coupling strength and thus $\Delta<0$ (not shown in figure~\ref{2sites_phases}).
We can thus conclude that for a given oscillation frequency $\omega\simeq O(J_0)$, a large enhancement over the static situation can be obtained, but strongly depends on the initial phase.
Therefore, it is crucial that the transport process can be initialized with a specific range of phases of the site oscillation rather than a uniformly distributed random phase.
At the same time, however, if one has an experimental way of setting or influencing the initial phase, the strong phase-dependence might be exploited to switch or to observe this effect experimentally.
\paragraph*{}
In the following sections, we consider longer chains and perform a deeper analysis of the behavior of $\Delta(\omega)$.
We anticipate that the main features identified in the dimer model are qualitatively present also in larger systems, and that the quick-transfer-and-locking strategy can be employed to shuttle the excitation along a longer structure.
In particular, we expect that $\Delta(\omega)$ exhibits an analogous oscillatory behavior that extends to positive values for a specific range of frequencies.
The identification of such optimal frequencies could in principle lead to a better control and to an enhancement of the efficiency of energy transfer in artificially designed systems via the modulation of the oscillation frequency.
\section{Multi-molecule oscillations}
\label{standing oscillations}
In this section, we consider longer linear chains as a first step towards a more suitable representation of nature-inspired systems, such as the amide-I oscillators along an $\alpha$-helix or the chromophores in a light harvesting complex.
The analysis performed in the dimer case is now extended to linear chains of $N$ molecules.
The underlying intuition is that the molecules carrying the excitation are embedded in a larger complex, e.g.\ a protein, but again are assumed to follow a concerted motion.
The particular environmental and boundary conditions change from instance to instance.
For this reason and for simplicity, we model the linear chain as a succession of molecules having the same mass and attached via a spring to their respective neighbors, as e.g.\ inspired by the amide units in an $\alpha$-helix.
For the first and last site we consider two possible situations:
both the ends are (i) completely free or (ii) both are in contact with two ``walls'' via two additional springs.
In the first case, the chain would represent an $\alpha$-helix structure without a narrow confinement in the longitudinal direction, while in the latter case the form of the surrounding protein determines an effective space in which the helix is confined.
For both situations, we study collective normal mode oscillations of low frequency as a relevant example of naturally occurring oscillations.
The transport efficiencies along the chain that are obtained under these dynamical conditions are compared to the ones in the static case, in which all the molecules are at the equilibrium distance.
We thus obtain the gain that is simply due to the presence of motion.
Alternatively, we compare to a configuration where the sites are maximally coupled, which yields~$\Delta$.
In the case of uniformly coupled, static chains of more than three sites, the excitation is no longer completely transferred to the last site with probability one, from where it may be trapped into the sink~\cite{Datta}.
Since, in addition, with increasing chain length the excitation is also more likely to decay into the environment, we expect lower absolute efficiencies than in the dimer toy model.
On the other hand, there is the possibility that the motion-induced enhancement might add up over several transfer step along the chain.
\begin{figure}
\centering
\includegraphics[width=0.45\textwidth]{Figures/sloshing.pdf}\hfill
\includegraphics[width=0.45\textwidth]{Figures/breathing.pdf}
\caption{Left: Sloshing mode of oscillation in which half of the system of two-level sites contracts and half expands alternately. Right:~Breathing mode in which the entire system contracts and expands periodically. The individual sites are coherently coupled, and the sink is connected via irreversible decay to the last site.}
\label{modes}
\end{figure}
Before comparing the asymptotic amounts of population in the sink, we briefly introduce the model that we use to derive the proper motion of the linear chains.
For simplicity, we assume that the forces between two neighboring sites are proportional to their relative displacements from the equilibrium positions:
this approximation, known as the harmonic approximation, holds as long as the displacements are relatively small.
To help in visualizing the system, one may think of the molecules forming the unidimensional lattice as connected by elastic springs (see figure~\ref{modes}).
In the chain of coupled oscillators, the internal force exerted on the $n$-th site is:
\begin{equation}
F_{n}=M\frac{d^{2}u_{n}}{dt^{2}}=K(u_{n-1}-u_{n})+K(u_{n+1}-u_{n}),
\end{equation}
where $K$ is the spring constant (intermolecular force), $M$ is the particle mass, and $u_{n}$ the displacement of the $n$-th site from its equilibrium position.
For the first and last site, we alternatively have only one spring or fix them with additional springs to the confining boundaries.
At this point we neglect the influence of random thermal forces on the particles.
An additional noise term in the position and thus in the coupling strength, as long as it is sufficiently small, will only slightly affect the effective Rabi frequency as compared to the case without noise, since the Rabi frequency it is effectively obtained by integrating over the coupling strength.
The particular shape of $J_n(t)$, and fluctuations of it will therefore not greatly influence the transport efficiency as long as a quick-transfer-and-locking strategy can be achieved.
A chain of $N$ coupled oscillators has a generic solution of the above equations of motion that can be expressed as a linear combination of normal modes.
The normal modes are independent, collective modes of oscillation in which all the sites move with the same periodicity and do not cause the excitation of other oscillatory motions.
Each of these modes is characterized by a frequency and a relative phase between the individual sites.
For the confined chain, the normal mode frequencies and the displacement of the $n$-th site are (see e.g.~\cite{Rosenstock})
\begin{eqnarray}
\label{freqConfined}
\omega_{q} &=2\omega_{0}\sin\left[\frac{q\pi}{2(N+1)}\right],
\qquad \text{and} \\
\label{displConfined}
u_{n}(t) &=\sum_{q=1}^N \frac{a_q d_0}{A_q} \sin\left(\frac{q\pi}{N+1}n\right) \sin(\omega_{q}t+\phi_q),
\end{eqnarray}
respectively, where $\omega_{0}=\sqrt{K/M}$, $A_q=\sin[q\pi/(N+1)]$ is a normalization constant, and $d_0$ is the equilibrium distance between neighboring sites.
For the chain with open ends the expressions are
\begin{eqnarray}
\label{freqOpen}
\omega_{q} &=2\omega_{0}\sin\left(\frac{q\pi}{2N}\right),
\qquad \text{and} \\
\label{displOpen}
u_{n}(t) &=\sum_{q=1}^N \frac{a_q d_0}{B_q} \cos\left[\frac{q\pi}{N}\left(n-\frac{1}{2}\right)\right] \sin(\omega_{q}t+\phi_q),
\end{eqnarray}
with $B_q=\cos[q\pi/(2N)]$.
Figure~\ref{modes} illustrates the normal mode with the lowest frequency for the confined chain, the sloshing mode, and for the open chain, the stretching or breathing mode.
\begin{figure}
\centering
\includegraphics[width=0.7\textwidth]{Figures/FinitModeSample.pdf}
\caption{Motion induced energy transfer gain over the uniformly coupled chain for different normal modes of $N=13$ confined sites. Oscillation amplitude constant is $a_q=1/24$ and $\phi_q=\pi$.
Only the modes $q=1, 2$, and 13 are shown. $P_\text{sink}^\text{st}(J_0)\approx 0.17$, and the flat part of $q=13$ for low $\omega_0$ amounts to a vanishing absolute sink population.}
\label{Nsites_QEnhancement}
\end{figure}
\paragraph*{}
For a more detailed study, we first analyze the effect of oscillations for individual normal modes of the confined chain.
This approach is consistent with the fact that slow coherent motion of the underlying structure is indeed observed in proteins where transport phenomena occur~\cite{Vos93,Vos00}.
We mimic this slow coherent motion by the low frequency collective modes of the chain.
The sloshing mode, the lowest lying of the normal modes, represents a contraction of one half of the system, and a simultaneous expansion of the other half. For higher modes, the system is divided into more parts, each of which can be considered to be in a local sloshing mode.
Figure~\ref{Nsites_QEnhancement} shows the increase of the asymptotic sink population over the static case for selected normal modes of a confined chain of coupled oscillators, when varying~$\omega_0$, i.e.\ spring constant and/or mass of the molecules.
The structural features that we found for the simple dimer case persist, and exhibit frequencies with increased and suppressed transfer efficiency.
For low frequencies, a suppression in the transport efficiency dominates the behavior.
In this frequency regime, the mechanical oscillation is almost static compared to the timescale of the coherent excitation transfer.
The excitation transfer thus happens on a chain with disordered couplings, which lead to localization of the excitation at the beginning of the chain, and hence to a decreased transfer efficiency.
For higher frequencies $\omega_0$, we obtain efficiency maxima that surpass the transport efficiency of the static, ordered chain.
As in the dimer model, these frequencies are in resonance with the wave-like propagation of the excitation through the chain.
For asymptotically large $\omega_0$ all curves converge to the efficiency of the static case with the time-averaged coupling strength between nearest neighbors (not shown in figure~\ref{Nsites_QEnhancement}). Since we fix $a_q=1/24$ for all modes, and the relative distance between two sites is also determined by a position- and mode-dependent prefactor in \eqref{displConfined} and~\eqref{displOpen}, the time-averaged coupling strengths differ between the normal modes, and between different sites of a single mode.
\begin{figure}
\centering
\begin{tabular}{rr}
\includegraphics[width=.5\textwidth]{Figures/BdeltaVsN.pdf} &
\includegraphics[width=.5\textwidth]{Figures/BreathingSampleN12.pdf} \\
\includegraphics[width=.5\textwidth]{Figures/BomgVsN.pdf} &
\includegraphics[width=.5\textwidth]{Figures/BsampleN48exph.pdf}
\end{tabular}
\caption{Left: Maximum energy transfer enhancement over the maximally coupled static chain $\Delta_\text{opt}$ (top) and the corresponding optimal oscillation frequency $\omega_\text{opt}$ (bottom) for the breathing mode for different chain lengths $N$ and oscillation amplitude constants $a\equiv a_1$.
Right:~Frequency dependence ($\omega\equiv\omega_1$) of the energy transfer enhancement as compared to the chain at rest, i.e.\ for~$J_0$, with amplitude constants $a_1=N/12$ (top), $a_1=N/48$ (bottom), and $\phi_1=0$.
The dotted lines include a detuning of the site energies due to the motion with coupling strength $\chi=10$.}
\label{NSites_optimalEnhancement}
\end{figure}
\paragraph*{}
In order to elucidate the dependence of the transport efficiency on the parameters that define the system, we investigate different amplitudes and chain lengths for a single normal mode.
Since the sloshing mode has no configuration where all the sites are closest and hence maximally coupled (half of the chain expands whereas the other half is compressed), we choose the breathing mode of the unconfined chain, where there exists such a configuration.
For the breathing mode, figure~\ref{NSites_optimalEnhancement} collects the data of the maximum enhancement of the transfer efficiency over the maximally coupled static case, i.e.\ $\Delta_\text{opt}=\max_{\omega_1}P_\text{sink}(\omega_1)-P_\text{sink}^\text{st}(J_{n,\text{max}})$, and the corresponding optimal oscillation frequency $\omega_\text{opt}$.
The amplitude constant $a\equiv a_1$ is chosen such that the maximal extension of the chain increases linearly with $N$, and the relative change in length of the entire chain during the mechanical oscillation, $\Delta L/L$, remains constant when changing~$N$.
In particular, the case of $a=N/48$ provides an instance where the relative distance between neighboring molecules is of the order of what is also found in vibrating $\alpha$-helices at room temperature~\cite{Pleiss91}.
\begin{figure}
\centering
\includegraphics[height=5cm]{Figures/deltaVSdephN48.pdf}%
\includegraphics[height=5cm]{Figures/BdephVsN.pdf}
\caption{Left: Decay of the motion-induced transport enhancement $\Delta$ for the breathing mode as a function of the dephasing rate for different chain lengths. Amplitude constant is $a\equiv a_1=N/48$, $\phi_1=0$, at a fixed frequency~$\omega_0$ that is optimal at $\gamma=0$.
Right:~Critical dephasing rate $\gamma_{c}$ dependent on the chain length $N$.}
\label{NSites_dephasing1}
\end{figure}
As for the dimer molecule, a positive $\Delta_\text{opt}$ cannot be achieved in the analogous classical description, and likewise here, for the quantum-coherent transport, positive values exist over a large range of chain lengths.
We observe that, in general, $\Delta_\text{opt}$ decreases with increasing $N$.
This is essentially due to the larger distance that the excitation needs to travel, and hence a proportionally increasing probability that the excitation is dissipated.
We attribute the increase for short chain lengths to the accumulation of the efficiency gain, which is observed for two sites, over several sites.
The decrease of $\omega_\text{opt}$ with increasing chain length qualitatively follows the inverse of the time at which the excitation is (partially) transferred to the last site in a homogeneously coupled chain, i.e.\ with increasing chain length, the time also increases at which the first maximum of the last site's population appears~\cite{Datta}.
We can apply a similarly intuitive argument to motivate the dependency of $\omega_\text{opt}$ on the oscillation amplitude factor~$a$.
When the amplitude factor $a$ of the oscillation is large, the inter-site coupling strength is subject to a large variation and therefore, when the molecules are closest to each other, the excitation propagates very fast such that a quick modulation of coupling strength is necessary to prevent the immediate reversion of the population.
We expect that this non-classical enhancement induced by oscillations can be observed not only for the chains with uniform site energy (i.e.\ with $\varepsilon_{n}=\varepsilon \ \forall n$), but also for systems with local energy disorder.
There, the varying coupling strength may effectively lift the disorder, since the localization depends on the ratio of detuning and coupling strength.
It is also worth mentioning that in the regime where $J_0\ll\gamma_{S}$, i.e.\ for a large decay rate into the sink, a quantum Zeno-type effect slows down, paradoxically, the absorption into the sink by projecting the system with high probability into a state of an unpopulated last site.
Therefore, in this regime, variations of the inter-site coupling strength due to oscillations have no pronounced effect on energy transfer efficiency.
On the other hand, in the regime of a weak sink rate and weak dissipation, the excitation undergoes many cycles before it is finally absorbed, thus amplifying the effect of the oscillations.
In order to test, how a local detuning of the (at equilibrium position) uniform site energies affects the observed transport efficiency enhancement, we add an exciton-vibration coupling term,
\begin{equation}
\label{exphCoupling}
H_\text{ex-vib} = \chi \Big(u_{n+1}(t)-u_n(t)\Big) \proj{n}{n},
\end{equation}
which also appears in $\alpha$-helices~\cite{Scott92}, to the Hamiltonian~\eqref{Hamiltonian}.
The coupling strength $\chi=10$ is chosen such that for $a=1/4$ in the dimer, the energies are of the same order as the exciton coupling between the sites.
Figure~\ref{NSites_optimalEnhancement} (bottom,right) compares the transfer efficiencies with (dotted lines) and without this detuning (solid).
Although the motion-induced efficiency gain suffers due to the detuning of the site energies and the thereby caused localization, the effect persists.
The claim that the observed enhancement is indeed a quantum feature is tested by adding a dephasing environment, which in general leads to the loss of quantum coherence.
This is modeled by an additional term in the master equation~\eqref{mastereq}:
\begin{equation}
\label{dephasing}
L_\text{deph}\rho= \gamma \sum_{n=1}^N \left(2\sigma^+_n\sigma^-_n\rho\sigma^+_n\sigma^-_n -\{\sigma^+_n\sigma^-_n,\rho\}\right)
\end{equation}
Since we expect quantum coherence to be a key requirement on which this effect relies, adding a dephasing environment should lead to a smaller enhancement.
Indeed, the simulations in figure~\ref{NSites_dephasing1} show that the enhancement over the static case with $J_\text{max}$ disappears at a specific critical dephasing rate $\gamma_{c}$. However, an enhancement of the transport efficiency over the static case with $J_0$ still exists, even for large dephasing rates (see figure~\ref{NSites_dephasing2}).
As expected, the critical dephasing rate becomes smaller as the system size increases due to the longer time over which the excitation is exposed to the coherence-destroying action of the environment.
Please note, that in \cite{Plenio08,Caruso09} an enhancement of the transport in the presence of dephasing is observed when adding dephasing to a static, disordered transport network.
In contrast, the enhancement observed here comes from the concerted motion of the molecules, and it is an enhancement with respect to the static case with the same dephasing.
This enhancement exists whether or not dephasing is present.
\begin{figure}
\centering
\includegraphics[height=5cm]{Figures/BsampleN12Deph.pdf}%
\includegraphics[height=5cm]{Figures/BsampleN48Deph.pdf}
\caption{Enhancement of energy transfer efficiency of the oscillating chain (breathing mode) as compared to the transfer efficiency of the resting chain coupled with $J_0$ as both suffer strong dephasing at rate $\gamma=1$ with $a_1=N/12$ (left) and $a_1=N/48$ (right), and $\phi_1=0$.}
\label{NSites_dephasing2}
\end{figure}
\section{Guided excitation transfer}
\label{guided transfer}
In the previous sections, we observed that motion and in particular oscillations can nontrivially interact with the quantum dynamics of the excitation such as to drive and enhance the energy transfer.
We now extend the previous cases to an externally driven scenario.
As a simple example, we investigate the sweeping of excited population guided by a Gaussian pulse that modifies the displacement and hence the interaction between the molecules as it travels along the chain.
A toy model could be implemented in a simulation with trapped ions by shining a laser with a suitable spatial intensity profile onto the chain, for example.
In a molecular context, a conceivable mean could be to attach the molecular chain to a nano-mechanical oscillator.
We model the pulse such that the distance between molecules $n$ and $n+1$, $d_n(t)=d_0-u(t)_{n,n+1}$, is changed from their equilibrium distance by
\begin{equation}
\label{guidedPulse}
u(t)_{n,n+1}=A\,d_0 \exp\left(-\frac{[(n-1) d_{0}-vt]^2}{2\sigma^2}\right),
\end{equation}
where $\sigma$ is the pulse width, and $v$ its velocity.
The displacement due to the pulse remains Gaussian as the pulse uniformly propagates along the chain with its center at the position of site $n$, i.e.~$nd_{0}=vt+d_0$. The effect of this pulse is to concentrate the molecules in a certain region of the chain and thereby couple them more strongly, whereas the molecule density and coupling remains unaffected outside the pulse.
Again, we would like to stress that there exist alternative handles to modulate the coupling strength, for example, a suitable change in the direction of the molecules' dipole moments.
We start the pulse centered around site $n=1$ at time $t=0$, and first observe the moving Gaussian pulse according to~\eqref{guidedPulse} for different values of its speed~$v$.
All lengths are measured in units of $d_0$, such that we effectively set $d_0=1$.
In figure~\ref{popGuided}~(left), we monitor the sink population as a function of time for different values of the pulse speed.
The continuous line shows the reference case without a pulse, i.e.\ for the uniformly coupled chain with sites equally spaced at distance~$d_0$.
Dashed and dotted lines indicate solutions where a pulse is present.
Clearly, there is an effective interplay between the moving Gaussian pulse and the dynamics of excitation transfer.
\begin{figure}
\centering
\includegraphics[width=0.5\textwidth]{Figures/Guided.pdf}%
\includegraphics[width=0.5\textwidth]{Figures/guidedExPhDeph.pdf}
\caption{Left: Sink population for a Gaussian wave packet sweeping along the coupled chain of $N=13$ sites. The solid line marks the reference case without pulse, dashed/dotted lines indicate data for pulses of different speed, but same width $\sigma=1$, and strength $A=1/6$.
Right:~Transport efficiency gain due to the pulse when dephasing at rate $\gamma$ is taken into account. Curves are obtained for the optimal speed for each~$\gamma$.}
\label{popGuided}
\end{figure}
When the speed of the Gaussian packet is zero, which means a stationary packet centered around the first site, it only compresses the distances in the beginning of the chain (but leaves the remaining distances unchanged), and thereby causes stronger, but disordered couplings.
Although a stronger coupling increases the speed of the excitation transfer between two sites, for $v=0$ the efficiency of energy transfer, i.e.\ the sink population in the long-time limit, is less than that of a uniform chain without a pulse being present (solid line in figure~\ref{popGuided} (left)).
In fact, when the couplings at the beginning of the chain are stronger than between the remaining sites, they cause a localization of the excitation population at the beginning, due to the caused disorder in the couplings.
Therefore, only a relatively small fraction of the population succeeds to reach the sink, whereas the localized population is finally dissipated into the environment.
With this suppression, we once again face a distinct feature of coherent energy transport. In the classical, diffusive energy transfer, an increase of the hopping rates always causes an increase of the transfer rates, and consequently leads to a higher energy transfer efficiency.
If the pulse moves, it will sweep along the excitation that is localized within the pulse.
In the limit of very distant sites, a good strategy is to apply a strong pulse that couples only two sites and moves at the speed given by the Rabi frequency, such that after a full transfer to the second site, the pulse moves and couples site two and three.
For a pulse that is too fast, only part of the excitation will be transferred to the next site, whereas if it is too slow, part of the excitation oscillates back onto the first site again.
In the present case, we cannot neglect the coupling outside the pulse.
Instead, we need to coordinate the two processes of localizing the excitation within the moving pulse, and the wave-like coherent transfer in the uniformly coupled chain.
Figure~\ref{popGuided}~(left) shows that, for the given parameters and a chain length of 13~sites, the dominant contribution to the sink population due to the wave-like spreading of the excitation arrives between times 6--8 (in units of $1/J_0$), with small additions at later times.
The standing wave packet (dotted line) localizes the excitation at the beginning of the chain, but a fraction still succeeds to arrive in the sink at the same time as in the uniformly coupled chain.
The increase of the sink population starts earlier than without the pulse being present, because the pulse also slightly increases the effective coupling strength for the rest of the chain, which leads to a faster propagation of the excitation.
A wave packet moving with the speed of one site per time unit consequently arrives at the end of the chain after having passed 12 sites at time~$t=12$, where it leaves part of the excitation that it has swept along, resulting in an additional increase of the sink population.
At the optimal speed of $v_\text{opt}=2.53\pm0.01$, the wave packet arrives at about the same time as the excitation would in the uniformly coupled chain, hence effectively localizing the excitation in the front of the wave-like propagation.
The optimum speed is proportional to the coupling strength, i.e.\ it increases linearly with~$J_0$, because the Rabi frequency that is given by the coupling strength (up to units) is the speed at which an excitation moves between two sites during the Rabi oscillation. Therefore, the pulse has to match the speed given by~$J_0$.
\begin{figure}
\centering
\includegraphics[width=0.5\textwidth]{Figures/GaussianUnnorm.pdf}%
\includegraphics[width=0.5\textwidth]{Figures/GaussianContour.pdf}
\caption{Energy transfer enhancement for a Gaussian pulse over the uniformly coupled chain with $J_\text{max}$. Chain length is $N=13$, and $A=1/6$. Contours of the surface (left) are plotted on the right.}
\label{packetParameters}
\end{figure}
The simulation is summarized in figure~\ref{packetParameters}, which shows the increase of the sink population for a Gaussian pulse as compared to the uniformly coupled chain, $\Delta=P_\text{sink}(v,\sigma)-P_\text{sink}^\text{st}(J_\text{max})$, i.e.\ at a fixed maximum intensity of the pulse.
For small pulse widths and low velocities the gradient of coupling strengths changes quickly, and influences only few sites.
Therefore, the efficiency suffers most due to the strong localization of the excitation within the pulse.
On the other hand, for large values of $\sigma$ the wave packet is broad, that is, it covers a larger region of the chain, and the gradient of coupling strengths changes more smoothly, which leads to less details in the velocity dependence.
In the limit of large pulse widths, the wave packet virtually covers the entire chain such that the reference case with $J_\text{max}$ is reached, and the efficiency gain approaches zero.
We also found that there is a global optimum ($v_\text{opt}=3.34\pm0.01$, $\sigma_\text{opt}=4.20\pm0.05$) where, at a given maximal intensity, the velocity and pulse width match best the coherent dynamics and achieve an increase of the transfer efficiency of more than $0.03$ over the static chain with maximal coupling strength. Note that the local optimum from figure~\ref{popGuided} with $\sigma=1$ corresponds to the maximum of the front-left cut of the surface in figure~\ref{packetParameters}~(left).
Finally, we also account for a possible dephasing, as caused by fast thermal noise, for example, which is formally implemented by adding the term~\eqref{dephasing} to the Lindblad equation.
The results are collected in the right panel of figure~\ref{popGuided}, and show that even under substantial dephasing the gain due to the presence of the pulse persists, and may even attain values in the classically forbidden region of enhancements $\Delta>0$ (not shown).
\section{Concluding remarks}
In this work, we demonstrate general features of energy transfer efficiency in mechanically oscillating systems, and relate our model to the biological scenarios of coherent energy transport in proteins.
Complementary to previous works, we focus on the driving of the excitation dynamics due to motion of the underlying molecular structure.
The time dependence of the inter-site coupling arises from its distance dependence together with relative motion of the sites, for example.
We generally find a motion-induced quantum enhancement of the excitation transfer efficiency over the equilibrium configuration of the linear chain, and furthermore over any static configuration that is met during the mechanical motion.
This effect distinguishes the quantum-coherent transport from the classical, diffusive transport.
If biological systems manage to utilize the modulation of coherent couplings by their motion, they can profit from the resulting enhancement of the transfer efficiency.
It would be interesting to extend our ideas beyond the linear setting to the enhancement of excitation transfer in light harvesting complexes in photosynthesis or light harvesting complexes in artificial devices.
\ac
Discussions with Leonor Cruzeiro and Victor Atanasov at the Workshop QuEBS 2009 in Lisbon, Portugal, are gratefully acknowledged.
We also thank J\"org Matysik for stimulating discussions and for bringing several references on transport in bacterial reaction centers to our attention.
This work was supported in part by the Austrian Science Fund (SFB FoQuS, J.M.C. through Lise Meitner Program) and the European Union (SCALA, NAMEQUAM)\@.
S.P.~acknowledges support from the UK EPSRC through the IRC-QIP.
\section*{References}
|
\section{Introduction}
Let $S \subseteq \mathbb{R}^d$ be a basic closed semialgebraic set, say $$S = \setcond{x \in \mathbb{R}^d}{p_1(x) \ge 0, \ldots, p_k(x) \ge 0} =: \{p_1\ge 0,\ldots,p_k \ge 0\},$$
where $p_i \in \mathbb{R}[X]$, $X := X_1,\ldots,X_d$.
It is known since the eighties that one can choose, for the description of $S$, polynomials $p_1,\ldots,p_k$ such that $k \le \frac{d(d+1)}{2}$ (compare \cite{MR1137812}). Scheiderer \cite{MR1005003} gave examples showing that this bound is sharp. However, in these examples $S$ admits points $x$ where the local dimension of $S$ at $x$ is $m$ for all $1 \le m \le d$. So one might ask, if the bound for $k$ equals $d$ for sets $S$ of constant local dimension. In her diploma thesis A.~Pauluhn \cite{Pauluhn90} showed that for $d \in \{2,3\}$ equal dimensional basic closed sets can be characterized by at most $2$ and $4$ polynomials, respectively. All this holds true, if $\mathbb{R}$ is replaced by an arbitrary real closed field $R$. It seems that more is not known. Several authors (see \cite{Bernig98}, \cite{Groetschel-Henk-2003}, \cite{Bosse-Groetschel-Henk-2005}, \cite{BosseDiss}) looked at the case where $S$ is a polytope, which might be interesting for applications (see also \cite{Henk-2007} for a survey on this topic). Also, they tried to find effective computations for suitable polynomials $p_i$ with $i=1,\ldots, s$, satisfying $S = \{p_1 \ge 0,\ldots, p_s \ge 0\}$ starting from the description $S= \{l_1 \ge 0, \ldots, l_k \ge 0\}$, where $l_i$ are linear forms (i.e., polynomials of degree at most one) and $k$ might be very large. Let $s=s(S)$ be the minimal possible value as above. One achieved the bound $s \le 2d-1$ in \cite{Bosse-Groetschel-Henk-2005}. In \cite{Groetschel-Henk-2003} one noticed that $s \ge d$ for polytopes and in \cite{Bosse-Groetschel-Henk-2005} one conjectured that $s=d$ for $d$-dimensional polytopes. The equality $s = d$ was shown in \cite{Bernig98} for polygons, in \cite{Averkov-Henk-2009} for simple polytopes and in \cite{Averkov-Henk-2009b} for three-dimensional polyhedra. The following two theorems are the main results of the manuscript. Theorem~\ref{av:thm} below presents a short proof of a generalization of the result for simple polytopes from \cite{Averkov-Henk-2009}. Theorem~\ref{every polyhedron is representable} computes $s$ for all $d$-dimensional polyhedra.
\begin{theorem} \label{av:thm} Let $S \subseteq \mathbb{R}^d$ be bounded and basic closed, say $S= \{q_1 \ge 0,\ldots,q_k \ge 0\}$, where $q_1,\ldots,q_k \in \mathbb{R}[X]$. Let $s \in \mathbb{N}$ be such that for each $x \in S$ there are at most $s$ polynomials $q_i$ among $q_1,\ldots,q_k$ where $q_i(x) =0.$ Then the following statements hold.
\begin{enumerate}[a)] \item \label{av:thm:a} $S=\{p_1,\ldots,p_{s+1} \ge 0\}$ for suitable $p_1,\ldots,p_{s+1} \in \mathbb{R}[X].$
\item \label{av:thm:b} If there are only finitely many points $x_1,\ldots,x_m \in S$ where exactly $s$ polynomials $q_i$ vanish, then $S = \{p_1 \ge 0,\ldots,p_s \ge 0\}$ for suitable $p_1,\ldots,p_s \in \mathbb{R}[X].$
\end{enumerate}
\end{theorem}
In view of Theorem~\ref{av:thm}, every simple $d$-dimensional polytope can be represented by $d$ polynomial inequalities. The above statement is also covered by the following
\begin{theorem} \label{every polyhedron is representable}
Let $S$ be a $d$-dimensional polyhedron in $\mathbb{R}^d$. Let $k$ be the maximal dimension of an affine space contained in $S.$ Then there exist $d-k$ polynomials $p_0,\ldots,p_{d-k-1}$ such that $S = \{p_0 \ge 0,\ldots,p_{d-k-1} \ge 0\}$. Furthermore, $S$ cannot be represented by less than $d-k$ polynomials.
\end{theorem}
In Section~\ref{separation} we present several separation theorems for semialgebraic sets. In Section~\ref{sect:simple:semialg} we prove Theorem~\ref{av:thm}. Finally, in Section~\ref{sect:polytopes} we prove Theorem~\ref{every polyhedron is representable}.
As in the above mentioned papers dealing with polynomial representations of polytopes, our work is semi-effective. That means, one has to check sequences of first order statements. As a consequence, in our theorems we can only compute the polynomials $p_i$ but one cannot bound their degrees in terms of the complexity of the ``input polynomials''. Equivalently, what we do below does not work over any real closed field.
We shall use the following notations. The Euclidean norm of $\mathbb{R}^d$ is denoted by $\|\,\cdot\,\|$. For the Euclidean topology of $\mathbb{R}^d$ we denote by $\rmcmd{cl}$, $\rmcmd{int}$, $\rmcmd{bd}$ the closure, interior and boundary, respectively. We write $\rmcmd{cl}^Z$ for the \term{Zariski closure}. Furthermore, the notations $\dim,$ $\rmcmd{aff},$ and $\rmcmd{relint}$ stand for dimension, affine hull, and relative interior, respectively. For $x \in\mathbb{R}^d$ and $\rho>0$ let $B(x,\rho):= \setcond{y \in \mathbb{R}^d}{\|x-y\| \le \rho}$ and $\mathcal{U}(x,\rho) := \setcond{y \in \mathbb{R}^d}{\|x-y\| < \rho}$.
We write as before $\{p_1 \ge 0,\ldots,p_k \ge 0\}$ instead of $\setcond{x \in \mathbb{R}^d}{p_1(x) \ge 0,\ldots,p_k(x) \ge 0}$ for polynomials $p_1,\ldots,p_k \in \mathbb{R}[X]$. Similarly we write $\{p_1>0,\ldots,p_k>0\}$ and $\{p_1=0,\ldots,p_k=0\}$. Note that $V:=\{p_1=0,\ldots,p_r=0\} = \{q=0\}$ for $q=p_1^2 + \cdots + p_r^2$. $q$ is also called \term{positive polynomial} for the algebraic set $V$.
\section{Separation} \label{separation}
We shall use standard inequalities on continuous semialgebraic functions, see \cite[Section~2.6]{MR1659509}.
\begin{theorem} \label{more than loj}
Let $S \subseteq \mathbb{R}^d$ be a closed semialgebraic set and let $f, g, h$ be continuous semialgebraic functions on $S$ with $\{g=0\} \cap S \subseteq \{f=0\} \cap S.$ Then there exists a positive polynomial $p$ and $N \in \mathbb{N}$ such that $|f^N h| \le |p g|$ on $S$.
\end{theorem}
The version for $h(X)=1$ follows from the results in \cite[Section~2.6]{MR1659509}. The version for $h(X)=1$ obviously implies the version for a general $h$.
For the special case that $S$ is bounded and $h(X)=1$ we may obviously choose $p$ to be a constant, which yields the well-known \term{H\"ormander-{\L}ojasiewicz Inequality}. The polynomial $p$ can be chosen to have a specific form $p(X) = (1+ \|X\|^2)^M$ with $M \in \mathbb{N}.$ Some consequences of Theorem~\ref{more than loj} (and, more specifically, H\"ormander-{\L}ojasiewicz's Inequality) will be useful in our subsequent derivations.
\begin{proposition} \label{f^N>g}
Let $A$ be an unbounded, closed semialgebraic set and $f, g$ be continuous semialgebraic functions on $A$ such that $f > 0$, $g \ge 0$ and $f(x) \rightarrow \infty$ as $x \in A$ and $x \rightarrow \infty$. Then there exists $\gamma > 0$ and $N \in \mathbb{N}$ such that $\gamma f^N \ge g$ on $A$.
\end{proposition}
\begin{proof}
Let $A = A_0 \cup A_1$, where $A_0, A_1$ are semialgebraic, closed, $A_0$ is bounded, $o \not\in A_1$ and $f \ge 2$ on $A_1$.
By Theorem~\ref{more than loj}, $g \le \alpha (1+ \|X\|^2)^M$ on $A$ for appropriate $\alpha >0$ and $M \in \mathbb{N}$. The inequality
\begin{equation} \label{f^N grows fast}
f^N \ge \alpha (1+ \|X\|^2)^M
\end{equation}
is fulfilled on $A_1$ if
\[
f\left(\frac{X}{\|X\|^2}\right)^{-N} \le \frac{\|X\|^{2M}}{\alpha (\|X\|^2+1)^M}
\]
for all $X$ with $\frac{X}{\|X\|^2} \in A_1$.
Let $\Tilde{A}_1:= \{o\} \cup \setcond{x \in \mathbb{R}^d \setminus \{o\}}{\frac{x}{\|x\|^2} \in A_1}$ and $\tilde{f}$ a semialgebraic function on $\Tilde{A}_1$ given by
\[
\Tilde{f}(x):=\begin{cases}
f\left(\frac{X}{\|X\|^2}\right)^{-1} & \mbox{if} \ x \in \Tilde{A}_1 \setminus \{o\}, \\
0 & \mbox{if} \ x=0.
\end{cases}
\]
Clearly, $\Tilde{A}_1$ is bounded and closed. Thus, for verification of \eqref{f^N grows fast} we need to show
\[
\Tilde{f}^N \le \frac{\|X\|^{2M}}{\alpha (\|X\|^2+1)^M} \qquad \mbox{on $\Tilde{A}_1$}.
\]
The existence of $N$ satisfying the above relation follows from {\L}ojasiewicz's inequality (by taking into account that $\tilde{f} \le \frac{1}{2}$ on $\Tilde{A}_1$).
The assertion follows by defining a $\gamma>1$ such that $\gamma f^N \ge g$ on $A_0$.
\end{proof}
\begin{definition} Let $S, T \subseteq \mathbb{R}^d$ be semialgebraic sets and let $p \in \mathbb{R}[X].$ We say that $p$ \term{separates} $S$ from $T$ if $p\ge 0$ on $S,$ $p \le 0$ on $T$, and $\{p=0\} \cap (S\cup T) \subseteq \rmcmd{cl}^Z(S \cap T).$
\end{definition}
The polynomials from Lemma~\ref{lambda:mu:lem} will be frequently used in our constructions.
\begin{lemma} \label{lambda:mu:lem}
Let $0< \delta < \rho$ and $m \in \mathbb{N}$. Then the following two statements hold true.
\begin{enumerate}[a)]
\item \label{kappa:part} There exists a polynomial $\kappa=\kappa_{\delta,\rho,m} \in \mathbb{R}[t]$ such that
\begin{enumerate}[i)]
\item $\kappa \ge 0$ on $[-\rho,\rho]$ and $\{\kappa = 0\} \cap [-\rho,\rho] = \{0\}$,
\item $\kappa \le \frac{1}{4^m}$ on $[-\rho,0]$,
\item $\kappa \ge 2$ on $[\delta,\rho]$,
\item $\kappa \le 3$ on $[0,\rho]$.
\end{enumerate}
\item \label{mu:part} There exists a polynomial $\mu :=\mu_{\delta,\rho} \in \mathbb{R}[t]$ such that
\begin{enumerate}[i)]
\item $\mu>0$ on $[-\rho,\infty[,$
\item $\mu<\frac{1}{2}$ on $[-\rho,0],$
\item $\mu>2$ on $[\delta,\infty[.$
\end{enumerate}
In particular, $\mu(t) \rightarrow \infty$ for $t \rightarrow \infty$
\end{enumerate}
\end{lemma}
\begin{proof}
\ref{kappa:part}) Consider a continuous function $\phi : [-\rho,\rho] \rightarrow \mathbb{R}$ such that $\frac{1}{4^{m+1}} \le t^2 \phi(t) \le \frac{1}{2 \cdot 4^m}$ for $t \in [-\rho,0]$, $t^2 \phi(t) \ge 2 + \frac{1}{3}$ for $t \in [\delta,\rho]$ and $t^2 \phi(t) \le 3 - \frac{1}{3}$ for $t \in [0,\rho]$. Now by explicit Stone-Weierstrass Approximation (cf. \cite[Theorem~8.8.5]{MR1659509}) of $\phi(t)$ by a polynomial $\pi(t)$ we get $\kappa(t) = t^2 \pi(t)$.
\ref{mu:part}) We set
$$
\mu_{\delta,\rho}(t) := \left( \frac{t+ 2 \rho - \delta /2}{2 \rho}\right)^{2k},
$$
where $k \in \mathbb{N}$ is sufficiently large.
\end{proof}
\begin{proposition} \label{sep:disjoint} \thmtitle{Separation of disjoint closed sets}
Let $S, T \subseteq \mathbb{R}^d$ be closed semialgebraic sets such that $S \cap T = \emptyset.$ Let $S$ be compact and basic closed, say $S = \{f_1 \ge 0,\ldots,f_k \ge 0\}$, where $f_1,\ldots,f_k \in \mathbb{R}[X]$. Then there exists a polynomial $p \in \mathbb{R}[X]$ which separates $S$ from $T.$
\end{proposition}
\begin{proof}
We define the polynomial mapping $F(X):=(f_1(X),\cdots, f_k(X))$ from $\mathbb{R}^d$ to $\mathbb{R}^k$. Then $F(S) \subseteq \mathbb{R}_{\ge 0}^k$ and $F(T) \subseteq \mathbb{R}^k \setminus \mathbb{R}_{\ge 0}^k$ are compact resp. closed semialgebraic sets. Choose $\rho>0$ such that $F(S) \subseteq [0,2 \rho]^k$. We define the polynomial $g_m(Y):= \rho^{2 m} k + \frac{1}{m} - \sum_{i=1}^k (Y_i-\rho)^{2 m}$ in indeterminates $Y:=Y_1,\ldots,Y_k$, where $m \in \mathbb{N}$. The semialgebraic set $G_m:= \{g_m \ge 0\} \subseteq \mathbb{R}^k$ approximates $[0,2 \rho]^k$ with any given precision (in the Hausdorff metric), as $m \rightarrow \infty$. Moreover, $[0,2\rho]^k$ is contained in the interior of $G_m$ for every $m \in \mathbb{N}$. Consequently, $G_m$ is disjoint with $F(T)$ if $m$ is sufficiently large. Hence we may define $p:=g_m \circ F$ with $m$ sufficiently large.
\end{proof}
\begin{remark}
Proposition~\ref{sep:disjoint} still holds, if $S$ is not necessarily basic. Also, it can be shown directly by Stone-Weierstrass Approximation, but the way we did it is more constructive.
\end{remark}
\begin{proposition} \label{merge:loc:glob} \thmtitle{Globalizing a local separator}
Let $S \subseteq \mathbb{R}^d$ be bounded and closed semialgebraic set and let $T \subseteq \mathbb{R}^d$ be a closed semialgebraic set. Let $r \in \mathbb{R}[X]$ be such that $r \ge 0$ on $\mathbb{R}^d$, $S \cap T = \{r =0 \}$ and $r(X) \rightarrow \infty$ as $\|X\| \rightarrow \infty$. Let $f \in \mathbb{R}[X]$ be such that $f$ separates $S \cap U$ from $T \cap U$ for an open, semialgebraic set $U$ with $S \cap T \subseteq U$. Then there exists a polynomial $p$ separating $S$ from $T$.
\end{proposition}
\begin{proof}
Since $r(x) \rightarrow \infty$ as $\|x\| \rightarrow \infty$, the set $\{r \le \varepsilon\}$ is bounded for every $\varepsilon \ge 0$. If $0 < \varepsilon \le \frac{1}{2}$ is small enough, then $\{r \le \varepsilon\} \subseteq U$. In view of Proposition~\ref{sep:disjoint}, there exists a polynomial $g$ that separates $S \cap \{r \ge \frac{\varepsilon}{4}\}$ from $T \cap \{r \ge\frac{\varepsilon}{4}\}$. We define $p:=f+ \left(\frac{2 r}{\varepsilon}\right)^N g$ with $N \in \mathbb{N}$ to be chosen below. In view of \L{o}jasiewicz's inequality we have $\frac{1}{2} |f| \ge \left|\frac{2 r}{\varepsilon}\right|^N |g|$ on $( S \cup T) \cap \{r \le \frac{\varepsilon}{4}\}$ for all sufficiently large $N$. On the other hand, by Proposition~\ref{f^N>g}, $\frac{1}{2} \left| \frac{2 r}{\varepsilon}\right|^N |g| \ge |f|$ on $(S \cup T) \cap \{r \ge \varepsilon\}$ for all sufficiently large $N$. Thus, $p$ separates $S$ from $T$ if $N$ is sufficiently large.
\end{proof}
\begin{proposition} \label{merge local sep}
\thmtitle{Merging local separators} Let $S, T \subseteq \mathbb{R}^d$ be closed semialgebraic sets and let $x_1,\ldots,x_m \in \mathbb{R}^d$. Assume that there exists $\rho>0$ such that for every $i \in \{1,\ldots,m\}$ one has $\{x_i\} = S \cap T \cap B(x_i,\rho)$ and one can find a polynomial $p_i$ separating $S \cap B(x_i,\rho)$ from $T \cap B(x_i,\rho)$. Then there exist $\rho>0$ and $p \in \mathbb{R}[X]$ such that $p$ separates $S \cap U$ from $T \cap U$ for $U:= \bigcup_{i=1}^m B(x_i,\rho)$.
\end{proposition}
\begin{proof}
Let $r_i:=\|X-x_i\|^2$. Choose $\rho>0$ to be small enough and a $\delta$ such that $\rho < \delta < \|X_i-X_j\| - \rho$ for $1 \le i < j \le m$. Then, by {\L}ojasiewicz's inequality, for a sufficiently large $N \in \mathbb{N}$ one has $\left(\frac{r_i}{\delta^2}\right)^N \le \frac{1}{2} |p_i|$ on $(S \cup T) \cap B(x_i,\rho)$ and $p_i + \left(\frac{r_i}{\delta^2}\right)^N > 0$ on $B(x_j,\rho)$ for $1 \le i < j \le m$. Thus, we may choose $$p:= \prod_{i=1}^m \left( p_i + \left(\frac{r_i}{\delta^2}\right)^N\right).$$
\end{proof}
The following proposition is a straightforward consequence of Propositions~\ref{merge:loc:glob} and \ref{merge local sep}.
\begin{proposition} \label{sep:finite:intersect} \thmtitle{Separation of sets with finite intersection} Let $S, T \subseteq \mathbb{R}^d$ be semialgebraic sets, $T$ closed, $S$ basic closed and bounded, such that $S \cap T$ is finite, say $S \cap T = \{x_1,\ldots, x_m\}.$ Then there exists a polynomial $p$ which separates $S$ from $T$ if and only if this holds locally, that means: There exists $\rho>0$ such that for $i \in \{1,\ldots,m\}$ there exists a polynomial $p_i$ which separates $B(x_i,\rho) \cap S$ from $B(x_i,\rho) \cap T.$
\end{proposition}
\begin{remark}
The following generalization of Proposition~\ref{sep:finite:intersect} holds true. Let $S, T \subseteq \mathbb{R}^d$ be semialgebraic, $S$ compact, $T$ closed. Then there exists a polynomial $p$ which separates $S$ from $T$ if and only if for any finite set of points $X$ there exists $\rho>0$ and a polynomial $q$ such that $q$ separates $S \cap U$ from $T \cap U$ for $U:=\left( \bigcup_{x \in X} B(x,\rho) \right)$. This follows from Proposition~6.10 (Chapter~VI) in \cite{MR1393194}, whose proof, however, is not at all constructive.
\end{remark}
\begin{proposition} \label{S:T1:T2} Let $S, T_1, T_2 \subseteq \mathbb{R}^d$ be semialgebraic sets, where $S$ is basic closed, $T_1$ is compact and $T_2$ is closed. Let $h$ be a non-negative polynomial with $\{h=0\} = \rmcmd{cl}^Z(S \cap T_1)$ and let $f \in \mathbb{R}[X]$ be an arbitrary polynomial. Assume that
\begin{enumerate}[a)]
\item \label{08.07.28,11:48} $S \cap T_2 = \emptyset,$
\item \label{08.07.28,11:49} $\rmcmd{int} (S \cap T_1)=\emptyset$,
\item \label{08.07.28,11:50} $h>0$ on $T_2$ and, for $x \in T_2,$ $h(x) \rightarrow \infty$, as $\|x\| \rightarrow \infty.$
\item \label{08.07.28,11:51} $\{f =0\} \cap T_1 \subseteq S \cap T_1.$
\end{enumerate}
Then there exists $p \in \mathbb{R}[x]$ such that
\begin{enumerate}[i)]
\item \label{08.07.28,11:52} $p \ge f$ on $S,$
\item \label{08.07.28,11:53} $\rmcmd{sign} p \le \rmcmd{sign} f$ on $T_1,$
\item \label{08.07.28,11:54} $p< 0$ on $T_2.$
\end{enumerate}
\end{proposition}
\begin{proof}
By assumption \ref{08.07.28,11:50}) there exists $\alpha>0$ such that $h\ge \alpha$ on $T_2.$ Choose $\delta$ with $0 < 2 \delta < \alpha$ and $\delta < \rho.$ We introduce the polynomial $\mu=\mu_{\delta,\rho}$ as in Lemma~\ref{lambda:mu:lem}\ref{mu:part}). Let $T_0 := T_1 \cap \{h \ge \delta \}.$ Then $S \cap T_0 = \emptyset.$ Taking into account assumption \ref{08.07.28,11:48}) and applying Proposition~\ref{sep:disjoint}, we can find a polynomial $q \in \mathbb{R}[X]$ with $q>0$ on $S$ and $q<0$ on $T_0 \cup T_2.$ Now, set $p:= f + q \cdot h^l \cdot [\mu \circ (h-\delta)]^{l+m}$, where $l, m \in \mathbb{N}$ are sufficiently large.
\end{proof}
\section{Small polynomial description of special semialgebraic sets} \label{sect:simple:semialg}
If $a_1,\ldots,a_k \in \mathbb{R}$, let $\sigma_i(a_1,\ldots,a_k)$ denote the $i$th \term{elementary symmetric} function of $a_1,\ldots,a_k$, that is,
$$\sigma_i(a_1,\ldots,a_k) := \sum_{1 \le j_1 < \ldots < j_i \le k} a_{j_1} \cdots a_{j_i}.$$
The basic observation is the following.
\begin{lemma} \label{el:sym:lemma}
Let $\rho > 0$ and $s, k \in \mathbb{N}$ with $s \le k.$ Then there exists $\varepsilon>0$ such that for all $a_1,\ldots,a_k \in \mathbb{R}$ with $|a_1| \le \varepsilon, \ldots, |a_s| \le \varepsilon$ and $a_{s+1} \ge \rho, \ldots, a_k \ge \rho$ one has $a_1,\ldots,a_s \ge 0$ if and only if $$\sigma_{k-s+1}(a_1,\ldots,a_k) \ge 0, \ldots, \sigma_{k}(a_1,\ldots,a_k) \ge 0.$$
\end{lemma}
\begin{proof}
The necessity is trivial. Let us prove the sufficiency. Consider the polynomial
$$
\prod_{i=1}^k (t+a_i) =\sum_{i=1}^k \sigma_i(a_1,\ldots,a_k) t^{k-i} + t^k
$$
For $i \le k-s$ there is a summand in the definition of $\sigma_i(a_1,\ldots,a_k)$ without factors belonging to $\{a_1,\ldots,a_s\}$. The summands containing a factor of $\{a_1,\ldots,a_s\}$ tend to zero for $\varepsilon \rightarrow 0.$ Hence for all sufficiently small $\varepsilon>0$ depending on $k, s, \rho$ and for $i \le k-s$ one has $\sigma_i(a_1,\ldots,a_k) >0.$ One has $\sigma_i(a_1,\ldots,a_k) \ge 0$ for $i > k-s,$ by the assumption. Consequently, the polynomial $\prod_{i=1}^k (t+a_i)$ has no positive roots, which implies that $a_1 \ge 0,\ldots,a_k \ge 0.$
\end{proof}
\begin{proof}[Proof of Theorem~\ref{av:thm}]
\ref{av:thm:a}) Let $p_i := \sigma_{k-s+i}(q_1,\ldots,q_k)$ for $i \in \{1,\ldots,s\}.$ Then $S \subseteq \{p_1 \ge 0,\ldots,p_s \ge 0\}.$ Moreover, by Lemma~\ref{el:sym:lemma}, for each $x \in \rmcmd{bd} S$ there is a neighborhood $B(x,\varepsilon)$ such that $S \cap B(x,\varepsilon) = \{p_1 \ge 0,\ldots,p_s \ge 0\} \cap B(x,\varepsilon).$ That means
\begin{equation} \label{S:cup:T:repr}
\{p_1 \ge 0,\ldots,p_s \ge 0\} = S \cup T,
\end{equation}
where $T \subseteq \mathbb{R}^d$ is semialgebraic, closed and
\begin{equation} \label{T:S:disj}
T \cap S = \emptyset.
\end{equation}
Now, according to Proposition~\ref{sep:disjoint}, we can choose $p_{s+1} \in \mathbb{R}[X]$ which separates $S$ from $T.$
\ref{av:thm:b}) We take $p_1,\ldots,p_{s}$ as before. Let $\{x_1,\ldots,x_m\}$ be the set of all points in $S$ where exactly $s$ polynomials $q_i$ vanish. We see that $R := ( \{p_1 \ge 0,\ldots, p_{s-1} \ge 0\} \setminus S) \cup \{x_1,\ldots,x_m\}$ is closed. Clearly, $R \cap S = \{x_1,\ldots,x_m\}.$ Moreover, for each $x_i$ there is a ball $B(x,\varepsilon)$ such that $p_s$ separates $B(x_i,\varepsilon) \cap S$ and $B(x_i,\varepsilon) \cap R$. Now, according to Proposition~\ref{sep:finite:intersect}, we can modify $p_s$ to a polynomial separates $S$ from $R.$
\end{proof}
\begin{remark}
Analogous semi-effective results can also be obtained for basic open sets by using similar methods.
\end{remark}
\section{Minimal description of polyhedra} \label{sect:polytopes}
A subset of $\mathbb{R}^d$ is said to be a \term{polyhedron} if it is the intersection of finitely many closed halfspaces. Bounded polyhedra are called \term{polytopes}. For background information on polytopes and polyhedra we refer to \cite{Ziegler-book-1995}. By $\mathcal{F}(P)$ we denote the set of all faces of $P.$ By $\mathcal{F}_k(P)$ we denote the set of all $k$-dimensional subfaces of $P.$ If the choice of $P$ is clear we merely write $\mathcal{F}_k$ and $\mathcal{F}.$ Polytopes of dimension $d$ are called \term{$d$-polytopes}. Faces of dimension $0$ and $\dim P-1$ are called \term{vertices} and \term{facets}, respectively. We introduce the \term{$k$-skeleton} (also called the set of all \term{$k$-extremal points}) of $P$ by $$\rmcmd{ext}_k P := \bigcup_{F \in \mathcal{F}_k(P)} F.$$ We also write $\rmcmd{ext} P:=\rmcmd{ext}_0 P$.
\begin{notationsremarks} \label{rem:polytopes} Let $S \subseteq \mathbb{R}^d$ be a \emph{polyhedron}.
\begin{enumerate}[a)]
\item For a $k$-face $F$ of $S$ we fix a degree-one polynomial $l_F$ such that $F = \{l_F = 0\} \cap P$ and $l_F \ge 0$ on $S.$
Then
$$D_k(S):= \bigcap_{F \in \mathcal{F}_k} \{l_F \ge 0\}$$
is called a \emph{$k$-support} of $S.$ The $k$-support depends on the choice of $l_F$'s. We have $D_{d-1}(S) = S.$
\item \label{08.07.29,15:59} For $l < k$ every $l$-face of $S$ is also an $l$-face of $D_k(S)$.
\item If $S$ is compact, $D_k(S)$ is not compact in general, unless $l_F$'s are chosen in a suitable way.
\item For a vertex $x \in S$ we set
\begin{align*}
S_x := \bigcap_{\overtwocond{F \in \mathcal{F}_{d-1}(S)}{x \in F}} \{l_F \ge 0\} & & \mbox{and} & & \mymark{S}_x := \bigcap _{\overtwocond{F \in \mathcal{F}_{d-1}(S)}{x \not\in F}} \{l_F > 0\}.
\end{align*}
Notice that $P_x$ is closed and $\mymark{P}_x$ is open.
\item \label{chain:of:k:supports} Now let $S$ be a polytope. Then there exists a sequence of $k$-supports $D_{-1}(S),\ldots, D_{d-1}(S)$ with
$$
S = D_{d-1}(S) \subseteq \cdots \subseteq D_0(S) \subseteq D_{-1}(S) = \mathbb{R}^d.
$$
and $D_k(S) \setminus \rmcmd{ext}_{k-1} S \subseteq \rmcmd{int} (D_{k-1})$ for $k=0,\ldots,d-1$.
\item For $k=0,\ldots,d-1$ one has
\begin{enumerate}[i)]
\item \label{08.07.29,16:25} $D_k(S) = \bigcap_{x \in \rmcmd{ext} S} D_k(S)_x.$
\item \label{08.07.29,16:26} There is a compact semialgebraic set $R_k$ such that $$S \subseteq \rmcmd{int}(R_k) \subseteq R_k \subseteq \bigcup_{x \in \rmcmd{ext} S} \mymark{D_k(S)}_x.$$
\end{enumerate}
Here i) holds, since each facet of $D_k(S)$ contains a vertex of $S,$ and ii) follows from the inclusion
\begin{equation} \label{08.09.17,17:25}
S \subseteq \bigcup_{x \in \rmcmd{ext} S} \mymark{D_k(S)}_x.
\end{equation}
Let us show \eqref{08.09.17,17:25}. Take an arbitrary $y \in S.$ Then there exists a unique face $G$ with $y \in \rmcmd{relint} G.$ Let $x$ be any vertex of $G$ and let $F$ be a facet with $x \not\in F.$ Then $y \not\in F,$ since otherwise we would have $\rmcmd{relint} G \cap F \ne \emptyset,$ which implies $G \subseteq F$ and by this $x \in F,$ a contradiction. Hence $l_F(y)>0.$ This yields \eqref{08.09.17,17:25}.
\end{enumerate}
\end{notationsremarks}
\begin{proposition} \label{main proposition}
For $k=0,\ldots,d-1$ there is a polynomial $p_k$ such that
\begin{enumerate}[a)]
\item $p_k \ge 0$ on $S$
\item \label{main prop cond 2} $p_k \le 0$ on $D_{k-1}(S) \setminus \rmcmd{int} (D_k(S))$
\item \label{main prop cond 3} $\{p_k = 0\} \cap (D_{k-1}(S) \setminus \rmcmd{int}(D_k(S))) \subseteq S$
\end{enumerate}
\end{proposition}
In \ref{main prop cond 2}) and \ref{main prop cond 3}) we could also replace $D_{k-1}(S) \setminus \rmcmd{int} (D_k(S))$ by $D_{k-1}(S) \setminus D_k(S)$, but for the proof by induction, which we give below, it is better to have it this way.
\begin{proof}[Proof of Proposition~\ref{main proposition}] The proof is by induction on $d$. For starting the induction argument we consider the cases $k=0$ and $k=d-1$ separately, which is done in Steps 1 and 2 of the proof. In the remaining steps we apply the inductive assumption to all vertex figures of $S$. Appropriately combining the polynomials associated to the vertex figures we generate polynomials associated to $S$. In Step~3 for each vertex of $S$ we construct a polynomial satisfying a)-c) in a small neighborhood of that vertex. We combine these polynomials in Step 6, thus getting a polynomial $r_k$ and show in Step 7 that $r_k$ fulfills the conditions a),b),c) locally, that means in a set $Q_{k-1}$ which we get by restriction to $R_k$. This is the main step. It uses decompositions of $Q_{k-1}$ which are explained in Step 4 and 5. Finally, in Step 8 we globalize $r_k$ in order to get $p_k$.
\emph{Step~1: $k=0$.} We need to show the existence of $p_0$ such that $p_0 \ge 0$ on $S$, $p_0 \le 0$ on $\mathbb{R}^d \setminus \rmcmd{int} D_0(S)$ and $\{p_0=0\} \cap (\mathbb{R}^d \setminus \rmcmd{int} D_0(S)) \subseteq S$. We use Proposition~\ref{sep:finite:intersect}, where $T:= \mathbb{R}^d \setminus \rmcmd{int}(D_0(S))$. Then $S \cap T = \rmcmd{ext} S $ is a finite set, and a local separation of $S$ from $T$ around each $x \in S \cap T$ can be achieved.
\emph{Step~2: $k=d-1$.} Let $\mathcal{F}_{d-1}(S)=\{F_1,\ldots,F_r\}$. We define $p_{d-1} := \prod_{i=1}^r l_{F_i}$. Clearly, $p_{d-1} \ge 0$ on $S$. Now let
$$S_i:= \{l_{F_i} \le 0\} \cap \bigcap_{\overtwocond{j=1,\ldots,r}{j \ne i}} \{l_{F_j} \ge 0\}.$$
Then by construction we have $D_{d-2}(S) \subseteq \bigcup_{i=1}^r S_i$, $p_{d-1} < 0$ on $(\bigcup_{i=1}^r S_i) \setminus S$ and $p_{d-1}=0$ on $\rmcmd{bd} S$. This proves b) and c).
The case $d=1$ is trivial. Cases 1 and 2 yield the assertion for $d=2$. Now assume that $d \ge 3$.
\emph{Step~3: Construction of a local solution $q_{i,k}$.} Let $\rmcmd{ext} S = \{x_0,\ldots,x_m\}$. For small $\varepsilon>0$ the hyperplane $\{l_{x_i} = \varepsilon\}$ intersects $S$, say $S^i := \{l_{x_i}= \varepsilon\} \cap S$, where we choose $\varepsilon$ such that $l_{x_i}(x_j) > \varepsilon$ for $j \ne i$. Also, we set $D_{i,k-1} = \{l_{x_i} = \varepsilon\} \cap D_k(S)$. So, we can use the inductive assumption to the $(d-1)$-dimensional polytope $S^i$ (in the $(d-1)$-dimensional affine space $\rmcmd{aff} S^i$). For all $\varepsilon>0$ as above $S^i$ remains the same up to a homothety. Let $1 \le k \le d-2$ be given. We want to construct a suitable polynomial $p_k$. For $i=1,\ldots,m$ let $p_{i,k-1}$ be a polynomial as in the assertion with respect to the polytope $S^i$ and the sets $D_{i,k-1}$ and $D_{i,k-2}$. Now let $q_{i,k}$ be the homogenization of $p_{i,k}$ with respect to the center $x_i$. Around $x_i$ the polynomial $q_{i,k}$ satisfies the properties a)-c). So, in order to generate $p_k$ we should combine the polynomials $q_{i,k}$ in a suitable way.
\emph{Step~4: Notations and Remarks.} For $k=0,\ldots,d-1$ let
\begin{equation} \label{compactified playground}
Q_{k-1} := R_k \cap (D_{k-1}(S) \setminus \rmcmd{int} (D_k(S)))
\end{equation}
Then $Q_{k-1}$ is compact. Let $G_1,\ldots,G_s$ be the facets of $D_k(S)$. (Here we keep $k$ fixed, so we omit the index $k$ at the $G_i$'s.) For $\delta>0$ and $x_i \in S_0 = \{x_1,\ldots,x_m\}$ let
\begin{equation*}
\mymark{D_k(S)}_{x_i,\delta} := \bigcap_{\overtwocond{j=1,\ldots,s}{x_i \not\in G_j}} \{l_{G_j} \ge \delta\}.
\end{equation*}
We have
\begin{equation*}
\mymark{D_k(S)}_{x_i,\delta} \subseteq \mymark{D_k(S)}_{x_i} \subseteq \rmcmd{cl} \mymark{D_k(S)}_{x_i} = \bigcap_{\overtwocond{j=1,\ldots,s}{x_i \not\in G_j}} \{l_{G_j} \ge 0\}.
\end{equation*}
By Remark~\ref{rem:polytopes}f)ii) there exists a $\delta>0$ such that
\begin{equation} \label{Q inclusion}
Q_{k-1} \subseteq \bigcup_{i=1}^m \mymark{D_k(S)}_{x_i,\delta}.
\end{equation}
\emph{Step~5: Notations and Remarks.} For every decomposition $\{1,\ldots,m\} = \{i\} \cup A \cup B$ let
\begin{align*}
Q_{k-1}^+(A) & := \bigcap_{\alpha \in A} Q_{k-1} \cap \mymark{D_k(S)}_{x_\alpha}, \\
Q_{k-1}^-(B) & := \bigcap_{\beta \in B} Q_{k-1} \setminus \mymark{D_k(S)}_{x_\beta}, \\
Q_{k-1}^0(i) & := Q_{k-1} \cap \mymark{D_k(S)}_{x_i,\delta}, \\
Q_{k-1}(i,A,B) & := Q_{k-1}^0(i) \cap Q_{k-1}^+(A) \cap Q_{k-1}^-(B).
\end{align*}
The sets $Q_{k-1}^0(i)$ and $Q_{k-1}^-(B)$ are compact, the other two are in general not. In view of \eqref{Q inclusion}, we have $Q_{k-1} = \bigcup_{(i,A,B)} Q_{k-1}(i,A,B)$, where the union runs over all partitions $\{1,\ldots,m\} = \{i\} \cup A \cup B$. Also, the sets $Q_{k-1}(i,A,B)$ form a partition of $Q_{k-1}(i)$ for a fixed $i$.
\emph{Step~6: Construction of a polynomial, which satisfies the assertion on $Q_{k-1}$.} Again $\rmcmd{ext} S = \{x_0,\ldots,x_m\}$, $G_1,\ldots,G_s$ are facets of $D_k(S)$ and $\delta$ is chosen as in Step~4. Moreover, by induction we constructed already polynomials $q_{i,k},$ $i=1,\ldots,m$. We choose $\rho>\delta$ such that $|l_{G_j}(x)| \le \rho$ for all $x \in R_k$, $j=1,\ldots,s$. Consider the polynomial $\kappa=\kappa_{\delta,\rho,m} \in \mathbb{R}[t]$ as in Lemma~\ref{lambda:mu:lem}.\ref{kappa:part}). We set
\begin{equation*}
r_{i,k} := q_{i,k} \prod_{\overtwocond{j=1,\ldots,t}{x_i \not\in G_j}} (\kappa \circ l_{G_j})^n
\end{equation*}
and
\begin{equation*}
r_k:= \sum_{i=1}^m r_{i,k},
\end{equation*}
where $n \in \mathbb{N}$ is to be fixed later. We claim the following.
\emph{Step~7: Verification that $r_k$ is a local solution.} For sufficiently large $n$ one has:
\begin{enumerate}[a')]
\item $r_k \ge 0$ on $S$,
\item $r_k \le 0$ on $Q_{k-1}$,
\item $\{r_k=0\} \cap Q_{k-1} \subseteq S$.
\end{enumerate}
Let us prove a'). We have $q_{i,k} \ge 0$ on $S$ for $i=1,\ldots,m$ and $\kappa \ge 0$ on $[-\rho,\rho]$, hence $r_{i,k} \ge 0$ on $S$ and finally $r_k\ge 0$ on $S$. In view of \eqref{Q inclusion} we may replace $Q_{k-1}$ by $Q_{k-1}(i,A,B)$ for a given decomposition $\{1,\ldots,m\} = \{i \} \cup A \cup B$. Clearly, we have $r_{i,k} \le 0$ on $Q_{k-1}^0(i)$ and $\{r_{i,k} = 0 \} \cap Q_{k-1}^0(i) \subseteq S$. Note that the factor $\prod_{\overtwocond{j=1,\ldots,t}{x_i \not\in G_j}} (\kappa \circ l_{G_j})^n$ is positive and arbitrarily large for sufficiently large $n$. Let
\begin{align*}
r_{k,A} := \sum_{\alpha \in A} r_{\alpha,k} & & \mbox{and} & & r_{k,B} := \sum_{\beta \in B} r_{\beta,k},
\end{align*}
so $r_k= r_{i,k} + r_{A,k} + r_{B,k}$. We have $r_{A,k} \le 0$ on $Q_{k-1}^0(i) \cap Q_{k-1}^+(A)$. So it is sufficient to show that $|r_{B,k}| \le \frac{1}{2} |r_{i,k}|$ on $Q_{k-1}^0(i) \cap Q_{k-1}^-(B)$. For this it is enough to show that $|r_{\beta,k}| \le \frac{1}{2m} |r_{i,k}|$ on $Q_{k-1}^0(i) \setminus \mymark{D_{k}(S)}_{x_{\beta}}$ for all $\beta \in B$. We write
$$
Q_{k-1}^0(i) \setminus \mymark{D_k(S)}_{x_\beta} = \bigcup_{\overtwocond{j=1,\ldots,t}{x_\beta \not\in G_j}} \left( Q_{k-1}^0(i) \cap \{l_{G_j} \le 0\}\right).
$$
So finally it remains to show that $|r_{\beta,k}| \le \frac{1}{2m} |r_{i,k}|$ on $Q_{k-1}^0(i) \cap \{l_{G_j} \le 0\}$ for given $\beta, j$ and all sufficiently large $n \in \mathbb{N}$. Clearly $\{r_{i,k}=0\} \subseteq \{r_{\beta,k}=0\}$ and $q_{\beta,k}$ is bounded on this set. Moreover, by the properties of $\kappa$, we have
$$\left| \prod_{\overtwocond{j=1,\ldots,t}{x_\beta \not\in G_j}} l_{G_j}\right| \le \frac{3}{4}.$$
Thus, the claim follows from the H\"ormander-{\L}ojasiewicz inequality (see Theorem~\ref{more than loj}).
\emph{Step 8: Conclusion.} Let $T_1 := Q_{k-1}$, $T_2 := D_{k-1}(S) \setminus \rmcmd{int} (R_k)$ and $h \in \mathbb{R}[X]$ be a non-negative polynomial which vanishes only on the Zariski closure of $S \cap T_1 = \rmcmd{ext}_k S$. Such a polynomial can easily be found with the additional property that $h > 0$ on $T_2$ and for $x \in T_2$: $h(x) \rightarrow \infty$ as $\|x\|\rightarrow \infty$. Finally, let $f=r_k$. For $S, T_1, T_2, h, f$ we apply Proposition~\ref{S:T1:T2}, which gives us a polynomial $p_k$ such that $p_k \ge r_k$ on $S$, $\rmcmd{sign} (p_k) \le \rmcmd{sign} (r_k)$ on $Q_{k-1}$ and $p_k<0$ on $D_{k-1}(S) \setminus \rmcmd{int} (R_k)$. That means $p_k$ satisfies a)-c).
\end{proof}
As an immediate consequence of Proposition~\ref{main proposition} we obtain
\begin{corollary} \label{polytopes:result}
Let $S \subseteq \mathbb{R}^d$ be a $d$-dimensional polytope. Then there are polynomials $p_0,\ldots,p_{d-1} \in \mathbb{R}[X]$ such that $S = \{p_0 \ge 0, \ldots, p_{d-1} \ge 0\}.$
\end{corollary}
In the proof of Theorem~\ref{every polyhedron is representable} we shall use the following observation, see \cite[Proposition~2.1]{Groetschel-Henk-2003} and also \cite[Section~6.5]{MR1659509} or use an argument with fans as in \cite[Chapter VI, Section~7]{MR1393194}.
\begin{proposition} \label{number of vanishing poly}
Let $S$ be a $d$-polyhedron in $\mathbb{R}^d$, let $S = \{q_1 \ge 0,\ldots,q_m \ge 0\}$ for $q_1,\ldots,q_m \in \mathbb{R}[X]$ and let $F$ be a face of $S.$ Then at least $d-\dim F$ polynomials $q_i$ vanish on $F$. In particular, $m \ge d- \dim F$.
\end{proposition}
\begin{proof}[Proof of Theorem~\ref{every polyhedron is representable}]
Most arguments presented below are also given in \cite{Averkov-Henk-2009b}, but for the sake of completeness we give the whole proof.
We write $S=S_0 \times \mathbb{R}^k$, where $\mathbb{R}^d = \mathbb{R}^{d-k} \times \mathbb{R}^k$ and $S_0 \subseteq \mathbb{R}^{d-k}$ in suitable coordinates. We may assume that $k< d$, so $S$ admits at least one $k$-face, hence, by Proposition~\ref{number of vanishing poly}, $S$ cannot be described by fewer than $d-k$ polynomials.
Conversely, every polynomial description of $S_0$ easily extends to $S$. Note also, that $S_0$ does not contain any line. So it remains to show that a polyhedron $S \subseteq \mathbb{R}^d$ which does not contain any line is representable by $S=\{p_0 \ge 0,\ldots,p_{d-1} \ge 0\}$ for suitable $p_0,\ldots,p_{d-1} \in \mathbb{R}[X].$
For this we consider $\mathbb{R}^d$ as affine subspace $\setcond{x \in \mathbb{R}^{d+1}}{x_{d+1}=1}$ in $\mathbb{R}^{d+1}$ and form the cone $C^0$ over $S$, that is $C^0 := \setcond{\lambda a}{a \in S, \ \lambda \ge 0}$. The set $C:= \rmcmd{cl} C^0$ is a polyhedral cone. Since $S$ does not contain a line, there is a linear form $l$ of $\mathbb{R}^{d+1}$ such that $C$ intersects the hyperplane $l=1$ properly, that is $S':=\{l=1\} \cap C$ is a polytope. By Corollary~\ref{polytopes:result} we can find polynomials $q_0,\ldots,q_{d-1}$ such that $S' = \{q_0 \ge 0,\ldots,q_{d-1} \ge 0, \ l=1\}$. Possibly after an appropriate modification with the help of $l$, we assume that $q_0,\ldots,q_{d-1}$ are homogeneous and of even degree. We would like that $\{q_0 \ge 0,\ldots,q_{d-1} \ge 0\} = C \cup (-C)$, but this is possibly not true, since $\{q_0 \ge 0,\ldots,q_{d-1} \ge 0, \ l=0\}$ may contain points other than $o$. In order to avoid the above situation we keep $q_1,\ldots,q_{d-1}$ as before and adjust $q_0$. We have $\{q_0=0\} \cap S = \rmcmd{ext} S$ and $\{q_0 \ge 0, \ l=1\}$ is compact. Then $q_0$ is negative semidefinite on $\{l=0\}$. We need to replace $q_0$ by a polynomial $p$ with $\{p \ge 0, \ l=0\} = \{o\}$. For this let $r_1,\ldots,r_m$ be (homogeneous) non-negative quadratic polynomials such that $\{r_1=0\},\ldots,\{r_m=0\}$ are the affine hulls of the extremal rays of $C$, and let $r:=r_1,\ldots,r_m$. Now let
$$
p:=q_0 l^{N-\deg q_0} - c r^N,
$$
where $c>0$ is sufficiently small and $N \in \mathbb{N}$ is sufficiently large. Then $\{p \ge 0, \ l=0\} = \{o\}$ and $S \subseteq \{p \ge 0, \ l=1\}$ by the H\"ormander-{\L}ojasiewicz inequality. Hence
$$
\{p\ge 0, \ q_1 \ge 0, \ldots,q_{d-1} \ge 0\} = C \cup (-C)
$$
Now we set $p_0(X_1,\ldots,X_d) := p(X_1,\ldots,X_d,1)$ and $p_i(X_1,\ldots,X_d) = q_i(X_1,\ldots,X_d,1)$ for $i=1,\ldots,d-1$.
\end{proof}
\begin{remark}
From the proofs we see that the polynomial representation $S = \{p_0 \ge 0,\ldots, p_{d-1} \ge 0\}$ of a $d$-dimensional polytope $S$ that we construct above satisfies the condition
\begin{equation} \label{p_i ext_i}
\{p_i = 0 \} \cap S \subseteq \rmcmd{ext}_i S
\end{equation}
for $i \in \{0,\ldots,d-1\}$. In fact, an arbitrarily representation of this size can easily be converted into a form satisfying the above properties. Let $S = \{q_0 \ge 0,\ldots,q_{d-1} \ge 0\}$ be an arbitrary polynomial representation of a $d$-dimensional polyhedron $S$. For $p_i:=\sigma_{i+1}(q_0,\ldots,q_{d-1})$ one has $S = \{p_0 \ge 0,\ldots, p_{d-1} \ge 0\}$ (see Lemma~\ref{el:sym:lemma}). Furthermore, in view of Proposition~\ref{number of vanishing poly}, $p_i$'s satisfy \eqref{p_i ext_i}.
\end{remark}
\small
\providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
\providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
\providecommand{\MRhref}[2]{%
\href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
}
\providecommand{\href}[2]{#2}
|
\section{Introduction}\label{intro}
Let $X$ be a regular tree of degree $q+1\geq3$ with vertex set $V(X)$. Let $T:X\rightarrow X$ be a nontrivial hyperbolic automorphism or \emph{translation} on $X$, which is defined as an element in $\text{Aut}(X)$ with no fixed vertices or edges (see for example \cite[Chapter 1]{figa}). Denote by $d(x,Tx)$ the combinatorial \emph{distance} from a vertex $x \in V(X)$ to its image vertex under $T$. Let $\Gamma$ be a group of translations in $\text{Aut}(X)$ such that $G=X/\Gamma$ is a \emph{finite, simple, non-bipartite} $(q+1)$-regular graph, and let $\mathcal{K}$ be the conjugacy class of a nontrivial element $T_0\in\Gamma$. For $t\in\mathbb{R}$, $x\in V(X)$, we define
\begin{equation}\label{N}
N_{\mathcal{K}} (x,t) = \# \big\{ T \in \mathcal{K} : d(x,Tx) \leq t \big\}
\end{equation}
In this article we study the limiting behaviour of this counting function for increasing $t$. To do this we define a function $G_{\mathcal{K}}: V(X) \times \mathbb{C} \rightarrow \mathbb{C}$ for fixed $\mathcal{K}$ as follows:
\begin{equation}\label{G1}
G_{\mathcal{K}} (x,s) = \sum_{T \in \mathcal{K}} q^{-d(x,Tx)s}
\end{equation}
\begin{lem}\label{meromorphic}
The function $G_{\mathcal{K}} (x,s)$ as defined above \eqref{G1} is absolutely convergent for $\text{Re}(s)>2$.
\end{lem}
\begin{pf}
Note $\sum_{T \in \mathcal{K}} q^{-d(x,Tx)s} < \sum_{y\in V(X)} q^{-d(x,y)s}$. We can rewrite the last sum in terms of spheres $S_n(x)=\{y\in V(X) : d(x,y)=n\}$ as $\sum_{n=0}^\infty |S_n(x)| q^{-ns}$. Observe that $|S_n(x)|=(q+1)q^{n-1}$ for $n\geq1$ and use this to find $$\sum_{n=0}^\infty |S_n(x)| q^{-ns} \leq \sum_{n=0}^\infty 2q^nq^{-ns} = 2\sum_{n=0}^\infty q^{n(1-s)}$$ which converges for $\text{Re}(s)>2$ by the geometric series formula.
\end{pf}
In fact even more is true. In due course we will see that $G_\mathcal{K}$ has a meromorphic extension to $\mathbb{C}$, which is holomorphic for $\text{Re}(s)>\frac{1}{2}$.
The \emph{axis} $\mathfrak{a}(T)$ of a non-trivial translation $T$ on $V(X)$ is the unique geodesic which is mapped to itself by $T$. The \emph{displacement length} $\mu(T)$ of $T$ is given by
$$\mu(T) = \min_{x \in V(X)} d(x,Tx) \geq 1 $$
Note that $\mu(T) \in \mathbb{N}$. It is easy to see that the minimum is attained exactly for those vertices that lie in $\mathfrak{a}(T)$, as the vertices in $\mathfrak{a}(T)$ are shifted along the axis by $\mu(T)$ vertices under the action of $T$. Let $\delta(x,T)$ be the distance from a point $x$ to the axis of a translation $T$, that is
$$\delta(x,T) = \min_{y \in \mathfrak{a}(T)} d(x,y)$$
We then observe
\begin{equation}\label{distance}
d(x,Tx) = \mu(T) + 2 \delta(x,T)
\end{equation}
An element $P \in \Gamma$ is called \emph{primitive} if $\nexists \ Q \in \Gamma$ and $n>1$ such that $P=Q^n$. For every $T\in\Gamma$, we can find a unique $P \in \Gamma$ so that we can write $T=P^k$ for some $k\geq1$, and we call this the \emph{standard representation} of $T$. Now write $k=\nu(T)$ and call it the \emph{multiplicity} of $T$. Clearly $\mathfrak{a}(T) = \mathfrak{a}(P)$, and $\mu(T) = \mu(P) \cdot \nu(T)$. It is easy to prove that primitivity, $\mu(T)$ and $\nu(T)$ are invariant under conjugation in $\Gamma$.
We can now state our main theorem.
\begin{thm}\label{t1}
Let $N_{\mathcal{K}}(x,n)$ be defined as in equation \eqref{N} for positive integers $n$. Then as $n\rightarrow\infty$, where $n-\mu(\mathcal{K})$ is even,
$$N_{\mathcal{K}}(x,n) \sim q^{\frac{n-\mu(\mathcal{K})}{2}} \frac{ \mu(\mathcal{K})}{\nu(\mathcal{K}) |G|}$$
with $\mu(\mathcal{K})$ and $\nu(\mathcal{K})$ as defined above. By $N_{\mathcal{K}}(x,n)\sim Cq^{n/2}$ we mean that $N_{\mathcal{K}}(x,n)q^{-n/2}\rightarrow C$ as $n$ goes to infinity.
\end{thm}
We have to rule out the case where $n-\mu(\mathcal{K})$ is odd in the theorem because $N_\mathcal{K}(x,n)$ is a step function, which changes values only at points where $n-\mu(\mathcal{K})$ is even due to equation \eqref{distance}.
We are dealing with a type of problem that counts the number of lattice points on a graph inside an increasing ball. This type of problem is often solved using a discrete version of Selberg's trace formula on a regular tree (see \cite{selberg} and for the discrete version \cite{ahumada}, \cite{brooks}, \cite{t&w} or \cite{v&n}). In this particular case however, we are not able to use the trace formula due to the following. The trace formula is obtained using a $\Gamma$-invariant function $G(x,y)$ (a so-called point-pair invariant) such that
\begin{equation}\label{point-pair}
G(\gamma x,y)=G(x,\gamma y)=G(x,y)\ \forall\ \gamma\in\Gamma
\end{equation}
(see \cite{selberg} or \cite{t&w}). A natural choice would be $G_{\mathcal{K}}(x,y,s) = \sum_{T\in\mathcal{K}} q^{-d(x,Ty)s}$, but one easily checks that this function doesn't satisfy equation \eqref{point-pair} for arbitrary $x,y$. Instead, in this article we follow the method of Huber \cite{huber} using the function defined in equation \eqref{G1} which satisfies $G_{\mathcal{K}}(x,s) = G_{\mathcal{K}}(\gamma x,s)$ for all $\gamma\in\Gamma$.
Some other lattice point problems on the hyperbolic plane are easier to prove on the regular tree. Take for example the full lattice point problem, where we count $N_\Gamma(x,y,n) = \# \{ \gamma \in \Gamma : d(x,\gamma y)\leq n\}$. A consequence of the spherical average result in \cite{douma} for spheres $S_n(x_0)$ of vertices at distance exactly $n$ from $x_0$, is the asymptotic $\# \{ \gamma \in \Gamma : \gamma y \in S_n(x_0) \} \rightarrow \frac{|S_n(x_0)|}{|V(X/\Gamma)|}$. Using this and the fact that a ball of radius $n$ is the disjoint union of the spheres $\{S_r(x_0),\ 0\leq r \leq n\}$, we obtain
$$N_\Gamma(x,y,n)\rightarrow \frac{|B_n(x_0)|}{|V(X/\Gamma)|}$$
where we note that $|B_n(x_0)|$ is approximately equal to $\frac{q+1}{q-1}q^n$. For a proof of the corresponding problem on the hyperbolic plane, see \cite{patterson} or \cite[p 261]{buser}. Here we have
$$N_\Gamma(x,y,t) \rightarrow \frac{\pi}{\text{area}(M)}\ e^t$$
where $M$ is the manifold obtained by taking the quotient $\mathbb{H}/\Gamma$.
Another lattice point problem is that of counting the number of primitive closed paths of length at most $n$ in a finite regular graph $X/\Gamma$, which is a discrete analogue of counting primitive closed geodesics of length $\leq n$ on a closed hyperbolic surface. The result in the discrete case can be found in \cite[p 71]{sunada}, where we find that this number $\pi(n)=\# \{\text{prime cycles of length}\leq n \}$ has the following asymptotic behaviour
\begin{itemize}
\item when $X$ is non-bipartite, $\pi(n)\sim\frac{q^n}{n}$ as $n\rightarrow \infty$
\item when $X$ is bipartite, $\pi(n)\sim2\frac{q^n}{n}$ where $n$ is even and goes to $\infty$
\end{itemize}
The continuous analogue is the Prime Number Theorem for compact hyperbolic surfaces (see for example \cite[Theorem 9.4.14]{buser}). See also \cite{p&s} for a counting result of closed geodesics in negatively curved manifolds under homological constraints. A corresponding result for regular graphs can be found in \cite[p 72]{sunada}.
\section{Proof of the Theorem}
The discrete Laplacian is an operator which acts on functions on the vertices of any $(q+1)$-regular graph $G$ as follows:
\begin{equation}\label{laplacian}
\Delta f(x) = \frac{1}{q+1} \sum_{d(x,y)=1} f(y)
\end{equation}
Assume that the graph $G=X/\Gamma$ has $|V(G)|=N+1$ vertices. Then there is an orthonormal basis $\{\varphi_i\}_{i=0}^N$ of eigenfunctions of the Laplacian on the graph $G$. We use the fact that $\Delta = \frac{1}{q+1} A$ where $A$ is the adjacency matrix of $G$, to see that this basis has $N+1$ elements, and that we can order the eigenvalues such that $1=\lambda_0 > \lambda_1 \geq \ldots \geq \lambda_N > -1$ (note the eigenvalue $-1$ is excluded as $G$ is not bipartite: see for example \cite[Lemma 1.8]{chung}).
Using the canonical projection map $\pi:X\rightarrow G$, we can lift $\{\varphi_i\}_{i=0}^N$ to a set $\{ \tilde\varphi_i = \varphi_i\circ\pi \} _{i=0}^N$ of functions on $X$. It is easy to show that each $\tilde\varphi_i$ is an eigenfunction of the Laplacian on $X$ with the same eigenvalue $\lambda_i$. By definition the functions $\tilde\varphi_i$ are $\Gamma$-invariant.
Recall we defined a function $G_\mathcal{K}(x,s)$ on $X$ which satisfies $G_\mathcal{K}(x,s) = G_\mathcal{K}(\gamma x,s)$ for all $\gamma\in\Gamma$, so $G_\mathcal{K}$ is $\Gamma$-invariant. We can therefore view it as a function on the quotient $V(X/\Gamma)\times\mathbb{C}=V(G)\times\mathbb{C}$, where we call it $g_\mathcal{K}(x,s)$. Write $g_\mathcal{K}(x,s) = \sum_{i=0}^N F_i(s) \varphi_i(x)$. The Fourier coefficients $F_i(s)$ are given by $F_i(s) = \sum_{x\in V(G)} g_\mathcal{K}(x,s) \varphi_i(x)$. Now `lift' $g_\mathcal{K}(x,s)$ back up to $X$ and get
\begin{equation}\label{G_K}
G_\mathcal{K}(x,s) = \sum_{i=0}^N F_i(s)\tilde\varphi_i(x)
\end{equation}
where
\begin{equation}\label{fourier}
F_i(s) = \sum_{x\in\mathfrak{F}} G_\mathcal{K}(x,s) \tilde\varphi_i(x)
\end{equation}
for a fundamental domain $\mathfrak{F}\subset V(X)$ of $\Gamma$ on $X$. Note that there is a canonical one-to-one correspondence between $\mathfrak{F}$ and $V(G)$.
Choose and fix a translation $T^* \in \mathcal{K}$ with the standard representation $T^* = P^{\nu(\mathcal{K})}$ for some primitive $P$. Let \mbox{$H = \langle P \rangle$} be the subgroup of $\Gamma$ generated by $P$. We can write $\Gamma$ as a disjoint union of right cosets of $H$, i.e.
$$ \Gamma = \bigcup_{n=1}^{\infty} H A_n$$
with a fixed set $\{ A_n \}_{n=1}^\infty \subset \Gamma$. Then the elements $A_n^{-1} T^* A_n = T_n$ are pairwise disjoint, and run through all of $\mathcal{K}$ as $n=1,2,\ldots,\infty$. Define
$$\mathfrak{F}^* = \bigcup_{n=1}^\infty A_n(\mathfrak{F})$$
where $A_n(\mathfrak{F})=\{ A_n x : x\in\mathfrak{F} \}$. One easily checks that $\mathfrak{F}^*$ is a fundamental domain of the cyclic group $H$.
\begin{lem}\label{G*}
The Fourier coefficients $F_i(s)$ are given by
$$ F_i (s) = q^{-s\mu(\mathcal{K})} \sum_{x \in \mathfrak{F}^*} q^{-2s\delta(x,P)} \tilde \varphi_i (x)$$
\end{lem}
\begin{pf}
Let $k=\nu(\mathcal{K})$, and use equation \eqref{fourier} and the observations above to get
\begin{align}
F_i(s) & = \sum_{x\in \mathfrak{F}} G_{\mathcal{K}} (x,s) \tilde \varphi_i (x) = \sum_{x \in \mathfrak{F}} \sum_{n=1}^\infty q^{-d(x,A_n^{-1} P^{k} A_n x) s} \tilde \varphi_i (x) \nonumber \\
& = \sum_{n=1}^\infty \sum_{x \in A_n(\mathfrak{F})} q^{-d(x,P^k x)s} \tilde \varphi_i (x) = \sum_{x\in\mathfrak{F}^*} q^{-d(x,P^{k}x)s} \tilde\varphi_i(x) \nonumber \\
& = q^{-sk\mu(P)} \sum_{x \in \mathfrak{F}^*} q^{-2s\delta(x,\mathfrak{a}(P))} \tilde \varphi_i (x) \label{e2}
\end{align}
where the last equality is due to equation \eqref{distance}. Note $k\mu(P) = \mu(\mathcal{K})$.
\end{pf}
Note that $\delta(x,P) = \delta(P^n x,P)$ for any integer $n$, as the axes of $P$ and $P^n$ coincide, and $\tilde\varphi_i(x) = \tilde\varphi_i(P^n x)$ as $P^n \in \Gamma$ and $\tilde\varphi_i(x)$ is $\Gamma$-invariant. This means the terms in the sum in equation \eqref{e2} is invariant under $H = \langle P \rangle$. Hence we can replace $\mathfrak{F}^*$ in the sum of \eqref{e2} by \emph{any} fundamental domain of $H$. Take a segment of $\mathfrak{a}(P)$ of length $\mu(P)$ and all branches emanating from the vertices in this segment, excluding the two branches that emenate from the vertices at the ends of the segment in the direction of the axis. The vertices in this set form a new fundamental domain for $H$, which we call $\mathfrak{F}_P$, and using the fact we can interchange fundamental domains of $H$ shown above, we now sum over the vertices in $\mathfrak{F}_P$ instead of $\mathfrak{F}^*$ in equation \eqref{e2}.
\begin{figure}[ht]
\centering
\psfrag{a(P)}{$\mathfrak{a}(P)$}
\psfrag{X}{$X$}
\psfrag{X/H}{$X/H$}
\includegraphics[height=4.5cm]{F_P3.pstex}
\caption{The bold vertices in $X$ belong to $\mathfrak{F}_P$}
\end{figure}
The structure of the quotient graph $X/H$ can easily be deduced from $\mathfrak{F}_P$ (see for example Figure 1, where $\mu(P)=3$). We now want to transfer the functions $\tilde\varphi_i$ from $\mathfrak{F}_P$ to functions on the vertices of $X/H$, and to do this we use the obvious one-to-one correspondence between the vertex sets $\mathfrak{F}_P$ and $V(X/H)$. Note that the edge relations are preserved, and call the new function $\overline{\varphi_i}:V(X/H)\rightarrow\mathbb{R}$ for $i=0,\ldots,N$. It is easy to show these are eigenfunctions of the Laplacian on $X/H$. Using these definitions and equation \eqref{e2} we now have
\begin{align}
F_i(s) & = q^{-s\mu(\mathcal{K})} \sum_{x\in\mathfrak{F}_P} q^{-2s\delta(x,P)} \tilde\varphi_i(x) \nonumber \\
& = q^{-s\mu(\mathcal{K})} \sum_{x\in V(X/H)} q^{-2s\delta'(x,P)} \overline\varphi_i(x) \label{e3}
\end{align}
where $\delta'(x,P)$ is the distance from the vertex $x$ to the central loop in $X/H$, which is exactly equal to $\delta(x,P)$ on $X$ (the central loop in Figure 1 is a triangle).
Define \emph{levels} in $X/H$ as follows:
$$L_n=L_n(X/H) = \{ x \in V(X/H) : \delta'(x,P)=n \} \quad \text{for} \ n\geq0$$
The \emph{radial average} of a function $f:V(X/H)\rightarrow\mathbb{R}$ with respect to these levels is $\frac{1}{|L_n|} \sum_{x \in L_n} f(x)$. A straightforward calculation shows that the radial average of $\overline{\varphi}_i (x)$ gives an eigenfunction $\Phi_i(x)$ of the Laplacian on $X/H$ with eigenvalue $\lambda_i$. Clearly $\Phi_i(x)=\Phi_i(y)$ whenever $\delta'(x,P)=\delta'(y,P)=n$, so we shall write $\Phi_i(n)$ from now on, where $n\in\mathbb{Z}^{\geq0}$. These $\Phi_i(n)$ are similar to the spherical functions with eigenvalue $\lambda_i$ on the tree as defined for example in \cite{t&w}. Use the facts that $V(X/H) = \bigcup_{n=0}^\infty L_n$ and $\sum_{x \in L_n} \overline \varphi_i (x) = |L_n| \Phi_i(n)$ in equation \eqref{e3} to obtain
\begin{equation}\label{e4}
F_i(s) = q^{-s\mu(\mathcal{K})} \sum_{n=0}^\infty |L_n| q^{-2sn} \Phi_i(n)
\end{equation}
\begin{lem}\label{lua}
There are constants $\alpha_i^\pm$, $u_i^\pm$ depending only on $\varphi_i$, $\lambda_i$ and $q$ such that
$$F_i(s) = q^{-s\mu(\mathcal{K})} \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})} \frac{q-1}{q} \Big( \frac{u_i^+ + u_i^-}{q-1} + \frac{u_i^+}{1-\alpha_i^+ q^{1-2s}} + \frac{u_i^-}{1-\alpha_i^- q^{1-2s}} \Big)$$
except in the case that $\lambda_i = \pm \frac{2\sqrt{q}}{q+1}$, where we obtain
\begin{align}
F_i(s) = \frac{1}{q} \ q^{-s\mu(\mathcal{K})} & \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})} \Phi_i(0) + q^{-s\mu(\mathcal{K})} \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})} \frac{q-1}{q} \frac{\Phi_i(0)}{1\mp q^{1/2-2s}} \label{fourier D=0} \\
& + q^{-s\mu(\mathcal{K})} \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})} \frac{q-1}{q} \Phi_i(0) \big(\frac{\sqrt{q}-1}{\sqrt{q}+1}\big)^{\pm1} \frac{\pm q^{1/2-2s}}{(1\mp q^{1/2-2s})^2} \nonumber
\end{align}
\end{lem}
\begin{pf}
Assume first that $\lambda_i\neq\pm\frac{2\sqrt{q}}{q+1}$. Note that $|L_0| = \mu(P)=\frac{\mu(\mathcal{K})}{\nu(\mathcal{K})}$, and $|L_n| = q^{n-1} (q-1) \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})}$ for $n\geq1$. We use $\Delta \Phi_i(n) = \lambda_i \Phi_i(n)$ to find the recursion relation
\begin{equation}\label{recursion}
\Phi_i(n+1) = \frac{q+1}{q}\lambda_i\Phi_i(n) - \frac{1}{q}\Phi_i(n-1)\quad \text{ for } n\geq 1
\end{equation}
and $(q+1)\lambda_i\Phi_i(0) = (q-1)\Phi_i(1)+2\Phi_i(0)$. This leads to $\Phi_i(n) = u_i^+ (\alpha_i^+)^n + u_i^- (\alpha_i^-)^n$ for constants $\alpha_i^\pm$ and $u_i^\pm$ defined by
\begin{equation}\label{alpha}
\alpha_i^\pm = \frac{q+1}{2q}\lambda_i \pm \frac{\sqrt{(q+1)^2\lambda_i^2-4q}}{2q}
\end{equation}
\begin{equation}\label{u}
u_i^\pm = \bigg(\frac{1}{2} \pm \frac{(q+1)^2\lambda_i-4q}{2(q-1)\sqrt{(q+1)^2\lambda_i^2-4q}} \bigg) \cdot \Phi_i(0)
\end{equation}
Note that $\alpha_i^+\neq\alpha_i^-$ as the square root is nonzero due to the exclusion of $\lambda_i=\pm\frac{2\sqrt{q}}{q+1}$. Using the geometric series formula twice in equation \eqref{e4}, taking care with the $n=0$ term, we obtain
\begin{align}
F_i(s) = \frac{1}{q} \ q^{-s\mu(\mathcal{K})} \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})} \Phi_i(0) & + q^{-s\mu(\mathcal{K})} \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})} \frac{q-1}{q} \frac{u_i^+}{1-\alpha_i^+ q^{1-2s}} \nonumber \\
& + q^{-s\mu(\mathcal{K})} \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})} \frac{q-1}{q} \frac{u_i^-}{1-\alpha_i^- q^{1-2s}} \label{fourier final}
\end{align}
Observe that for the infinite sums to converge, we need $|\alpha_i^\pm q^{1-2s}|<1$. For $i\neq0$ it is easy to check that $|\lambda_i|<1$ and $|\alpha_i^\pm| <1$ for $i\neq0$, and that there is a real number $\sigma_0<\frac{1}{2}$ so that the sums obtained from equation \eqref{e4} converge for $\text{Re}(s)>\sigma_0$. However for the eigenvalue $\mu_0=1$ we have $\alpha_0^+=1$ and the infinite sum for $F_0(s)$ will only converge for $\text{Re}(s)>\frac{1}{2}$.
The square root in equation \eqref{alpha} is zero iff we have $\lambda_i=\pm\frac{2\sqrt{q}}{q+1}$, which implies $\alpha_i^+=\alpha_i^-=\alpha_i=\pm\frac{1}{\sqrt{q}}$. In this case $\Phi_i(n)= ( 1 + n\big(\frac{\sqrt{q}-1}{\sqrt{q}+1}\big)^{\pm1} ) \Phi_i(0)\alpha_i^n$. Using the geometric series formula and the series $\sum_{i=1}^\infty ix^i = \frac{x}{(1-x)^2}$, we obtain the required expression from \eqref{e4}. For the convergence of the two infinite sums obtained here we require $|q^{1/2-2s}|<1$ which implies $\text{Re}(s)>\frac{1}{4}$, which is consistent with the general case above.
\end{pf}
Calculating $\alpha_0^-$ and $u_0^\pm$ using $\widetilde \varphi_0(x) = \frac{1}{\sqrt{|G|}} \ \forall \ x\in V(G)$ we obtain
\begin{equation}\label{G final}
G_{\mathcal{K}} (x,s) = \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})|G|q^{s\mu(\mathcal{K})+1}} \big( 1+ \frac{q-1}{1-q^{1-2s}} \big) + \sum_{i=1}^{N} F_i(s) \widetilde\varphi_i(x)
\end{equation}
This is a meromorphic extension of $G_{\mathcal{K}} (x,s)$ defined in \eqref{G1} to the complex plane, which is holomorphic for $\text{Re}(s)>\frac{1}{2}$.
To finish the proof of our theorem, we use a refined version of the Tauberian theorem by Wiener-Ikehara from \cite{g&v} (see also \cite[chapter III Theorem 5.4]{korevaar}), which is a refinement of the Tauberian theorem in \cite{wiener}. This theorem requires a function $f(s)$ which converges for $\text{Re}(s)>1$ and has a simple pole at $s=1$. We choose $f(s) = G_{\mathcal{K}} (x,\frac{s}{2})$. The residue of $f(s)$ at $s=1$ is
\begin{equation}\label{A}
\text{Res}(f(s), 1) = \lim_{s\rightarrow1} (s-1)G_{\mathcal{K}} (x,\frac{s}{2}) = \frac{\mu(\mathcal{K})(q-1)}{\nu(\mathcal{K})|G|q^{(\mu(\mathcal{K})/2)+1}} \ \frac{1}{\ln q} :=A
\end{equation}
That means $f(s) = \frac{A}{s-1} + g(s)$ for some function $g(s)$ which is analytic for $\text{Re}(s)>1$. Check that $g(s)$ is analytic for $\text{Re}(s)\searrow1$ when $|\text{Im}(s)|<\frac{2\pi}{\ln q}$. Indeed, $g(s)$ has poles wherever $f(s)$ does, except we have removed the pole at $s=1$. As $f(s)$ has poles on the line $l=\{s:\text{Im}(s)=1\}$ at $s=1+ki\frac{2\pi m}{\ln q}$ for any $m\in\mathbb{Z}$, $g(s)$ has no poles on $l$ for $|\text{Im}(s)|<\frac{2\pi}{\ln q}$. Note also $A>0$.
Recall $N_{\mathcal{K}}(x,t) = \# \{ T \in \mathcal{K} : d(x,Tx) \leq t \}$. For fixed $x\in V(X)$, let
$$S(t) = N_{\mathcal{K}} (x,2t) = \# \{ T \in \mathcal{K} : d(x,Tx) \leq 2t \}$$
This is a non-decreasing step-function, which vanishes for $t<0$.
We now have all the ingredients we need to apply the theorem, which in our notation reads as follows.
\begin{thm}
Let $S(t)$ vanish for $t<0$, be non-decreasing, continuous from the right and such that
$$f(s) = \int_{0}^\infty q^{-s t} d S(t), \quad s = \sigma_1 + i \sigma_2$$
exists for $\text{Re}(s)=\sigma_1 >1$. Suppose that for some number $\rho>0$, there is a constant $A$ (necessarily $\geq0$) such that the analytic function
$$g(s) = f(s) - \frac{A}{s-1}, \quad s = \sigma_1 + i \sigma_2, \quad \sigma_1 >1$$
converges to a boundary function $g(1+i \sigma_2)$ in $L^1 (-\rho < \sigma_2 < \rho)$ as $\sigma_1 \searrow 1$. Let $\tau$ be the supremum of all possible numbers $\rho$. Then
$$ \frac{2\pi/\tau}{e^{2\pi/\tau}-1} A \leq \liminf_{t\rightarrow\infty} q^{-t} S(t) \leq \limsup_{t\rightarrow\infty} q^{-t} S(t) \leq \frac{2\pi/\tau}{1-e^{-2\pi/\tau}} A $$
\end{thm}
In our case $\tau = 2\pi/\ln q$. Using this and equation \eqref{A} we obtain
\begin{equation}\label{inf sup}
\frac{1}{q} \ \frac{\mu(\mathcal{K})}{\nu(\mathcal{K}) |G| q^{\mu(\mathcal{K})/2}} \leq \liminf_{t\rightarrow\infty} \ q^{-t}S(t) \leq \limsup_{t\rightarrow\infty} \ q^{-t}S(t) \leq \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})|G|q^{\mu(\mathcal{K})/2}}
\end{equation}
These estimates no longer depend on the choice of $x$.
Notice that as a consequence of \eqref{distance} when $\mu(\mathcal{K})$ is even, $S(t)$ will jump only at integer values of $t$ (the case where $\mu(\mathcal{K})$ is odd works similarly, except jumps occur only when $t+\frac{1}{2}$ is an integer). In this case, writing $m=[t]$ or equivalently
\begin{equation}\label{m}
t=m+\varepsilon\text{ with }m\in\mathbb{Z}\text{ and }\varepsilon\in [0,1)
\end{equation}
we have
\begin{equation}\label{t=m}
S(t)=S(m) \ \forall \ \varepsilon \in [0,1)
\end{equation}
Letting $a=\frac{\mu(\mathcal{K})}{\nu(\mathcal{K})|G|q^{\mu(\mathcal{K}) / 2}}$ we obtain the following estimates for the $\liminf$ and $\limsup$ respectively:
\begin{equation}\label{inf}
\frac{a}{q} \leq \liminf_{m\rightarrow\infty,m\in\mathbb{N}} q^{-(m+\varepsilon)} S(m) = q^{-\varepsilon} \liminf_{m\rightarrow\infty,m\in\mathbb{N}} q^{-m} S(m) \leq a
\end{equation}
\begin{equation}\label{sup}
\frac{a}{q} \leq \limsup_{m\rightarrow\infty,m\in\mathbb{N}} q^{-(m+\varepsilon)} S(m) = q^{-\varepsilon} \limsup_{m\rightarrow\infty,m\in\mathbb{N}} q^{-m} S(m) \leq a
\end{equation}
As this must hold for all $\varepsilon \in [0,1)$, we obtain
\begin{equation*}
\liminf_{m\rightarrow\infty,m\in\mathbb{N}} q^{-m} S(m) = \limsup_{m\rightarrow\infty,m\in\mathbb{N}} q^{-m} S(m) = \lim_{m\rightarrow\infty,m\in\mathbb{N}} q^{-m} S(m) = a
\end{equation*}
This means that for large integers $n=2m$ we have an approximation of $S(m)$ and hence $N_{\mathcal{K}}(x,n)$ as follows
$$S(m) \sim q^{m-\frac{\mu(\mathcal{K})}{2}} \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})|G|} \qquad \qquad N_{\mathcal{K}}(x,n) = S(\frac{n}{2}) \sim q^{\frac{n-\mu(\mathcal{K})}{2}} \frac{\mu(\mathcal{K})}{\nu(\mathcal{K}) |G|}$$
where from the definition of $m$ we require $n-\mu(\mathcal{K})$ to be even (because when $\mu(\mathcal{K})$ is odd, we take $m\in\frac{1}{2}+\mathbb{Z}$ in equation \eqref{m}). Using equation \eqref{t=m} it is clear that for any real $t$ such that $\mu(\mathcal{K})+2r\leq t < \mu(\mathcal{K})+2r+2$ for a non-negative integer $r$, we have $N_\mathcal{K}(x,t) = N_\mathcal{K}(x,\mu(\mathcal{K})+2r)$.
\textsc{Remark:}
We now discuss why we required that $G$ was non-bipartite. Most of the proof above can be used to show a weaker result, but Theorem \ref{t1} will not hold for bipartite $G$. The method of proof works for the bipartite case up to Lemma \ref{lua}, where we have to consider the effects on $G_\mathcal{K}(x,s)$ of $\lambda_N=-1$, the eigenvalue of the Laplacian which occurs exactly when $G$ is bipartite. The series for $F_N(s)$ requires $\text{Re}(s)>\frac{1}{2}$ to converge, and we obtain
\begin{align*}
G_{\mathcal{K}} (x,s) = & \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})|G|q^{s\mu(\mathcal{K})+1}} \big( 1+ \frac{q-1}{1-q^{1-2s}} \big) + \sum_{i=1}^{N-1} F_i(s) \widetilde\varphi_i(x) \\
& + \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})|G|q^{s\mu(\mathcal{K})+1}} \big(1+\frac{q-1}{1+q^{1-2s}}\big)
\end{align*}
As before the residue of $f(s)=G_\mathcal{K}(x,\frac{s}{2})$ at $s=1$ equals $A$, but now the function $g(s)=f(s)-\frac{A}{s-1}$ only converges to a boundary function for $|\text{Im}(s)|<\frac{\pi}{\ln q}=\tau$. The Tauberian theorem can still be applied, but only shows
\begin{align*}
\frac{1}{q^2} \frac{2q}{q+1} \ \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})|G|q^{\mu(\mathcal{K})/2}} \leq & \liminf_{t\rightarrow\infty} \ q^{-t}S(t) \\
\leq & \limsup_{t\rightarrow\infty} \ q^{-t}S(t) \leq \frac{2q}{q+1} \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})|G|q^{\mu(\mathcal{K})/2}}
\end{align*}
The difference between this estimate and that in equation \eqref{inf sup} is that the left and right estimates differ by a factor $q^2$ instead of the factor $q$ obtained in the non-bipartite case, and it is no longer possible to deduce a precise limit from the estimates (note that $S(t)$ can jump at any integer value $t$). All we can say here is
\begin{align*}
\frac{1}{q^2} \frac{2q}{q+1} \ \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})|G|q^{\mu(\mathcal{K})/2}} \leq & \liminf_{t\rightarrow\infty} \ q^{-t/2}N_\mathcal{K}(x,t) \\
\leq & \limsup_{t\rightarrow\infty} \ q^{-t/2}N_\mathcal{K}(x,t) \leq \frac{2q}{q+1} \frac{\mu(\mathcal{K})}{\nu(\mathcal{K})|G|q^{\mu(\mathcal{K})/2}}
\end{align*}
Acknowledgements: The author wishes to thank N.~Peyerimhoff for many helpful discussions. This work forms part of the author's PhD research, which is supported by the EPSRC.
|
\section{Introduction}
\label{intro}
Fractional Brownian motion (fBm) has become very important tool in modern probability and statistical modeling. Fractional Brownian motion is defined as a Gaussian process with certain covariance structure. Besides of this definition, fBm could be represented equivalently as an integral of a deterministic kernel with respect to an ordinary Brownian motion. In fact, there exist at least two such kernels: Mandelbrot-Van Ness kernel that has infinite support and compactly supported Molchan-Golosov kernel.
Fractional Brownian motion allows to model dependency because of its covariance structure. Hence it is a popular model in many applications. There is one parameter, namely Hurst parameter $H$ that describes the whole dependence structure. For Hurst parameter $H>\frac{1}{2}$ the process has long-range dependence property. For $H<\frac{1}{2}$ the increments are negatively correlated and for $H=\frac{1}{2}$ the increments are independent i.e. we come to the ordinary Brownian motion case. Of course, fBm has also several other properties such as self-similarity and stationarity of increments. Despite of all these properties, fBm is neither semimartingale nor Markov process (excluding the Brownian motion case $H=\frac{1}{2}$).
If one is interested in fBm because of its correlation structure, one might not need exactly fBm but just some process with the same covariance structure. The law of a Gaussian process is determined uniquely by its second order structure. However, if we drop the assumption of Gaussianity, then the covariance structure does not determine the law uniquely. Thus, there are several possible ways to generalize fractional Brownian motion to the case of fractional L\'evy processes. By choosing different ways of generalization, we preserve different properties of fBm.
In this paper we define fractional L\'evy processes by two different integral transformations. This means basically that we take the integral representation of fractional Brownian motion with respect to an ordinary Brownian motion and replace the driving Brownian motion with a general square integrable L\'evy process. The processes that we will end up with, share the covariance structure of fractional Brownian motion. However, these processes could be more flexible in modeling than fractional Brownian motion, since the driving L\'evy noise is more general than the Gaussian one. For example, we might be able to capture such a phenomenon as a shock in the stock market (jump of the driving L\'evy process) that affects the market with delay and has some long term impacts. The applications in different fields of science might also be possible.
Fractional L\'evy processes by Mandelbrot-Van Ness representation were first defined by~\cite{benassi}. The theory was developed further by~\cite{marquardt}. Molchan-Golosov transformation has been used in fractional L\'evy process setting for defining fractional subordinators in~\cite{bender}. The general definition for fractional L\'evy processes by Molchan-Golosov transformation is new to the best of our knowledge.
There are also several other related concepts in the literature. One of the best known are fractional stable motions, see~\cite{samorodnitsky}. However, fractional stable motions are not fractional L\'evy processes in the sense that they would share the covariance structure of fBm.
\section{Preliminaries}
Let $(\Omega,\mathcal{F},(\mathcal{F}_t)_{t\in \R},\Pro)$ denote a fixed filtered probability space.
\subsection{Definition and some properties of fBm}
\begin{defin}[Fractional Brownian motion]
Let $H\in(0,1)$. Fractional Brownian motion with Hurst parameter $H$ is a zero mean Gaussian process $B^H$ with covariance
\begin{equation*}
\E B^H_t B^H_s=\frac{1}{2}\left ( |t|^{2H}+|s|^{2H}-|t-s|^{2H}\right ).
\end{equation*}
\end{defin}
Besides of this definition, we can represent fractional Brownian motion in many equivalent ways. For example, we can choose integral representations with respect to an ordinary Brownian motion.
On one hand, let $(W_t)_{t\in \R}$ be two-sided Brownian motion, i.e. $W_t=W^{(1)}_t$, when $t\geq 0$ and $W_t=-W^{(2)}_{-t}$, when $t<0$. Here $W^{(1)},W^{(2)}$ are independent Brownian motions. Then it holds that
\begin{equation*}
\left(B^H_t\right )_{t\in \R
{=}\left (\int_{-\infty}^t f_H(t,s)dW_s\right)_{t\in \R},
\end{equation*}
where the Mandelbrot-Van Ness kernel $f_H$ is given by
\begin{equation*}
f_H(t,s)=C_H\left ( (t-s)_+^{H-\frac{1}{2}}-(-s)_+^{H-\frac{1}{2}}\right), \quad s,t\in \R,
\end{equation*}
and the constant is given by
\begin{equation*}
C_H=\left ( \int_0^\infty \left ( (1+s)^{H-\frac{1}{2}}-s^{H-\frac{1}{2}}\right )^2ds+\frac{1}{2H}\right )^{-\frac{1}{2}}=\frac{\big(2H \sin \pi H
\Gamma(2H)\big)^{1/2}}{\Gamma(H+1/2)}.
\end{equation*}
On the other hand, fBm can be represented as well on compact interval by
\begin{equation*}
\left (B^H_t\right)_{t\geq 0}=\left (\int_0^tz_H(t,s)dW_s\right)_{t\geq 0},
\end{equation*}
where the Molchan-Golosov kernel $z_H$ is given by
\begin{equation*}
z_H(t,s)=c_H (t-s)^{H-\frac{1}{2}}F\left ( \frac{1}{2}-H,H-\frac{1}{2},H+\frac{1}{2},\frac{s-t}{s}\right), \quad 0< s< t<\infty,
\end{equation*}
and $z_H(t,s)=0$ otherwise. Here the Gauss' hypergeometric function $F$ of $x\in \R$ with parameters $a,b,c$ is defined by
\begin{equation*}
F(a,b,c,x)=\sum_{j=0}^\infty \frac{(a)_j(b)_j}{(c)_j}\frac{x^j}{j!},
\end{equation*}
where $(a)_0=0$ and $(a)_k=a\cdot(a+1)\dots (a+k-1)$ for $k\in \N$. The constant $c_H$ is given by
\begin{equation*}
c_H=\frac{1}{\Gamma(H+\frac{1}{2})}\left ( \frac{2H\Gamma(H+\frac{1}{2})\Gamma(\frac{3}{2}-H)}{\Gamma(2-2H)} \right)^{\frac{1}{2}}.
\end{equation*}
For $H>\frac{1}{2}$ we have the following simplified form of the kernel
\begin{equation*}
z_H(t,s)=\Big(H-\frac{1}{2}\Big)c_Hs^{\frac{1}{2}-H}\int_s^t u^{H-\frac{1}{2}}(u-s)^{H-\frac{3}{2}}du, \quad 0<s<t<\infty.
\end{equation*}
Note that we do not need the definition of any two-sided process for the compact interval Molchan-Golosov representation of fBm. For more details on the integral representations of fBm, see for example~\cite{jost} or~\cite{nualart}.
\subsection{L\'evy processes}
Consider now the conventions related to L\'evy processes. By the well-known L\'evy-Khinchine theorem, the characteristic function of a L\'evy process $L$ at time $t\geq 0$ can be represented as
\begin{equation*}
\E e^{iuL_t}=\exp \left (t \Psi(u)\right ),
\end{equation*}
where the characteristic exponent $\Psi$ is given by
\begin{equation*}
\Psi(u)=i\gamma u-\frac{1}{2}\sigma^2u^2+\int_\R \left ( e^{iux}-1-iux1_{\{|x|<1\}}\right )\nu(dx),
\end{equation*}
with $\gamma \in \R$, $\sigma^2\geq 0$ and $\nu$ being a measure concentrated on $\R \backslash \{0\}$ and satisfying
\begin{equation*}
\label{levymeasurecondition}
\int_\R (1\wedge x^2)\nu(dx)<\infty;
\end{equation*}
see for instance~\cite{kyprianou}~or~\cite{sato}. We call $(\gamma,\sigma^2,\nu)$ the characteristic triplet of $L$. From now on we assume that $\E L_1^2<\infty$ and $\E L_1=0$. For simplicity we assume that there is no Gaussian component, i.e. $\sigma^2=0$. With these assumptions we see that the characteristic function can be written as
\begin{equation*}
\E e^{iuL_t}=\exp \left (t \int_{\R}\left (e^{iux}-1-iux\right )\nu(dx)\right ),
\end{equation*}
as in~\cite{bender2}.
\subsection{FLp by infinite interval transformation}
Fractional L\'evy processes by infinite interval transformation were defined for $H \in (\frac{1}{2},1)$ in~\cite{marquardt}. However, the $L^2$- definition in~\cite{marquardt} can be extended for $H\in (0,\frac{1}{2})$ as well.
\begin{defin}[Two-sided L\'evy processes]
A two-sided L\'evy process or a L\'evy process on $\R$, $(L_t)_{t\in \R}$, is defined as $L_t=L^{(1)}_t$ if $t\geq 0$ and $L_t=-L^{(2)}_{-(t-)}$ if $t<0$, where $L^{(1)},L^{(2)}$ are independent and identically distributed L\'evy processes. We say that the characteristic triplet and exponent of $L^{(1)}$ are the characteristic triplet and exponent of $L$, respectively.
\end{defin}
\begin{defin}[FLp by Mandelbrot-Van Ness transformation]
\label{flpmvn}
Let $(L_t)_{t \in \R}$ be a L\'evy process defined on $\R$. Furthermore, assume that $\E L_1=0$ and $\E L_1^2<\infty$ and $L$ does not have Gaussian component. For $H\in\left(0,1\right)$ we say that
\begin{equation*}
X_t=\int_{-\infty}^t f_H(t,s)dL_s
\end{equation*}
is the fractional L\'evy process by Mandelbrot-Van Ness transformation (fLpMvN), where the stochastic integral is understood as a limit in probability of elementary integrals or in $L^2$-sense.
\end{defin}
Also the following facts follow from~\cite{marquardt}:
\begin{itemize}
\item The integral can be understood pathwise if $H>\frac{1}{2}$ as an improper Riemann integral.
\item The paths of fLpMvN are continuous when $H>\frac{1}{2}$ and even H\"older continuous of order $H-\frac{1}{2}$ on compacts.
\item Fractional L\'evy processes are never self-similar. This is proved for the case $H>\frac{1}{2}$, but the same proof works for the whole range $H\in(0,1)$.
\end{itemize}
In this paper, we contribute to the theory of fLpMvN by proving Theorem~\ref{flpmvnqv} on quadratic variation of fLpMvN.
\section{FLp as a result of compact interval transformation}
The main contribution of this paper is the theory of fractional L\'evy processes obtained via compactly supported Molchan-Golosov transformation. Convenient feature of these processes is that we do not need their infinite history.
\begin{defin}[FLp by Molchan-Golosov transformation]
\label{flpmgdef}
Let $(L_t)_{t\geq 0}$ be a L\'evy process without Gaussian component such that $\E L_1=0$ and $\E L_1^2<\infty$. Let $H \in(0,1)$. We call the stochastic process
\begin{equation*}
Y_t=\int_0^t z_H(t,s)dL_s
\end{equation*}
fractional L\'evy process by Molchan-Golosov transformation (fLpMG to be short). Here $z_H$ is the Molchan-Golosov kernel.
\end{defin}
The definition is understood as taking the limit in probability of elementary integrals in the sense of~\cite{rajput}~or~\cite{marquardt}.
\begin{rem}
It is also possible to include Gaussian component by considering a sum of fLpMG driven by pure jump L\'evy process and an independent fBm with the same Hurst parameter.
\end{rem}
The following theorem is the main result of the paper. It basically states that Brownian motion is the only process with slight moment assumptions such that the both integral transformations give the same process (in distribution).
\begin{theor}\label{mainthm}
Let $H\in (\frac{1}{2},1)$ and $L$ be a (two-sided) L\'evy process with non-degenerate L\'evy measure s.t. $\E L_1=0$.
1) If $\E |L_1|^3<\infty$ and $\E L_1^3\neq0$, then fLpMvN $X$ and fLpMG $Y$ driven by $L$ have different finite dimensional distributions.
2) If $\E L_1^4<\infty$, then fLpMvN $X$ and fLpMG $Y$ driven by $L$ have different finite dimensional distributions.
\end{theor}
The proof of this theorem is presented in section~\ref{mainproof}. Here it does not matter if the driving L\'evy process has Gaussian component or not because the proof is based on the 4th cumulant. In fact one does not need these moment assumptions, if the driving L\'evy process happens to be a compound Poisson process.
\begin{prop}
\label{mainimproved}
Let $L$ be a non-degenerate compound Poisson process s.t. $\E L_1=0$ and $\E L_1^2<\infty$. Then fLpMvN $X$ and fLpMG $Y$ have different finite dimensional distributions.
\end{prop}
Proof of the proposition is presented in section~\ref{mainproof}. A picture of the paths of the two fractional L\'evy processes is in Figure~\ref{flppathpic}. The driving L\'evy process is a compound Poisson process with jump sizes $\pm 1$. Note the different behavior near origin.
\begin{figure}
\begin{center}
\includegraphics[width=6cm]{mgflppath.eps}
\includegraphics[width=6cm]{mvnflppath.eps}
\caption{A path of fLpMG (left) and fLpMvN (right) with L\'evy measure $\lambda(\delta_{-1}+\delta_{1})$ and $H=0.75$. The different behavior of the two processes can be seen near origin, even though the driving paths are not the same.}
\label{flppathpic}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=6cm]{mgflppath25.eps}
\includegraphics[width=6cm]{mvnflppath025.eps}
\caption{A path of fLpMG (left) and fLpMvN (right) with $H=0.25$.}
\label{flppathpic2}
\end{center}
\end{figure}
\begin{rem}
An fLpMG driven by $L$ is adapted to the natural filtration $\mathcal{F}^L_t= \sigma \{L_u|0\leq u\leq t\}$. This is not the case for fLpMvN.
\end{rem}
\begin{prop}
An fLpMG $Y$ can be considered as the $L^2$- limit of approximating step functions. Moreover, the following $L^2$- isometry holds
\begin{equation*}
\E Y_t^2 = ||z_H(t,\cdot)||^2_{L^2([0,t])}\E L_1^2.
\end{equation*}
\end{prop}
\begin{proof}
This is a direct consequence of Proposition~2.1.~of~\cite{marquardt}.
\end{proof}
\begin{prop}[Autocovariance function]
\begin{equation*}
\E Y_tY_s=\frac{\E L_1^2}{2}\left ( t^{2H}+s^{2H}-|t-s|^{2H}\right ),
\end{equation*}
where $s,t\geq 0$.
\end{prop}
\begin{proof}
By $L^2$- isometry we have that
\begin{equation*}
\E Y_t^2 = ||z_H(t,\cdot)||^2_{L^2([0,t])}\E L_1^2.
\end{equation*}
We use the same argument for the increment (for the $L^2$- isometry, see \cite{marquardt} proposition~2.1.). Thus
\begin{equation*}
\E (Y_t-Y_s)^2=\E L_1^2\int_0^{t \vee s} (z_H(t,u)-z_H(s,u))^2du=\E L_1^2 \cdot |t-s|^{2H}.
\end{equation*}
Now
\begin{equation*}
\E Y_t Y_s=\frac{1}{2}\left ( \E Y_t^2 + \E Y_s^2 -\E (Y_t-Y_s)^2\right )=\frac{\E L_1^2}{2}\left (t^{2H}+s^{2H}-|t-s|^{2H}\right ).
\end{equation*}
\end{proof}
Besides the $L^2$- interpretation, we have also a partial result on the pathwise construction of fLpMG.
\begin{prop}[Pathwise construction]
\label{pathwise}
Let $H \in \left ( \frac{1}{2},1\right )$ and $L$ be a compound Poisson process with characteristic triplet $(0,0,\nu)$ such that $\E L_1=0$ and $\E L_1^2<\infty$. Then
\begin{equation*}
Y_t=\int_0^tz_H(t,s)dL_s=-\int_0^t \left ( \frac{d}{ds}z_H(t,s)\right )L_sds \quad \text{(almost surely)}.
\end{equation*}
\end{prop}
\begin{proof}
Fix $\omega \in \Omega$. A. s. there exists $\epsilon>0$ such that $L_s=0$ for all $s\in[0,\epsilon]$. The kernel $z_H(t,\cdot)$ is continuous when $s\neq 0$ by~\cite{jost}. Also its derivative is continuous on $(0,t)$. We get now by Lemma~2.1. of~\cite{eberlein} that
\begin{align*}
&\int_0^tz_H(t,s)dL_s=\int_{\epsilon}^t z_H(t,s)dL_s\\
=&z_H(t,t)L_t-z_H(t,\epsilon)L_\epsilon-\int_{\epsilon}^t \left ( \frac{d}{ds}z_H(t,s)\right )L_sds\\
=&-\int_0^t \left ( \frac{d}{ds}z_H(t,s)\right )L_sds.
\end{align*}
\end{proof}
The problem with the pathwise construction of fLpMG (when not in compound Poisson case) is that for $H>\frac{1}{2}$, the Molchan-Golosov kernel $z_H(t,\cdot)$ does not vanish at the origin like the Mandelbrot-Van Ness kernel does. However, the paths of the fLpMG are continuous when $H>\frac{1}{2}$ as is illustrated by the following theorem.
\begin{prop}
\begin{enumerate}
\item For $H>\frac{1}{2}$, an fLpMG $Y$ on $[0,T]$ has a. s. H\"older continuous paths of any order strictly less than $H-\frac{1}{2}$.
\item For $H<\frac{1}{2}$, an fLpMG $Y$ has discontinuous sample paths with positive probability.
\item For $H<\frac{1}{2}$, an fLpMG $Y$ has unbounded sample paths with positive probability.
\end{enumerate}
\end{prop}
\begin{proof}
Let $H>\frac{1}{2}$. It holds that
\begin{equation*}
\E |Y_t-Y_s|^2=|t-s|^{2H}=|t-s|^{1+2(H-\frac{1}{2})}.
\end{equation*}
The first assertion follows now from the Kolmogorov-Chentsov theorem. See for example~\cite{karatzasshreve}.
Let now $H<\frac{1}{2}$. We know that in this case the mapping $t\mapsto z_H(t,s)$ is unbounded and discontinuous for all $s\in(0,T)$. Thus by theorem~4 of~\cite{rosinski2} we know that the sample paths of $Y$ are unbounded with positive probability and also discontinuous with positive probability.
\end{proof}
\begin{rem}
Analogously one can prove that an fLpMvN has unbounded and discontinuous paths with positive probability when $H<\frac{1}{2}$.
\end{rem}
Besides continuity, the sample paths have also the zero quadratic variation property for $H>\frac{1}{2}$. This is illustrated in the following theorem where we compute the quadratic variation over the dyadic sequence of partitions.
\begin{theor}
\label{flpmgqv}
Let $(Y_t)_{t_{\geq 0}}$ be a fLpMG with $H>\frac{1}{2}$. Then for all $t>0$ it holds that
\begin{equation*}
\sum_{j=1}^{2^n} \left ( Y_{\frac{j}{2^n}t}-Y_{\frac{j-1}{2^n}t}\right )^2\rightarrow 0\quad \text{a. s. when } n \rightarrow \infty.
\end{equation*}
\end{theor}
\begin{proof}
Set
\begin{equation*}
V_n=\sum_{j=1}^{2^n}\left ( Y_{\frac{j}{2^n}t}-Y_{\frac{j-1}{2^n}t}\right )^2.
\end{equation*}
Now we have that
\begin{equation*}
\E V_n=(\E L_1^2) 2^n \left ( \frac{t}{2^n}\right )^{2H}=(\E L_1^2) t^{2H}2^{-n(2H-1)}.
\end{equation*}
We obtain using Markov inequality (see e.g.~\cite{jacodprotter}) that
\begin{align*}
&\sum_{n=1}^\infty \Pro\left (|V_n-\E V_n|\geq \frac{1}{n}\right )\leq \sum_{n=1}^\infty n\E |V_n-\E V_n|\\\leq &\sum_{n=1}^\infty n 2 \E V_n=2t^{2H} (\E L_1^2) \sum_{n=1}^\infty n 2^{-n(2H-1)}<\infty.
\end{align*}
We use now Borel-Cantelli theorem (see~\cite{jacodprotter}) and obtain that
\begin{equation*}
V_n -\E V_n\rightarrow 0, \quad \text{as } n \rightarrow \infty \text{ a. s.}
\end{equation*}
On the other hand $\E V_n \rightarrow_{n\rightarrow \infty}0$. Thus $V_n\rightarrow_{n\rightarrow \infty}0$ a. s.
\end{proof}
Note that the same proof works also in the case of fLpMvN. Thus, we have the following theorem.
\begin{theor}
\label{flpmvnqv}
Let $X$ be an fLpMvN with $H>\frac{1}{2}$. Then for $-\infty<s<t<\infty$ it holds that
\begin{equation*}
\sum_{j=1}^{2^n} \left ( X_{s+\frac{j}{2^n}(t-s)}-X_{s+\frac{j-1}{2^n}(t-s)}\right )^2 \rightarrow 0 \quad \text{a.s. when }n \rightarrow \infty.
\end{equation*}
\end{theor}
\begin{prop}[Characteristic function]
\label{mgflpcf}
Let $u_1,\dots,u_n \in \R$ and $0<t_1<\dots<t_n<\infty$. Then
\begin{equation*}
\E \left ( \exp \left ( i\sum_{j=1}^n u_j Y_{t_j}\right )\right )= \exp \left ( \int_\R \Psi \left ( \sum_{j=1}^n u_j z_H(t_j,s)\right )ds\right ),
\end{equation*}
where $\Psi$ is the characteristic exponent of the driving L\'evy process $L$. Moreover, $Y_t$ is infinitely divisible for all $t\geq 0$.
\end{prop}
\begin{proof}
This follows, for example, from~\cite{rajput}.
\end{proof}
\begin{prop}
The increments of fLpMG are not always stationary.
\end{prop}
\begin{proof}
Consider fLpMG $Y$ driven by a compound Poisson process $L$ with L\'evy measure $\nu=\frac{1}{2}(\delta_1+\delta_{-1})$, where $\delta_\cdot$ denotes the Dirac delta. For $\epsilon>0$
\begin{equation*}
\Pro(Y_\epsilon-Y_0=0)\geq \Pro(L_t=0 \quad\forall t\leq \epsilon)= 1-e^{-\epsilon}.
\end{equation*}
On the other hand, consider set
\begin{equation*}
A=\{\omega \in \Omega \text{ s.t. } \#\{s\in[0,1]\text{ s.t. }\Delta L_s\neq 0\}=1 \text{ and } \#\{s\in (1,\epsilon] \text{ s.t. }\Delta L_s\neq 0\}=0\}.
\end{equation*}
In set $A$, there is one jump time $S\in[0,1]$ and $\Delta L_S=\pm 1$. It follows that
\begin{equation*}
Y_{1+\epsilon}-Y_1=\left ( \frac{d}{ds}z_H(1+\epsilon,s)-\frac{d}{ds}z_H(1,s)\right )_{s=S}\Delta L_S\neq 0,
\end{equation*}
by proposition~\ref{pathwise}. We also have that
\begin{equation*}
\Pro(Y_{1+\epsilon}-Y_1=0)\leq 1-\Pro(A)=1-e^{-1}(1-e^{-\epsilon})<1-e^{-\epsilon},
\end{equation*}
when $\epsilon>0$ small enough. Thus, $Y_\epsilon-Y_0\stackrel{d}{\neq}Y_{1+\epsilon}-Y_1$. Hence, the increments of fLpMG $Y$ are not stationary.
\end{proof}
\begin{theor}[Self-similarity]
Fractional L\'evy process by Molchan-Golosov transformation cannot be self-similar for $H\in(0,1)$.
\end{theor}
\begin{proof}
Assume that the process $Y$ is self-similar with some index $\alpha$. Then we have for all $c>0$ that
\begin{equation*}
\left (Y_{ct}\right)_{t\geq 0}\stackrel{d}{=}\left(c^\alpha Y_t\right)_{t\geq 0}.
\end{equation*}
The characteristic function of $Y$ is given by theorem~\ref{mgflpcf}. On the other hand
\begin{align*}
&\E e^{iuY_t}=\E e^{iuc^{-\alpha}Y_{ct}}\\
=&\exp\left ( \int_0^{ct} \int_{\R}\left ( e^{ic^{-\alpha}uxz_H(ct,s)}-1-ic^{-\alpha}uxz_H(ct,s)\right)\nu(dx)ds \right )\\
=&\exp\left ( \int_0^{t}\int_{\R}\left ( e^{ic^{-\alpha}c^{H-\frac{1}{2}}ux z_H(t,s)}-1-ic^{-\alpha}c^{H-\frac{1}{2}}uxz_H(t,s)\right )\nu(dx)cds\right )\\
=&\exp\left(\int_0^t \int_{\R}\left ( e^{iuxz_H(t,s)}-1-iuxz_H(t,s)\right )c\nu\left(c^{\alpha-H+\frac{1}{2}}dx\right)ds\right ).
\end{align*}
Note that $\alpha-H+\frac{1}{2}=0$ implies that $Y\stackrel{d}{=}cY$ for all $c>0$ which means that $Y=L=0$ identically. We define for $r>0$ the translation operator $T_r$ of measures on $\R$ for $B \in \mathcal{B}(\R)$ by
\begin{equation*}
(T_r\nu)(B)=\nu\left(r^{-1}B\right).
\end{equation*}
Now the L\'evy measure of infinitely divisible random variable $c^{-\alpha}Y_{ct}$ is given by $cT_r\nu_{Y_t}$, where $r=c^{-H+\frac{1}{2}+\alpha}$ and
\begin{equation*}
\nu_{Y_t}(B)=\int_0^t\int_{\R}1_B\left ( z_H(t,s)x\right )\nu(dx)ds.
\end{equation*}
The drift parameter of $Y_1$ is $\gamma_{Y_1}$. It follows from the uniqueness of the generating triplet and self-similarity property that it holds for all $b>0$
\begin{equation*}
\nu_{Y_t}=b^{1/\left(H-\frac{1}{2}-\alpha\right)}T_b\nu_{Y_t}.
\end{equation*}
Denote now $\beta=\frac{-1}{H-\frac{1}{2}-\alpha}$. Let $\mu$ be the distribution of random variable $Y_1$ and let $\hat{\mu}(u)$ be the characteristic function of $\mu$. Random variable $bY_1$ has now characteristic function $\hat{\mu}(bu)$. Because $Y_1$ is infinitely divisible, we can use proposition~11.10~of~\cite{sato} that the triplet of $bY_1$ is $(\gamma(b),0,T_b\nu_{Y_1})$ for some $\gamma(b)$. On the other hand, $\hat{\mu}(u)^{b^\beta}$ is an infinitely divisible characteristic function with triplet $(b^\beta\gamma_{Y_1},0,b^\beta\nu_{Y_1})$. Thus we have for any $b>0$ some $d$ such that
\begin{equation*}
\hat{\mu}(u)^{a}=\hat{\mu}(bu)e^{idu},
\end{equation*}
where $a=b^\beta$. We note that $\R_+\mapsto \R_+: b\mapsto b^\beta$ is one-to-one. Thus, $\mu$ follows a stable law with index $\beta$. The index $\beta \in (0,2]$ by definition~13.5.~of~\cite{sato}. By theorem~14.1.~of~\cite{sato}, $\beta=2$ corresponds to Gaussian case and is thus impossible.
It follows now that $\E Y_t^2=\infty$, which contradicts the fact that $Y_t\in L^2(\Omega,\Pro)$. Thus, fLpMG can never be self-similar of any order $\alpha$.
\end{proof}
\begin{rem}
In~\cite{bender}, the authors define fractional subordinators by Molchan-Golosov transformation using pathwise Riemann-Stieltjes integration. However, these processes are not fLpMG as considered here, since subordinators are increasing L\'evy processes and here we consider only zero mean L\'evy processes. Also the integration concept there is different.
\end{rem}
\section{Relation of the two fLp concepts}
The connection between fractional L\'evy processes by Molchan-Golosov transformation and Mandelbrot-Van Ness transformation is basically the same as in the fBm case. The result in fLp case is new.
Let $H\in\left( 0,1\right )$ and $L$ be a two-sided L\'evy process without Brownian component satisfying $\E L_1=0$ and $\E L_1^2<\infty$. Let $s>0$ and set
\begin{equation*}
Y^s_t=\int_0^tz_H(t,u)dL_{u-s}=c_H\int_0^t(t-u)^{H-\frac{1}{2}}\tilde{F}\left ( \frac{u-t}{u}\right)dL_{u-s}, \quad t \in [0,\infty),
\end{equation*}
which is in fact fLpMG with Hurst parameter $H$. Here
\begin{equation*}
\tilde{F}(x)=F\left ( \frac{1}{2}-H,H-\frac{1}{2},H+\frac{1}{2},x\right),
\end{equation*}
where $F$ is the Gauss' hypergeometric function. Define the time shifted process
\begin{equation*}
Z^s_t=Y^s_{t+s}-Y^s_s, \quad t\in[-s,\infty).
\end{equation*}
In the fBm case this would also be fBm, but in fLpMG case we do not have the stationarity of the increments and we are lacking such an interpretation. Now we substitute $v=u-s$ and obtain a.s. that
\begin{equation*}
Z^s_t=c_H\left ( \int_{-s}^t (t-v)^{H-\frac{1}{2}}\tilde{F}\left ( \frac{v-t}{v+s}\right )dL_v-\int_{-s}^0 (-v)^{H-\frac{1}{2}}\tilde{F}\left( \frac{v}{v+s}\right )dL_v\right ).
\end{equation*}
By~\cite{jost2}, $\tilde{F}(0)=1$ and thus we obtain formally as $s\rightarrow \infty$ that
\begin{equation*}
Z^\infty_t:=\frac{c_H}{C_H}X_t:=c_H \int_\R \left ( (t-v)_+^{H-\frac{1}{2}}-(-v)_+^{H-\frac{1}{2}}\right )dL_v, \quad t\in \R.
\end{equation*}
\begin{theor}
\label{flpmvnflpmgconnection}
For every $t\in \R$ there exist constants $S,C>0$ such that
\begin{equation*}
\E \left ( Z^s_t-Z^\infty_t\right )^2\leq Cs^{2H-2}, \quad \text{for } s>S.
\end{equation*}
\end{theor}
\begin{proof}
We obtain
\begin{align*}
&\frac{1}{c_H^2\E L_1^2}\E \left ( Z^s_t-Z^\infty_t\right )^2\\
=&\int_\R \left(\left((t-v)^{H-\frac{1}{2}}-(-v)^{H-\frac{1}{2}}\right)1_{(-\infty,-s)}(v)\right)^2dv\\
+&\int_\R \left((t-v)^{H-\frac{1}{2}}\left ( \tilde{F}\left ( \frac{v-t}{v+s}\right )-1\right)1_{(-s,t)}(v)\right.\\
-&\left.(-v)^{H-\frac{1}{2}}\left ( \tilde{F}\left ( \frac{v}{v+s}\right)-1\right)1_{(-s,0)}(v)\right)^2dv,
\end{align*}
by $L^2$- isometry and independence of increments of $L$. The claim follows now from the proof of Theorem~3.1.~of~\cite{jost2}.
\end{proof}
\section{Wiener integration}
Here our goal is to define suitable Wiener integrals with respect to fLpMG. In contrary to the case of fLpMvN, we use the fractional integration on a compact interval instead of the whole real line. We will define the space $L^2_H([0,T])$ of integrands as in the case of compact interval Wiener integrals in fBm case.
Let $g$ be a function defined on $[0,T]$ and $I_-^{H-\frac{1}{2}}$ be the right-sided Riemann-Liouville integral operator of order $H-\frac{1}{2}$ as in~\cite{jost}. Define operator
\begin{equation*}
(K^Hg)(s)=\Gamma(H+\frac{1}{2})c_H s^{\frac{1}{2}-H}\left ( I_-^{H-\frac{1}{2}}\left ((\cdot)^{H-\frac{1}{2}}g(\cdot)\right )\right )(s), \quad s,t\in[0,T].
\end{equation*}
Now it holds by~\cite{jost} that
\begin{equation*}
z_H(t,s)=\Gamma(H+\frac{1}{2})c_Hs^{\frac{1}{2}-H}\left ( I_-^{H-\frac{1}{2}} (\cdot)^{H-\frac{1}{2}}1_{[0,t)}\right )(s),
\end{equation*}
for $0\leq t\leq T$. Define now the space
\begin{equation*}
L_H^2([0,T])=\{g \in L^1([0,T])| K^Hg\in L^2([0,T])\},
\end{equation*}
equipped with norm $||g||_{L_H^2([0,T])}:=||K^Hg||_{L^2([0,T])}$. Now we are ready for the definition of Wiener integral.
\begin{defin}[Wiener integral for fLpMG]
\label{wiflpmg}
Let $H\in(0,1)$, $Y$ be a fLpMG driven by L\'evy process $L$. For $g\in L^2_H([0,T])$ the Wiener integral with respect to fLpMG is defined as
\begin{equation*}
\int_0^T g(s)dY_s=\int_0^T (K^H g)(s)dL_s.
\end{equation*}
\end{defin}
Note that the definition is completely analogous to the definition of compact interval Wiener integrals in the fBm setup. Now, let $g$ be a step function, which means that
\begin{equation}
\label{stepfunction}
g(s)=\sum_{j=1}^n a_j 1_{(s_{j-1},s_j]}(s),
\end{equation}
where $a_0,\dots,a_n\in \R$ and $0=s_0<s_1<\dots<s_n=T$. Now $g \in {L}^{2}_{H}([0,T])$ and we have the following result.
\begin{lemma}
Assume $H\in(0,1)$, $Y$ is fLpMG driven by $L$ and $g$ a step function defined by equation~(\ref{stepfunction}). It holds that
\begin{equation*}
\int_0^T g(s)dY_s= \sum_{j=1}^n a_j \left(Y_{s_{j}}-Y_{s_{j-1}}\right).
\end{equation*}
\end{lemma}
\begin{proof}
It is clear from the definition that the integral of a step function is linear. We will prove the claim for indicator functions. The general claim follows from the linearity. Set $g(s)=1_{(s_1,s_2]}$. Now
\begin{align*}
&\int_0^T g(s)dY_s=\int_0^T \left ( K^H g\right )(s)dL_s=\int_0^T \left (K^H\left(1_{(0,s_2]}-1_{(0,s_1]}\right)\right )(s)dL_s\\
=&Y_{s_2}-Y_{s_1}.
\end{align*}
\end{proof}
Obviously the following isometry holds for a step function $g$
\begin{equation}
\label{mgintisometry}
||\int_0^T (K^Hg)(s)dL_s||_{L^2(\Pro)}^2=\E L_1^2\int_0^T (K^Hg)^2(s)ds=\E L_1^2 \cdot ||g||^2_{{L}^{2}_{H}([0,T])}.
\end{equation}
Next we restrict ourselves to the case $H\in\left(0,\frac{1}{2}\right)$. In this case $L^2_H([0,T])$ is complete (\cite{sottinen}) and the step functions are dense in $L^2_H([0,T])$. Thus we can make the following alternative definition. Note that both the definitions yield the same Wiener integral.
\begin{defin}
\label{altdef}
Let $Y$ be a fLpMG with driving L\'evy process $L$ and Hurst index $H\in\left(0,\frac{1}{2}\right )$. Let $g\in L^2_H([0,T])$ and let $\{g_k\}_{k\in \Z_+}$ be a sequence of step functions converging to $g$ in $L^2_H([0,T])$. We define the Wiener integral of $g$ with respect to $Y$ as follows
\begin{equation*}
\int_0^T g(s)dY_s=L^2(\Pro)-\lim_{k\rightarrow \infty} \int_0^T g_k(s)dY_s.
\end{equation*}
\end{defin}
Note that the definition does not depend on the approximating sequence.
\section{Financial application}
Next we will construct an arbitrage free model including fractional L\'evy processes. This is a (geometric) mixed Brownian motion and fractional L\'evy process model. The no-arbitrage result is analogous to the result in the case of mixed Brownian motion and fractional Brownian motion.
In the following, $Z$ may be either fLpMG or fLpMvN with $H>\frac{1}{2}$ and $W$ is an ordinary Brownian motion independent of $Z$. Let $\sigma,\epsilon>0$. Define the mixed process by
\begin{equation}
\label{Uprocess}
U_t=\sigma Z_t +\epsilon W_t.
\end{equation}
\begin{theor}
\label{noarbitrage}
Let the market model be given by $(\Omega,\mathcal{F},\exp{U},(\mathcal{F}_t^U),\Pro)$ and let $\Phi$ be a stopping-smooth trading strategy, where we use the conventions of~\cite{bendersottinenvalkeila2}. Then $\Phi$ is not an arbitrage opportunity.
\end{theor}
\begin{proof}
We will check the assumptions of Theorem~5 of~\cite{bendersottinenvalkeila2} and then we are done. The two conditions to be checked, are the quadratic variation property and the conditional small ball property.
Both fLpMvN and fLpMG are continuous path processes with zero quadratic variation over the dyadic partitions, see Theorems~\ref{flpmgqv} and~\ref{flpmvnqv}. Thus $U$ has the quadratic variation of Brownian motion over these partitions.
Moreover, $U$ has conditional full support (CFS) w.r.t. its own filtration by Theorem~3.1 ~of~\cite{pakkanen}. Since $U$ has CFS w.r.t. $(\mathcal{F}^U_t)$ on $\R$, it follows that $\exp{U}$ has CFS w.r.t. $(\mathcal{F}^U_t)$ on $\R_+$. This is equivalent to the conditional small ball property of~\cite{bendersottinenvalkeila2} by Lemma~2.3 ~of~\cite{pakkanen}.
\end{proof}
The exact definition for stopping-smooth strategies is not given in this paper, because it is rather technical. According to~\cite{bendersottinenvalkeila2} the chosen strategies cover hedges for many European, lookback and Asian options. Thus, it is an economically meaningful class.
This mixed model is a natural way for modeling shocks in financial markets. The Brownian motion part corresponds to the ordinary noise in the market and the fractional L\'evy process part to sudden shocks in the market. On the other hand the fractional L\'evy process has the covariance structure of fBm, this allows to model for long-range dependence.
The no-arbitrage result holds for both fLp concepts, but from the modeling point of view they are different. If one wants to have stationary increments of $U$, one should use fLpMvN. If one wants to avoid history from $-\infty$, one should use fLpMG instead. In real world, there is always the time $0$ when the trading began. Hence fLpMG might be more natural choice. However, this modeling question is rather delicate.
If in the model~(\ref{Uprocess}), $Z$ is of bounded variation, then $U$ is a semimartingale with Brownian motion as the martingale part of the decomposition. However, this model has long-range dependence property.
\section{Proofs}
\label{mainproof}
First we prove a lemma about the connection of the normalizing constants of the different integral representations.
\begin{lemma} For any $H\in(0,1)$
\begin{equation*}
C_H=c_H.
\end{equation*}
\end{lemma}
\begin{proof}
First of all by~\cite{mishura}
\begin{equation*}
C_H=\frac{(2H \sin{(\pi H)}\Gamma{(2H)})^\frac{1}{2}}{\Gamma{(H+\frac{1}{2})}}.
\end{equation*}
Now we have that
\begin{align*}
&C_H-c_H=\frac{1}{\Gamma(H+\frac{1}{2})}\left ( (2H\sin{(\pi H)}\Gamma(2H))^\frac{1}{2}-\left (\frac{2H\Gamma(H+\frac{1}{2})\Gamma(\frac{3}{2}-H)}{\Gamma(2-2H)}\right )^{\frac{1}{2}}\right )\\
=&\frac{1}{\Gamma(H+\frac{1}{2})}\sqrt{\frac{2H}{\Gamma(2-2H)}}\left ( (\sin{(\pi H)}\Gamma(2H)\Gamma(2-2H))^\frac{1}{2}-(\Gamma(H+\frac{1}{2})\Gamma(\frac{3}{2}-H))^\frac{1}{2}\right ).
\end{align*}
For the difference we have now that
\begin{align*}
&(\sin{(\pi H)}\Gamma(2H)\Gamma(2-2H))^\frac{1}{2}-(\Gamma(H+\frac{1}{2})\Gamma(\frac{3}{2}-H))^\frac{1}{2}\\
=&(\sin{(\pi H)}(2H-1)\Gamma(2H-1)\Gamma(2-2H))^\frac{1}{2}-((H-\frac{1}{2})\Gamma(H-\frac{1}{2})\Gamma(\frac{3}{2}-H))^\frac{1}{2}\\
=&\sqrt{H-\frac{1}{2}}\left ( \left (2\sin{(\pi H)}\frac{\pi}{\sin{\pi(2H-1)}}\right )^\frac{1}{2}-\left( \frac{\pi}{\sin{\pi(H-\frac{1}{2})}}\right)^\frac{1}{2}\right )\\
=&\sqrt{\pi (H-\frac{1}{2})}\left ( \left ( -\frac{2\sin{(\pi H)}}{\sin{(2\pi H)}}\right )^\frac{1}{2}-\left ( \frac{-1}{\cos{\pi H}}\right )^\frac{1}{2}\right )\\
=&\sqrt{\pi (H-\frac{1}{2})\sin{2\pi H}\cos{(\pi H)}}\left ( (-2\sin{(\pi H)\cos{(\pi H)}})^\frac{1}{2}-(-\sin(2\pi H))^\frac{1}{2}\right )=0.
\end{align*}
The previous computation is for $H>\frac{1}{2}$, but an analogous computation goes through for $H<\frac{1}{2}$ as well.
\end{proof}
Next we present some results about finiteness of the moments of different kernels.
\begin{lemma}
\label{ikm}
Let $H>\frac{1}{2}$ and $K>2$. Then for any $t>0$
\begin{equation*}
\int_0^t (z_H(t,s))^{K}ds=\infty,
\end{equation*}
when $H\geq \frac{1}{2}+\frac{1}{K}$.
\end{lemma}
\begin{proof}
\begin{align*}
&(H-\frac{1}{2})^{-K}c_H^{-K}\int_0^t (z_H(t,s))^{K}ds
=\int_0^t s^{K(\frac{1}{2}-H)}\left ( \int_s^t u^{H-\frac{1}{2}}(u-s)^{H-\frac{3}{2}}du\right )^{K}ds\\
\geq&\int_0^t s^{K(\frac{1}{2}-H)}\left ( \int_s^t (u-s)^{H-\frac{1}{2}}(u-s)^{H-\frac{3}{2}}\right )^K ds\\
=&\int_0^t s^{K(\frac{1}{2}-H)}\left ( \frac{1}{2H-1}(t-s)^{2H-1}\right )^Kds.
\end{align*}
We note that the factor
\begin{equation*}
\left ( \frac{1}{2H-1}(t-s)^{2H-1}\right )^K
\end{equation*}
is bounded and also bounded away from zero in some neighborhood of the origin. Thus the last integral is finite if and only if
\begin{equation*}
\int_0^t s^{K(\frac{1}{2}-H)}ds<\infty.
\end{equation*}
Thus the integral $\int_0^t (z_H(t,s))^Kds=\infty$ if $K(\frac{1}{2}-H)\leq 1$ i.e. $H\geq \frac{1}{2}+\frac{1}{K}$.
\end{proof}
\begin{lemma}
\label{fkm}
For $H>\frac{1}{2}$ and any $t>0$
\begin{equation*}
\int_{-\infty}^t (f_H(t,s))^Kds<\infty, \quad K\geq 2.
\end{equation*}
\end{lemma}
\begin{proof} From self-similarity, is sufficient to consider only $t=1$. We have that
\begin{align*}
&C_H^{-K}\int_0^1 (f_H(1,s))^Kds=\int_{-\infty}^0 \left ( (1-s)^{H-\frac{1}{2}}-(-s)^{H-\frac{1}{2}}\right )^Kds+\frac{1}{K(H-\frac{1}{2})+1}.
\end{align*}
For the first term we get
\begin{align*}
&\int_{-\infty}^0 \left ( (1-s)^{H-\frac{1}{2}}-(-s)^{H-\frac{1}{2}}\right )^Kds=\int_0^{\infty}\Big((1+s)^{H-\frac{1}{2}}-s^{H-\frac{1}{2}}\Big)^Kds\\
=& \int_0^1 z^{K(\frac{1}{2}-H)}((1+z)^{H-\frac{1}{2}}-1)^Kz^{-2}dz=\int_0^1 z^{K(\frac{1}{2}-H)-2}((1+z)^{H-\frac{1}{2}}-1)^Kdz.
\end{align*}
By the Lagrange theorem we have that for some $x\in[0,z]$
\begin{equation*}
(1+z)^{H-\frac{1}{2}}-1=\Big(H-\frac{1}{2}\Big)(1+x)^{H-\frac{3}{2}}z\leq \Big(H-\frac{1}{2}\Big)z.
\end{equation*}
Now we have that
\begin{align*}
&\int_0^1 z^{K(\frac{1}{2}-H)-2}((1+z)^{H-\frac{1}{2}}-1)^Kdz\leq\Big(H-\frac{1}{2}\Big)^K\int_0^1 z^{K(\frac{1}{2}-H)-2}z^Kdz\\
=&\Big(H-\frac{1}{2}\Big)^K \int_0^1 z^{K(\frac{3}{2}-H)-2}dz<\infty,
\end{align*}
since $K\left (\frac{3}{2}-H\right)-2>2\left(\frac{3}{2}-1\right)-2=-1.$
\end{proof}
\begin{rem} It follows from two above lemmas that for any $H\geq \frac{1}{2}+\frac{1}{K}$ and $t>0$ we have inequality
$\int_{-\infty}^t (f_H(t,s))^Kds<\int_0^t (z_H(t,s))^Kds$.
\end{rem}
Now we want to establish similar inequalities for $K=3, 4$ and $\frac{1}{2}< H< \frac{1}{2}+\frac{1}{K}$.
\begin{lemma}
\label{mgbound}
1) For any $\frac{1}{2}< H< \frac{5}{6}$ and any $t>0$ we have the inequality
$\int_{-\infty}^t (f_H(t,s))^3ds<\int_0^t (z_H(t,s))^3ds$.
2) For any $\frac{1}{2}< H< \frac{3}{4}$ and any $t>0$ we have the inequality
$\int_{-\infty}^t (f_H(t,s))^4ds<\int_0^t (z_H(t,s))^4ds$.
\end{lemma}
\begin{proof} The proof is similar in both the cases so consider the first one. It is better to normalize the integrals, and for the normalized Molchan-Golosov kernel we obtain that
\begin{align*}
&\frac{1}{C_H^3}\int_0^1(z_H(1,s))^3ds\\
=&\Big(H-\frac{1}{2}\Big)^3\int_0^1 s^{\frac32-3H}\left (\int_s^1 u^{H-\frac{1}{2}}(u-s)^{H-\frac{3}{2}}du\right)^3ds.\\
\end{align*}
Note that integration by parts leads to the equality
\begin{equation*}
\Big(H-\frac{1}{2}\Big)\int_s^1 u^{H-\frac{1}{2}}(u-s)^{H-\frac{3}{2}}du=(1-s)^{H-\frac{1}{2}}-\Big(H-\frac{1}{2}\Big)\int_s^1 u^{H-\frac{3}{2}}(u-s)^{H-\frac{1}{2}}du.
\end{equation*}
Further, for $s\leq u\leq 1$ we have that
\begin{equation*}
(u-s)^{H-\frac{1}{2}}=u^{H-\frac{1}{2}}\Big(1-\frac{s}{u}\Big)^{H-\frac{1}{2}}\leq u^{H-\frac{1}{2}}(1-s)^{H-\frac{1}{2}},
\end{equation*}
whence
\begin{align*}
&(1-s)^{H-\frac{1}{2}}-\Big(H-\frac{1}{2}\Big)\int_s^1 u^{H-\frac{3}{2}}(u-s)^{H-\frac{1}{2}}du\\\geq&(1-s)^{H-\frac{1}{2}}
\Big(1-\Big(H-\frac{1}{2}\Big)\int_s^1u^{2H-2}du\Big)=\frac{1}{2}(1-s)^{H-\frac{1}{2}}(1+s^{2H-1}).
\end{align*}
Denote $p=H-\frac12, \, \frac12< H < \frac56$ and $\alpha_p^1=\frac{1}{C_H^3}\int_0^1(z_H(1,s))^3ds.$
Then $p\in(0,\frac13)$ and we obtain from previous estimates that
\begin{align*}
&\alpha_p^1\geq\frac{1}{8}\int_0^1s^{-3p}(1-s)^{3p}(1+s^{2p})^3ds\\
=&\int_0^1s^{-3p}(1-s)^{3p}s^{3p}ds+\frac{1}{8}\int_0^1s^{-3p}(1-s)^{3p}((1+s^{2p})^3-8s^{3p})ds\\
&=\int_0^1(1-s)^{3p}ds+\frac{1}{2}\int_0^1s^{-3p}(1-s)^{3p}(1-s^p)^{2}\Big(\frac{1+2s^{2p}+s^{4p}}{4}+\frac{s^p+s^{3p}}{2}+s^{2p}\Big)ds\\
&=:\alpha_p^{11}+\alpha_p^{12}.
\end{align*}
Evidently, $\alpha_p^{11}=\frac{1}{3p+1}$. For $\alpha_p^{12}$
we use the following simple bounds: on the interval $s\in[0,1]$ $1\geq s^{kp}, k=1,2,3,4$
and for $p\in(0,\frac13),\,s\in[0,1]$ $(1-s)^{3p}\geq 1-s^{3p}.$
Therefore, \begin{align*}
&\alpha_p^{12}\geq\frac32\int_0^1s^p(1-s^{3p})(1-s^p)^{2}ds\\
&=\frac32\int_0^1(s^p-2s^{2p}+s^{3p}-s^{4p}+2s^{5p}-s^{6p})ds\\
&=\frac32\Big(\frac{1}{p+1}-\frac{2}{2p+1}+\frac{1}{3p+1}-\frac{1}{4p+1}+\frac{2}{5p+1}-\frac{1}{6p+1}\Big)\\
&=9p^3\frac{38p^2+21p+3}{\prod_{k=1}^6(kp+1)}.
\end{align*}
On the other hand, for the normalized Mandelbrot-Van Ness kernel we can use the same reasonings as in the proof of Lemma \ref{fkm} and obtain in the above notations that
\begin{align*}
&\alpha_p^2:=C_H^{-3}\int_0^1 (f_H(1,s))^3ds\\
&=\int_{-\infty}^0 \left ( (1-s)^{p}-(-s)^{p}\right )^3ds+\frac{1}{3p+1}\\
&=\int_0^1 s^{-3p-2}((1+s)^{p}-1)^3ds+\frac{1}{3p+1}\\
&\leq p^3\int_0^1 s^{-3p+1}ds+\frac{1}{3p+1}=\frac{p^3}{2-3p}+\frac{1}{3p+1}.
\end{align*}
To establish the inequality $\alpha_p^{1}>\alpha_p^{2}$, that is equivalent to the statement of the lemma, it is sufficient to prove that
$$9p^3\frac{38p^2+21p+3}{\prod_{k=1}^6(kp+1)}>\frac{p^3}{2-3p}, \,p\in \Big(0, \frac13\Big),$$
or $$9\frac{38p^2+21p+3}{\prod_{k=1}^6(kp+1)}>\frac{1}{2-3p}, \,p\in \Big(0, \frac13\Big).$$
For technical simplicity, diminish the left-hand side and compare the functions $f_1(p)=9\frac{36p^2+21p+3}{\prod_{k=1}^6(kp+1)}=27\frac{12p^2+7p+1}{\prod_{k=1}^6(kp+1)}$ and $f_2(p)=\frac{1}{2-3p}$. Of course, to finish the proof, it is sufficient to establish the inequality$f_1(p)\geq f_2(p),\,p\in(0,\frac13)$.
Note that $f_2(p)$ increases from $\frac12$ to $1$. The derivative of $f_1(p)$ can be estimated, up to positive multiplier, as
\begin{align*}
&(24p+7)\prod_{k=1}^6(kp+1)-(12p^2+7p+1)\sum_{r=1}^6 r\prod_{k=1, k\neq r}^6(kp+1)\\
&\leq \prod_{k=1}^5(kp+1)((24p+7)(6p+1)-21(12p^2+7p+1))\\
&=-\prod_{k=1}^5(kp+1)(108p^2+81p+14)<0.
\end{align*}
It means that $f_1(p)$ decreases on $(0,\frac13),$ and it decreases from $27$ to $\frac{243}{160}>1$ that completes the proof.
\end{proof}
\begin{rem} The integral with respect to Molchan--Golosov kernel can be bounded from below in terms of Beta- or Gamma-functions. These estimates are more sharp that obtained in the proof of Lemma \ref{mgbound}; however, we can not proceed with them otherwise than numerically.
Indeed, for example, in the case $K=4$ we can estimate \begin{align*}
&\frac{1}{C_H^4}\int_0^1(z_H(1,s))^4ds\geq\frac{1}{16}\int_0^1s^{-4p}(1-s)^{4p}(1+s^{2p})^4ds\\
&= \frac{1}{16}B(3-4H,4H-1)
+\frac{1}{4}B(2-2H,4H-1)+\frac{1}{4}B(2H,4H-1)\\&+\frac{1}{16}B(4H-1,4H-1)+\frac{3}{8}\frac{1}{4H-1}=:g_1(p).
\end{align*}
As before, $\frac{1}{c_H^4}\int_0^1(f_H(1,s))^4ds\leq\frac{p^4}{5-4p}+\frac{1}{4p+1}=g_2(p).$
The difference $g_1(p)-g_2(p), p=H-\frac12,$ is presented on Figure \ref{dif} in terms of Hurst parameter $H$.
\begin{figure}
\begin{center}
\includegraphics[width=6cm]{difference2.eps}
\caption{The difference of for the integrals of Molchan-Golosov and Mandelbrot-van-Ness kernels for $K=4$ as the function of $H$.}
\label{dif}
\end{center}
\end{figure}
\end{rem}
Now we are ready for the proof of the main result, i.e., Theorem \ref{mainthm}.
\begin{proof}
With assumptions $\E L_1=0$ and $\E L_1^2<\infty$ we can write the characteristic exponent of the driving L\'evy process as follows
\begin{equation*}
\Psi(t)=-\frac{1}{2}\sigma^2 u^2+\int_{\R} \left ( e^{iux}-1-iux \right )\nu(dx).
\end{equation*}
Prove only the second statement, the first one can be proved in a similar way. We use the representation formula for the characteristic function of fractional L\'evy process and get that the fourth cumulant of the fLpMG $Y$ is given by
\begin{equation*}
\int_0^t (z_H(t,s))^4ds \int_\R x^4 \nu(dx).
\end{equation*}
Analogously the fourth cumulant of the fLpMvN $X$ is
\begin{equation*}
\int_{-\infty}^t (f_H(t,s))^4ds \int_\R x^4\nu(dx).
\end{equation*}
Note that iff the L\'evy measure $\nu$ is nondegenerate, $\int_\R x^4\nu(dx)>0$.
Our aim is now to prove that with the assumption $\E L_1^4<\infty$ the fourth cumulants of different fLp's are different. This will prove that the different fractional L\'evy processes have different distributions.
For $H\geq \frac{3}{4}$ the 4th cumulant of fLpMG $Y$ is infinite by lemma~\ref{ikm}. On the other hand the corresponding cumulant for fLpMvN $X$ is finite by lemma~\ref{fkm}.
The case $H\in(\frac{1}{2},\frac{3}{4})$ uses Lemma \ref{mgbound}, and we immediately obtain the proof.
\end{proof}
Next we proof Proposition~\ref{mainimproved}.
\begin{proof}
Let $L$ be a compound Poisson process with parameter $\lambda$ s.t. $\E L_1=0$ and $\E L^2_1<\infty$. We have for fLpMG $Y$ by proposition~\ref{pathwise} $\Pro (Y_t=0)\geq e^{-\lambda t}$ for $t>0$. On the other hand, for $t>0$ we can decompose fLpMvN $X$ to two independent components
\begin{equation*}
\int_{-\infty}^{-1}f_H(t,s)dL_s+\int_{-1}^tf_H(t,s)dL_s.
\end{equation*}
Probability that $L$ jumps exactly once on $[-1,t)$ is $\lambda(1+t)e^{-\lambda (1+t)}$ and probability that $L$ does not have jumps on $[-1,t)$ is $e^{-\lambda(1+t)}$. Now it is easy to see that
\begin{equation*}
\Pro (X_t=0)\leq \lambda(1+t)e^{-\lambda (1+t)} \wedge e^{-\lambda(1+t)} < e^{-\lambda t}\leq\Pro(Y_t=0)
\end{equation*}
for $t>0$ small enough.
\end{proof}
\section*{Acknowledgements}
We would like to thank Esko Valkeila for his fruitful comments. H. Tikanm\"aki has been supported financially by \textit{The Finnish Graduate School in Stochastics and Statistics} and \textit{Academy of Finland}, grant 212875. Yu. Mishura was supported financially by Finnish Academy of Science and Letters, Vilho, Yrj\"o and Kalle V\"ais\"al\"a Foundation.
\bibliographystyle{plain}
|
\section{Introduction}
Wigner matrices have a ubiquitous presence in science; from the computation of molecular quantum states, through the description of solitons in particle physics and convolution of beam and sky algorithms in astronomy, they are needed to sometimes very high quantum numbers making fast and accurate algorithms that calculate them important. Other methods have been developed that calculate these matrices exactly but with sub-optimal performance to very high angular momenta~\cite{risbo}, or approximately but very efficiently~\cite{rowe}, but none that calculates them exactly and quickly to almost arbitrarily high angular momentum. Two such methods are presented in this paper and applied to a convolution algorithm between beam and sky. The following section gives some basic properties of Wigner matrices, and this is followed by a section describing the algorithm. The fourth section describes its application to convolution and a summary is presented at the end.
\section{Wigner matrices}
Wigner matrix elements\footnote{For a nice review of Wigner matrices, see \cite{varshalovich}.}, typically denoted by $\Dlmm$, are the eigenfunctions of the Schr\"{o}dinger equation for a symmetric top and form an irreducible basis of the Lie group SU(2), and the rotation group SO(3); the angles $\alpha$, $\beta$ and $\gamma$ are the Euler angles that define the orientation of the top. As basis functions of SU(2), the $\Dlmm$ satisfy the standard angular momentum relations
\begin{eqnarray}\label{llplus1}
\hat{\text{J}}^2\Dlmm &=& l(l+1)\Dlmm \\ \label{mproj}
\hat{\text{J}}_{z}\Dlmm &=& m \Dlmm \\ \label{mprojp}
\hat{\text{J}}_{z^\prime}\Dlmm &=& m^\prime \Dlmm ~,
\end{eqnarray}
where $l$ labels the irreducible representation of SU(2) and also corresponds to the quantum number representing the total angular momentum of the eigenfunction; $-l\le m,m^\prime\le l$ are the quantum numbers representing the projections of the total angular momentum on two $z$-axes rotated with respect to each other as described below.
The Euler angles are defined as three rotations: a rotation $\gamma$ about the $z$-axis that rotates the $x$ and $y$ axes $\to$ $x^\prime$ and $y^\prime$; this first rotation is followed by a rotation $\beta$ about the new $y^\prime$-axis rotating $x^\prime$ and $z$ axes $\to$ $x^{\prime\prime}$ and $z^\prime$; the final rotation $\alpha$ is about $z^\prime$. In the basis we are using as defined by \Eq{mproj} and \Eq{mprojp}, the operators $\hat{\text{J}}_{z}$ and $\hat{\text{J}}_{z^\prime}$ are diagonal and $\Dlmm$ has the form
\begin{eqnarray}\label{wignerdef}
\Dlmm = e^{-im\alpha}\dlmm e^{-im^\prime \gamma}
\end{eqnarray}
where
\begin{eqnarray}\label{reducedwignerdef}
\dlmm = \langle lm | \exp\left[ -i\frac{\beta}{\hbar} \hat{\text{J}}_y \right] | lm^\prime \rangle~.
\end{eqnarray}
$\dlmm$ is called the reduced Wigner matrix element and consists of the overlap of a spherical harmonic with another spherical harmonic that has been rotated by an angle $\beta$ about the y-axis.
The differential equation satisfied by $\dlmm$ is
\begin{eqnarray}\label{wigeq}
& &\frac{\text{d}^2 \dlmm}{\text{d}\beta^2} + \text{cot}\beta \frac{\text{d} \dlmm}{\text{d}\beta} +\\ \nonumber
& & ~~~~~~ +\left( \frac{2mm^\prime \cos\beta - m^2 - m^{\prime 2}}{\sin^2\beta} + l(l+1)\right) \dlmm = 0~.
\end{eqnarray}
From the Schr\"{o}dinger equation in \Eq{wigeq}, it is possible to extract 3-term recursion relations that relate reduced Wigner matrix elements that differ in their quantum numbers. In principle, it is possible to use such relations to calculate the $\dlmm$. 3-term recursion relations can be unstable, which limits their usefulness unless the potential pitfalls are identified and avoided. Two examples of these relations are
\begin{eqnarray}\label{mrec}
\frac{-m + m^\prime \cos\beta}{\sin\beta} \dlmm = \frac{1}{2}\sqrt{(l+m^\prime)(l-m^\prime +1)}d^l_{mm^\prime-1}(\beta) & & \nonumber
\\
+\frac{1}{2}\sqrt{(l-m^\prime)(l+m^\prime +1)}d^l_{mm^\prime+1}(\beta)~,& &
\end{eqnarray}
and
\begin{eqnarray}\label{lrec}
\left[ \cos\beta - \frac{mm^\prime}{l(l+1)}\right] \dlmm = \frac{\sqrt{(l^2-m^2)(l^2-{m^\prime}^2)}}{l(2l+1)}d^{l-1}_{mm^\prime}(\beta) & & \nonumber
\\
+\frac{\sqrt{[(l+1)^2-m^2][(l+1)^2-{m^\prime}^2]}}{(l+1)(2l+1)}d^{l+1}_{mm^\prime}(\beta)~.& &
\end{eqnarray}
Generally, 3-term recursion relations will have two linearly independent solutions, $f_n$ and $g_n$~\cite{recipes}; these solutions can be oscillatory or exponentially decreasing or increasing. In the non-oscillatory case, $f_n$ is the {\it minimal} solution if
\begin{eqnarray}
\frac{f_n}{g_n} \to 0 ~~~\text{as}~~~ n \to \infty~,
\end{eqnarray}
while $g_n$ is the {\it dominant} solution. For solutions to the Schr\"{o}dinger equation, exponentially increasing/decreasing solutions appear only in the region where a particle can not classically exist because of energy conservation, but where a wave function can be non-zero in quantum mechanics. In the case of a rigid rotor~\cite{edmonds}, the kinetic energy of a spherically symmetric rotor is:
\begin{eqnarray}\label{classic}
2IT = p_{\beta}^2 + \frac{1}{\sin^2\beta} (p_{\gamma}^2+p_{\alpha}^2 -2p_{\alpha}p_{\gamma}\cos\beta )
\end{eqnarray}
In classical mechanics, $p_{\beta}^2>0$. In quantum mechanics, the quantization of \Eq{classic} means substituting $p_{\alpha}\to -i\partial/\partial\alpha$, $p_{\beta}\to -i\partial/\partial\beta$ and $p_{\gamma}\to -i\partial/\partial\gamma$. These substitutions combined with an eigenfunction of the form (\ref{wignerdef}) and the additional substitution $2IT\to l(l+1)\Dlmm$ inferred from \Eq{llplus1} yields \Eq{wigeq} . Since $p_{\beta}^2$ corresponds to the first two terms of \Eq{wigeq}, one concludes that classically we would have
\begin{eqnarray}\label{classiccond}
l(l+1) + \frac{2mm^\prime \cos\beta - m^2 - m^{\prime 2}}{\sin^2\beta} \ge 0 ~ .
\end{eqnarray}
Wherever \Eq{classiccond} is not satisfied, the solutions will be exponentially suppressed or divergent. When solving the Schr\"{o}dinger equation for the physical solutions, the divergent solutions are simply put to zero. When using the 3-term recursion relations, the divergent solution can be 'sniffed' out because of round-off errors and the recursions quickly fail. One special case where this cannot happen is when $m,m^\prime=0$ where \Eq{classiccond} is always satisfied since $l\ge 0$. In that case, \Eq{lrec} is stable and can be used to calculate $d^l_{00}(\beta)$ to very high $l$ extremely accurately.
For the cases where $m,m^\prime\ne 0$, we can still use 3-term recursion relations provided we do so in the right direction in the quantum number being varied. For example, looking at \Eq{mrec}, one can either calculate each $\dlmm$ for increasing $m^\prime$ or decreasing $m^\prime$. In one direction, the divergent solution will be growing while in the other it will be shrinking. To determine the direction in which \Eq{mrec} is stable, one need only consider \Eq{classiccond}. Assume you are interested in evaluating all the reduced Wigner matrix elements $d^l_{0m^\prime}(\beta)$ for $0\le m^\prime \le l$ using \Eq{mrec}; you can choose to begin your recurrence with $d^l_{00}(\beta)$ and increasing $m^\prime$ or begin from $d^l_{0l}(\beta)$ and decreasing $m^\prime$. To use \Eq{mrec} in a stable manner, you need start from $d^l_{0l}(\beta)$ and decrease $m^\prime$. Putting $m=0$ in \Eq{classiccond} yields the new condition
\begin{eqnarray}\label{classiccondm0}
l(l+1) - \frac{ m^{\prime 2}}{\sin^2\beta} \ge 0 ~ ,
\end{eqnarray}
where it is seen that as $m^\prime$ increases from 0 (taking for example $\beta=\pi/4$), we approach the non-classical and violate \Eq{classiccondm0} at $m^\prime \ge \sin\beta\sqrt{l(l+1)}$; increasing $m^\prime$ further means sampling the divergent dominant solution of the Schr\"{o}dinger equation from which \Eq{mrec} is derived. It is then clear that the stable direction to use \Eq{mrec} is for decreasing $|m^\prime|$. From \Eq{classiccond}, it is seen quite generally that the recursion relations (\ref{mrec}) and (\ref{lrec}) will be stable provided they are used in the direction of decreasing $| m^\prime |$ and increasing $l$ respectively.
In addition to Eqs~(\ref{mrec}) and (\ref{lrec}), a third recursion relation in $\beta$ can be derived by discretizing the derivatives in \Eq{wigeq} with the relations
\begin{eqnarray}\label{firstder}
f^\prime(x) &\cong& \frac{f(x+\epsilon) - f(x-\epsilon)}{2\epsilon} + O(\epsilon^2 f^{\prime\prime\prime}) \\
f^{\prime\prime}(x) &\cong& \frac{f(x+\epsilon) + f(x-\epsilon) - 2f(x)}{\epsilon^2} + O(\epsilon f^{\prime\prime\prime})~.
\end{eqnarray}
Substituting into \Eq{wigeq} yields
\begin{eqnarray}\label{betarec}
& &\left[\epsilon^2\left( \frac{2mm^\prime \cos\beta - m^2 - m^{\prime 2}}{\sin^2\beta} + l(l+1)\right) - 2 \right] d^l_{mm^\prime}(\beta) \cong \nonumber \\
& & ~~ \left( \frac{\epsilon\text{cot}\beta}{2} - 1 \right) d^l_{mm^\prime}(\beta - \epsilon) - \left( \frac{\epsilon\text{cot}\beta}{2} + 1 \right) d^l_{mm^\prime}(\beta + \epsilon) + \\ \nonumber
& & ~~~~~~~~~~~~~~~~~+ O(\epsilon^3 d^{l^{\prime\prime\prime}}_{mm^\prime}(\beta))~.
\end{eqnarray}
From \Eq{classiccond}, it is seen that this recursion relation should be used for increasing $\beta$ if $0<\beta<\pi/2$ and decreasing $\beta$ if $\pi/2<\beta<\pi$. Examples of these conclusions are given in \Fig{dlmmplots}. The top plot shows the change in the behavior of $\dlmm$ with increasing $l$ as one moves from the non-classical to classical regions; in that case, the angle $\beta$ was chosen so that \Eq{classiccond} is satisfied only when $l\ge 100$. The middle plot shows the variation of $\dlmm$ with $m^\prime$. With $\beta=0.52331$, $l=1000$ and $m=0$, the transition from non-classical to classical regimes occurs at $m^\prime = 500$. The last plot shows the variation of $\dlmm$ with $\beta$; the value of $m^\prime=71$ was chosen so that the transition from non-classical to classical occurs at $\beta=\pi/4$. A noticeable feature of all three plots is the tallest peak is always the first peak after the transition to the classical region. This is qualitatively understandable from \Eq{reducedwignerdef} where the reduced Wigner matrix is seen to characterize the overlap between two spin states after a rotation. In the classical limit of large $l$, the angle $\omega$ of the spin direction of a quantum object with the $z$ axis is given by
\begin{eqnarray}\label{classicalAng}
| lm \rangle ~~ \text{:} ~~ \cos\omega \approx {m\over\sqrt{l(l+1)}}~.
\end{eqnarray}
One might expect that the overlap would be greatest when the 'classical' spins are aligned after the rotation about the $y$-axis. From Eqs.~(\ref{classiccond}) and (\ref{classicalAng}), we can show that this is the case when
\begin{eqnarray}\label{betaclassic}
\beta=\text{acos}\left(\frac{m^\prime}{\sqrt{l(l+1)}}\right)-\text{acos}\left(\frac{m}{\sqrt{l(l+1)}}\right)~.
\end{eqnarray}
In our example, $m=0$ and \Eq{betaclassic} reads $\sin(\beta)=m^\prime/[l(l+1)]$ and the overlap is greatest at the transition point.
\begin{figure}
\resizebox{8.7 cm}{!}{\includegraphics*{conviqtPaper_1col_300ppi_unscaled.png}}
\caption{The top plot shows the variation of $d^{l}_{010}(0.0996687)$ for $10\le l\le 500$; the middle plot shows the variation of $d^{500}_{0m^\prime}(0.52331)$ for $0\le m^\prime\le 500$; the third plot shows the variation of $d^{100}_{071}(\beta)$ for $0\le\beta\le \pi/2$.}\label{dlmmplots}
\end{figure}
\section{Algorithm}
The evaluation of the $\dlmm$ from Eqs.~(\ref{mrec}) and (\ref{lrec}) requires starting values for the recursions. For large $l$ however, those values are often vanishingly small and cannot
be represented by any of the IEEE 754 floating-point data formats which are used on practically all current computer hardware. Starting the recursions with 0 and 1 does not help because there will come a point where the $\dlmm$ become too big to be represented numerically. Two solutions to this problem are presented here.
\subsection{$\dlmm$ ratios}
From \Eq{wigeq} and the plots of \Fig{dlmmplots}, it is clear that the $\dlmm$ vary smoothly with varying $l$, $m$, and $\beta$. As a result, a recursion relation of ratios should always be finite in the non-classical region where the $\dlmm$ are not oscillatory, and one only has to worry about singularities in the ratios in the classical/oscillatory region where the denominators could vanish if evaluated at a zero of the $\dlmm$. In this ratio-based method, Eqs.~(\ref{mrec}) and (\ref{lrec}) can be rewritten
\begin{eqnarray}\label{ratiom}
& &\frac{d^l_{mm^\prime}}{d^l_{mm^\prime-1}} = \\ \nonumber
& & ~~\frac{ \sqrt{ (l+m^\prime)(l-m^\prime+1) } }{ m\,\text{cosec}\beta - m^\prime\text{cotan}\beta - \sqrt{ (l-m^\prime)(l+m^\prime+1)} \frac{ d^l_{mm^\prime+1} }{ d^l_{mm^\prime} } } \\ \label{ratiol}
& &\frac{d^l_{mm^\prime}}{d^{l+1}_{mm^\prime}} = \\ \nonumber
& & \frac{ (l\!+\!1) \sqrt{ (l^2\!-\!{m^\prime}^2)(l^2\!-\!m^2) } }{ (2l\!+\!1)( mm^\prime\! -\! \cos\!\beta ) - l \sqrt{ [(l\!+\!1)^2\!-\!{m^\prime}^2][(l\!+\!1)^2\!-\!m^2] } \frac{ d^{l-1}_{mm^\prime} }{ d^l_{mm^\prime} } } ~.
\end{eqnarray}
Using
\begin{eqnarray}
\frac{d^l_{ml}}{d^l_{ml-1}} &=& \frac{ \sqrt{l/2}\sin\beta }{ l\cos\beta - m}~~\text{and} \\
\frac{d^l_{ml}}{d^{l+1}_{ml}} &=& \sqrt{ \frac{ (l+m+1)(l-m+1) }{ 2l+1 } } [ (l+1)\cos\beta - m ]^{-1}~,
\end{eqnarray}
the ratios can be calculated recursively down to $m^\prime=0$ if using \Eq{ratiom} or up to a $l=\lmax$ if using \Eq{ratiol}. For example, in the case where all the $d^l_{0m^\prime}$ for $l\ge m^\prime \ge 0$ are required, one would start with \Eq{ratiom} to calculate:
\begin{eqnarray}\label{ratiosm0}
\frac{d^l_{0l}(\beta)}{d^l_{0l-1}(\beta)},~\frac{d^l_{0l-1}(\beta)}{d^l_{0l-2}(\beta)},~\dots,~\frac{d^l_{02}(\beta)}{d^l_{01}(\beta)},~\frac{d^l_{01}(\beta)}{d^l_{00}(\beta)}~.
\end{eqnarray}
To then calculate the $d^l_{0m^\prime}$, one would need to know $d^l_{00}$. Fortunately, the $d^l_{00}$ are easy to calculate because their recursion relation do not contain exponential solutions as remarked under \Eq{classiccond}:
\begin{eqnarray}\label{dl00}
\cos\beta d^l_{00}(\beta)=\frac{l}{2l+1}d^{l-1}_{00} + \frac{l+1}{2l+1}d^{l+1}_{00}
\end{eqnarray}
Once $d^l_{00}(\beta)$ has been calculated, $d^l_{01}$ can be calculated from the ratios; in order, each $d^l_{0m}(\beta)$ can be calculated by multiplying adjacent ratios until $d^l_{0l}(\beta)$ has been evaluated. In the case where $d^l_{1m^\prime}$ for $l\ge m^\prime \ge 0$ are also needed, the set of ratios
\begin{eqnarray}\label{ratiosm1}
\frac{d^l_{1l}(\beta)}{d^l_{1l-1}(\beta)},~\frac{d^l_{1l-1}(\beta)}{d^l_{1l-2}(\beta)},~\dots,~\frac{d^l_{12}(\beta)}{d^l_{11}(\beta)},~\frac{d^l_{11}(\beta)}{d^l_{10}(\beta)}~.
\end{eqnarray}
are next computed. To then calculate the $d^l_{1m^\prime}$, one needs to know $d^l_{10}$. Fortunately, $d^l_{01}=-d^l_{10}$ has previously been calculated and all the $d^l_{1m^\prime}$ can be obtained up to $d^l_{1l}$. In this fashion, all the $\dlmm$ can be calculated up to a desired $m=m_{\text{max}}$.
The special and extremely rare case where a ratio $d^l_{mm^\prime+1}/d^l_{mm^\prime}$ is infinite can be handled by using \Eq{mrec} where $\dlmm$ is set to zero and substituting an infinite ratio and a null ratio for a single finite ratio:
\begin{eqnarray}
\frac{d^l_{mm^\prime+1}}{d^l_{mm^\prime}},~\frac{d^l_{mm^\prime}}{d^l_{mm^\prime-1}} \to \frac{d^l_{mm^\prime+1}}{d^l_{mm^\prime-1}} = -\frac{\sqrt{(l+m^\prime)(l-m^\prime+1)}}{\sqrt{(l-m^\prime)(l+m^\prime-1)}}
\end{eqnarray}
Note that in contrast to the method described in \cite{risbo}, the column of matrix elements $d^l_{ml}(\beta)\to d^l_{m0}(\beta)$ can be evaluated without having to calculate every single $d^{l^\prime}_{mm^\prime}(\beta)$ for $l^\prime < l$.
The same tricks can be applied to the recursion relation in $l$ for the evaluation of $d^l_{lm^\prime}(\beta)\to d^{\lmax}_{lm^\prime}(\beta)$. First calculate the column of elements $d^{\lmax}_{l\lmax}(\beta)\to d^{\lmax}_{l0}(\beta)$ and then calculate the ratios
\begin{eqnarray}\label{ratiosl0}
\frac{d^l_{lm^\prime}(\beta)}{d^{l+1}_{lm^\prime}(\beta)},~\frac{d^{l+1}_{lm^\prime}(\beta)}{d^{l+2}_{lm^\prime}(\beta)},~\dots,~\frac{d^{\lmax-2}_{lm^\prime}(\beta)}{d^{\lmax-1}_{lm^\prime}(\beta)},~\frac{d^{\lmax-1}_{lm^\prime}(\beta)}{d^{\lmax}_{lm^\prime}(\beta)}~.
\end{eqnarray}
Knowing $d^{\lmax}_{lm^\prime}(\beta)$ allows us to evaluate $d^{\lmax-1}_{lm^\prime}(\beta)\to d^l_{lm^\prime}(\beta)$.
For a particular $\beta$, it is sufficient to compute the elements of $\dlmm$ for $0 \le m \le l$ and $-l \le m^\prime \le l$ to know the entire matrix $d^l(\beta)$. To fill out the rest of the matrix, the symmetry relations in appendix~\ref{appa} can be used.
\subsection {Wigner matrix elements by $l$-recursion}\label{wigl}
Another way to deal with the underflow problem is to start from \Eq{lrec} with the following initialization values
\begin{align}
&& &d^l_{l,m}(\beta)& &=& &A \, \left( \cos\frac{\beta}{2}\right)^{l+m} \, \left(-\sin\frac{\beta}{2}\right)^{l-m}& &&\label{ini1}\\
&& &d^l_{-l,m}(\beta)& &=& &A \, \left( \cos\frac{\beta}{2}\right)^{l-m} \, \left(\sin\frac{\beta}{2}\right)^{l+m}& &&\\
&& &d^l_{-l,m}(\beta)& &=& &A \, \left( \cos\frac{\beta}{2}\right)^{l+m} \, \left(\sin\frac{\beta}{2}\right)^{l-m}& &&\\
&& &d^l_{-l,m}(\beta)& &=& &A \, \left( \cos\frac{\beta}{2}\right)^{l-m} \, \left(-\sin\frac{\beta}{2}\right)^{l+m}\text{,}& &&\label{ini2}
\end{align}
where $A=\sqrt{(2l)!/(l+m)!(l-m)!}~$. As far as equations (\ref{ini1}) to (\ref{ini2}) are concerned, the underflow problem
can be avoided by simply calculating the logarithm of the absolute value
of the matrix element and storing its sign separately.
Equation (\ref{ini1}), e.g., then transforms to
\begin{eqnarray}
\ln \left|d^l_{l,m}(\beta)\right|&=&0.5\left(\ln((2l)!)-\ln((l+m)!)-\ln((l-m)!\right)\nonumber \\
&+& (l+m)\ln\left|\cos(\beta/2)\right| + (l-m)\ln\left|\sin(\beta/2)\right| \label{logstart}
\end{eqnarray}
In cases where one of the last two terms is $-\infty$, the recursion in $l$ can be stopped
immediately, since all subsequent values will be zero.
The logarithms of the faculties are easily precomputed, so that the seed
value for the recursion can be obtained in $\mathcal O(1)$ operations.
Since the result of eq.\ (\ref{logstart}) is in some circumstances much smaller
than the individual terms on the right-hand side, cancellation errors may
reduce the number of significant digits of the result.
In order to have the highest accuracy that can be achieved without sacrificing
too much performance, the computation of the seed value is carried
out with extended IEEE precision (corresponding to the C++ data type
{\tt long double}).
The recursion relation (\ref{lrec}) itself unfortunately cannot be computed
conveniently in logarithms; therefore a way must be found to represent floating
point values with an extreme dynamic range, which does not incur a high
performance penalty.
This was implemented by representing a floating-point number $v$ using an
IEEE double precision value $d$ and an integer scale $n$, such that
\begin{equation}
v=d\cdot S^n\text{,}
\end{equation}
where either $d=0$, or $S^{-1}\le|d|\le S$ and $S$ (the ``scale factor'') is a
positive constant that can be represented as a double-precision IEEE value.
Using this prescription, $v$ does not have a unique representation as a
$(d,n)$-pair, but this is not a problem.
Similar techniques have been in use since at least three decades in numerical
algorithms; for a recent example see the spherical
harmonic transform routines of the HEALPix package.
It is advantageous to choose a scale factor which is an integer power of 2,
because multiplying or dividing by such a factor only affects the exponent of a
floating-point value stored in binary format, and is therefore exact
(ignoring possible under- or overflows).
In order to avoid frequent re-scaling of $d$, the scale factor should also be
rather large; the value adopted for our implementation is $2^{90}$.
Using this representation for the $d^l_{mm'}(\beta)$, the recursion is
performed until either $\lmax$ is reached, or the matrix element has
become large enough to be safely represented by a normal double-precision
variable (the threshold value used in the code is $2^{-900}$).
In the latter case the remaining computations up to $\lmax$
are done with standard floating-point arithmetic, which is signi\-fi\-cant\-ly faster.
\section{Convolution}
One area where fast and efficient techniques of computing $\dlmm$ are particularly valuable is in $4\pi$ convolution~\cite{Wandelt:2001gp}. For the convolution of two fields $b(\Omega)$ and $s(\Omega)$ defined on a sphere, the following integral must be calculated:
\begin{eqnarray}
c = \int\text{d}\!\Omega b^*\!(\Omega)s(\Omega)
\end{eqnarray}
In the physical application where $b(\Omega)$ is a beam from a horn located on a slowly rotating space telescope that scans the sky (denoted $s(\Omega)$) as it orbits the sun (WMAP or Planck missions, e.g.), a large number of convolutions must be performed to account for every possible orientation $(\alpha, \beta, \gamma)$ of the satellite
\begin{eqnarray}\label{conv}
c(\Omega^\prime) = \int\text{d}\!\Omega [\text{R}(\Omega^\prime)b(\Omega) ]^*s(\Omega)
\end{eqnarray}
where $\text{R}(\Omega^\prime)$ is a rotation matrix that rotates the beam to a particular orientation of the satellite and is defined
\begin{eqnarray}\label{rot}
\text{R}(\alpha, \beta, \gamma) \text{Y}_{lm}(\theta,\phi) = \sum_{m^\prime=-l}^{l} D^l_{m^\prime m}(\alpha, \beta, \gamma) \text{Y}_{lm^\prime}(\theta,\phi) ~,
\end{eqnarray}
and where the $\text{Y}_{lm}(\theta,\phi)$ are spherical harmonics. The $ \text{Y}_{lm}(\theta,\phi)$ are related to the $\Dlmm$ through the relation
\begin{eqnarray}
\text{Y}_{lm}(\theta,\phi) = (-1)^m \sqrt{ \frac{ 2l+1 }{ 4\pi } } D^l_{0m}(0,\theta,\phi)
\end{eqnarray}
and can therefore be calculated using the methods described above.
For beams with significant side-lobes stemming from reflections of light far away from the line of sight as is the case for both the WMAP and Planck missions, the beams can cover a significant portion of the sky and full-sky convolutions are necessary; as shown in ref~\cite{Wandelt:2001gp}, such full-sky convolutions are much faster when performed in harmonic space instead of pixel space. We now describe a very fast and massively parallel method to perform full-sky convolutions in harmonic space.
\subsection{\tt CONVIQT}
{\tt conviqt} (CONvolution VIa the Quantum Top equation) is a fast 4$\pi$ convolution algorithm that relies on fast computational methods for reduced Wigner matrix elements. Starting from Eqs.~(\ref{conv}) and (\ref{rot}), the beam and sky fields can be expanded on the spherical harmonic basis to yield
\begin{eqnarray}\label{conviqt}
c (\alpha, \beta, \gamma) &=& \sum_{\mb=-\mmax }^{\mmax}\sum_{\msky=-\lmax }^{\lmax} e^{-i\mb\alpha}e^{-i\msky \gamma} C_{\mb\msky}(\beta) ~, \\ \label{convdef}
C_{\mb\msky}(\beta) &\equiv& \sum_{l=0}^{\lmax} {b^*_{l\mb}} \dconv s_{l\msky}~,
\end{eqnarray}
where $b_{l\mb}$ and $s_{l\msky}$ are the spherical components of the beam and sky fields.\footnote{For ease of reading, only the scalar case is described since the generalization to polarized maps and beams is easily accomplished by evaluating \Eq{convdef} for the additional pairs $(b_{l\mb}^G,~s_{l\msky}^G)$ and $(b_{l\mb}^C,~s_{l\msky}^C)$.} Typically, the $b_{l\mb}$ are negligible for some $\mmax<\mb\ll\lmax$. Noting that the number of $\beta$ angles needed for the convolution scales as $\lmax$, the evaluation of \Eq{conviqt} scales as $O(\lmax^2\mmax\log(\lmax))$ after the use of the Fast Fourier Transform algorithm to perform the summations. The numerically expensive part of \Eq{conviqt} is the computation of \Eq{convdef} which scales as $O(\lmax^3\mmax)$. Two separate computations of \Eq{convdef} scale as $O(\lmax^3\mmax)$: the computation of the $\dconv$, and the evaluation of the sum over $0\le l \le \lmax$ for every single $m$, $\msky$, and $\beta$. The fast methods described in the previous section are used to compute the $\dconv$. To evaluate the sum for each $\beta$, a massively parallel MPI-based approach is used since the $C_{\mb\msky}(\beta)$ are uncorrelated between the different $\beta$ and can be computed by different tasks. Additional acceleration techniques for both the computation of the $\dconv$ and for the evaluation of the sums over $l$ are described in the following sub-section.
\subsection{Acceleration techniques for the $\dconv$}
In simulations of the measurement of cosmic microwave background, convolutions appear repeatedly especially if Monte Carlos are required.
Since the generation of Wigner matrix elements is typically the most computationally
intensive part of the convolution algorithm, a large effort was made to
increase its efficiency. This has two aspects: first to compute the matrix
elements as quickly as possible, but also to decide (if possible) which matrix
elements are too small to contribute measurably to the result and skip their
generation altogether.
\subsubsection{Skipping unneeded calculations}
When performing convolutions, especially within the context of Monte Carlo simulations where many convolutions with the same $\lmax$ and $\mmax$ are needed, it is computationally profitable to skip unneeded calculations.\footnote{Note that it is generally not more efficient to evaluate all of the $\dconv$ before hand because of the disk space required and the large amount of time needed to read them in; it is more efficient to calculate them on the fly.} Some terms in the sum of \Eq{convdef} need not be included because their $\dconv$ are vanishingly small. To determine which terms to exclude we turn to \Eq{classiccond} where three general possibilities are considered:
\begin{itemize}
\item $\mb$ and $\msky$ are of similar magnitude and much smaller than $l$.
\item $\mb$ and $\msky$ are of similar magnitude and of the same order of magnitude as $l$.
\item $\mb$ and $\msky$ are of widely differing magnitude with one much smaller than $l$ and one of similar magnitude.
\end{itemize}
In each of these possibilities, the neglected $\dconv$ are those evaluated at $\beta$ angles that correspond to the non-classical, exponentially suppressed region. Before explaining how these $\dconv$ are identified, {\tt conviqt}'s nested structure should be described. In {\tt conviqt}, the outermost loop deals with $\mb$ (which ranges from 0
to $\mmax$; nested into the $\mb$-loop
is the $\msky$-loop ranging from $-\msky$ to $\msky$; nested in the $\msky$-loop is the loop over the $\beta$ processed by that particular task.\footnote{A note to remind the reader that {\tt conviqt} is parallelized in $\beta$; each task will perform the convolutions in a subset of all the complete set of $\beta$ where $C_{\mb\msky}(\beta)$ must be calculated.} Finally, the innermost loop is that over $l$ which is where the condition is applied. To derive the minimum $l$ such that outside the parameter space defined by $l_{\text{min}}\le l \le \lmax$, $\dconv$ is negligible, we use \Eq{classiccond} to write
\begin{eqnarray} \nonumber
l_{\text{min}} &=& -\frac{1}{2} + \frac{1}{2}\left[ 1 + \frac{4}{\sin^2\beta} ( \msky^2 + \mb^2 - 2 \msky\mb\cos\beta ) \right]^{1/2} \\ \label{lmin}
&\cong& \frac{\sqrt{\msky^2 + \mb^2 - 2 \msky\mb\cos\beta}- \text{\tt offset}}{\sin\beta}
\end{eqnarray}
where {\tt offset}$>$0 and ensures that the $\dconv$ neglected are well within the non-classical region and suppressed. Calling $m_{\text{big}}$ the larger of $|\mb|$ or $|\msky|$, it is noted that $l \ge m_{\text{big}}$. To determine the {\tt offset}, we go back to the three possibilities listed above; of those, $l_{\text{min}} $ will generally equal $m_{\text{big}} $ in the cases where $\mb$ and $\msky$ are of similar magnitude; only in the third case will we generally have $l_{\text{min}} > m_{\text{big}}$, namely when $\mb$ and $\msky$ are of widely differing magnitude. From ref.~\cite{rowe}, this case can approximately be written as a harmonic oscillator wave function:
\begin{eqnarray}\label{asymp}
\dconv &\to& (-1)^{l-\msky}(\sqrt{l}\sin\beta_m)^{-1/2} u_{l-\msky}(\sqrt{l}(\beta-\beta_m) \\
u_\nu(x) &=& (\sqrt{\pi}2^\nu\nu!)^{-1/2}H_\nu(x)e^{-x^2/2}
\end{eqnarray}
where $H_\nu(x)$ is a Hermite polynomial and $\cos\beta_m=\mb/l$. Since we are dealing with orders of magnitude, it is not necessary to evaluate \Eq{asymp} exactly to determine {\tt offset}, only to calculate an estimate from the factor $\exp(-x^2/2)$. We have found that using {\tt offset}$\ge\lmax/20$ gives extremely accurate results. Finally it is noted that $d^l_{\mb\msky}(0) = \delta_{\mb\msky}$ and for that special case we put $l_{\text{min}}>\lmax$ when $\mb\ne\msky$ and no sum over $l$ is performed.
Thus, \Eq{lmin} is used to estimate whether the absolute values of all
$\dconv$ for a given combination of $l$, $\mb$, $\msky$ and $\beta$
lie below a certain threshold; if this is the case, the generation of these
values can be skipped entirely. In particular, the most efficient version of the code written was one where the $d^{l_{\text{min}}}_{\mb\msky}$ and $d^{l_{\text{min}}+1}_{\mb\msky}$ were pre-calculated for $0\le \mb \le \mmax$, $-\lmax \le \msky \le \lmax$ and $\beta$ subset for a particular task, and read-in as seeds to the recursion relation of \Eq{lrec}. That way, none of the unneeded $\dconv$ were calculated during the convolution. This required an extra code to pre-compute the $\dconv$. In the end, we opted for a single code based on the evaluation of the reduced Wigner matrix elements as described in section~\ref{wigl} because of the low overhead and maintenance as well as the high efficiency.
In this approach, we calculate all the $\dconv$ on the fly, but only include the relevant $\dconv$ in the final $l$-loop. This code is self-contained and easier to maintain at a very minimal cost in performance. As the $\dconv$ recursion is performed, the code checks the absolute values
of the generated $d^l_{mm'}(\beta)$ and records the $l$ index at which
a predefined threshold $\varepsilon$ (typically set to $10^{-30}$) is crossed
for the first time.
Due to the limited dynamic range of IEEE data types, values below
this threshold have no measurable influence on the convolution result and
can therefore be neglected during the final summation loop, which saves
a significant amount of CPU time.
\subsubsection{Precomputed values}
In this single code approach, a precomputation strategy well-suited to the loop structure was adopted:
\begin{itemize}
\item At the beginning of each run, we compute just once $\ln(\cos(\beta/2))$,
$\ln(\sin(\beta/2))$ and $\cos\beta$ for all $\beta$ at
which we need the Wigner matrix, and as mentioned above, $\ln n!$ up to
$n=2\lmax$.
\item Also at the beginning we compute the tables
\begin{equation*}
P_i=\sqrt{1/(i+1)}\quad \text{and}\quad Q_i=\sqrt{i/(i+1)}
\end{equation*}
for $i$ in the range of 0 to $2\lmax+1$, which
are needed for the next precomputation step.
\item inside the second loop (i.e.\ for every combination of $\mb$ and $\msky$)
we compute the tables
\begin{eqnarray*}
F_{0,l}&=&(l+1)(2l+1)P_{l+\mb}P_{l-\mb}P_{l+\msky}P_{l-\msky}\text{,}\\
F_{1,l}&=&\mb\msky/(l(l+1))\text{, and}\\
F_{2,l}&=&Q_{l+\mb}Q_{l-\mb}Q_{l+\msky}Q_{l-\msky}(l+1)/l\text{.}
\end{eqnarray*}
\end{itemize}
After all these preparations, eq.\ (\ref{lrec}) boils down to
\begin{equation}
d^{l+1}_{\mb\msky}(\beta) = F_{0,l}\left(\cos\beta-F_{1,l}\right)
d^l_{\mb\msky}(\beta) - F_{2,l}d^{l-1}_{m\msky}(\beta)\text{,}
\end{equation}
which corresponds to only five quick-to-compute floating-point operations.
The space overhead for the additional tables is $\mathcal{O}(\lmax)$,
which is insignificant compared to the $\mathcal{O}(l^2_{\text{max}})$ memory
requirement of the whole
convolution code. This also means that for the reasonable assumption of
$\lmax\lesssim 10^4$ all data required for the recursion
fit conveniently into current processors' Level-2 caches.
\subsubsection{Use of $\dlmm$ symmetries}
The use of the $\dlmm$ symmetries considerably cut the computational cost of the full sky convolution. In particular, \Eq{pithetasym} relates the computed values of $\Cmm$ at $\beta < \pi/2$ to those at $\beta > \pi/2$. In \Eq{pithetasym}, the phase factor $(-1)^{l}$ is accounted for by splitting the sum in \Eq{convdef} into even and odd $l$. The phase factor $(-1)^{-m^\prime}$ is accounted for by using the relations
\begin{eqnarray}\label{blmslmsym}
b_{lm}=(-1)^m b_{l-m}^*~, ~~s_{lm}=(-1)^m s_{l-m}^*
\end{eqnarray}
and further splitting the odd and even sums of \Eq{convdef} into real and imaginary parts. In addition, the symmetry of \Eq{mmsym} and \Eq{blmslmsym} can be used to show
\begin{eqnarray}
C_{-\mb-\msky}(\beta) = \Cmm^*
\end{eqnarray}
speeding up the computation of $\Cmm$ by another factor of two.
\subsection{Example simulations}
To determine the accuracy of {\tt conviqt}, a detailed comparison with the stable release of the {\tt LevelS totalconvolver} \cite{Wandelt:2001gp, Reinecke:2006fv} currently compiled on the {\tt planck} cluster at the National Energy Research Science Council (NERSC) was performed. {\tt LevelS} is a simulation package for the generation of time ordered data (TOD) by the Planck satellite~\cite{planck}. {\tt Totalconvolver} and {\tt conviqt} both calculate a data cube that is fully compatible with LevelS. For both codes, data cube is composed of convolved points calculated at a polar angle $\theta$, a longitudinal angle $\phi$ and a particular beam orientation (a rotation about the beam axis) $\psi$.
\subsubsection{data cube comparison}
For $\lmax=2000$, GRASP beams for LFI-19a, $\mmax = 9$, offset = 30 and a polarized CMB map, there were 153634399 points in the data cube. Taking the difference between the totalconvolver and conviqt data cubes (residual values below), we had:
\begin{tabular}{|c|c|c|c|c|}
\hline
{\tt offset} & avr &$\sigma$ & Max & rel \\ \hline
2000 (exact) & -1.7e-16 & 4.1e-13 & 1.4e-11 & 4.0e-8 \\ \hline
30 & -1.2e-16 & 1.3e-11 & 5.6e-10 & 1.3e-6 \\ \hline
15 & -1.6e-16 & 3.0e-9 & 1.1e-7 & 2.9e-4\\
\hline
\end{tabular}
where 'avr' refers to the average difference of the two data cubes ({\tt conviqt}($\theta,\phi,\psi$)-{\tt totalconvolver}($\theta,\phi,\psi$)), $\sigma$ is the variance of that residual data cube, 'Max' refers to the maximum value found in the residual data cube, and 'rel' refers to the ratio of $\sigma$ to the variance of the {\tt totalconvolver} data cube. We see that with {\tt offset}=2000, the two data cubes agree to approximately 8 significant digits; for {\tt offset}=30,15 they agree to 6 and 4 significant digits respectively.
\subsection{Performance}
\begin{figure}
\resizebox{8.7 cm}{!}{\includegraphics*{secGbytesPlot-scaled.png}}
\caption{The top plot compares the wall clock performance of conviqt and totalconvolver at $\lmax$ intervals of 256 starting at $\lmax=256$; the middle plot compares the memory needs in GBytes of the two codes. The lower plot shows the ratio of wall clock and Gbytes of the two codes.} \label{secgb}
\end{figure}
On a single processor on Jacquard, for lmax=2000 with a GRASP beam (LFI-19a) of $\mmax=9$, offset=30, MC=T including polarisation, conviqt had the following performance
\begin{tabular}{|c|c|c|c|c|}
\hline
{\tt Code} & Clock (secs) & gbytes \\ \hline
{\tt conviqt} & 349 & 0.37 \\ \hline
{\tt totalconvolver} & 1120 & 0.41 \\ \hline
\end{tabular}\\
where 'Clock' refers to the time the it took to complete the convolution according to the wall clock, while 'gbytes' refers to the total memory consumed. These numbers were obtained using NERSC's Integrated Performance Monitoring (IPM) on a 712-CPU Opteron cluster called {\tt Jacquard} running a Linux operating system; each processor on {\tt Jacquard} runs at a clock speed of 2.2GHz with a theoretical peak performance of 4.4GFlop/s. Unlike {\tt totalconvolver}, {\tt conviqt} is a massively parallel code which can be run on machines with distributed memory; running it on a single processor shows that {\tt conviqt} is intrinsically faster and more efficient than {\tt totalconvolver}. The above table is for the case where {\tt offset}~$<$~$\lmax$, i.e., the case where {\tt conviqt} sacrifices precision for the sake of a speedier convolution.
For the general case where {\tt conviqt} and {\tt totalconvolver} calculate the same thing, we use the single code approach to obtain the following table:
\begin{tabular}{|c|c|c|c|c|}
\hline
& \multicolumn{2}{|c|}{\tt conviqt} & \multicolumn{2}{|c|}{\tt totalconvolver} \\ \hline
$\lmax$ & secs & GBytes & secs & GBytes \\ \hline
256 & 1.9657 & 1.51806e-02 & 11.726 & 2.32534e-02 \\ \hline
512 & 9.8691 & 1.76401e-02 & 54.634 & 5.00412e-02 \\ \hline
768 & 27.982 & 2.16856e-02 & 144.51 & 9.46579e-02 \\ \hline
1024 & 61.797 & 2.73190e-02 & 294.71 & 1.57092e-01 \\ \hline
1280 & 117.13 & 3.45383e-02 & 513.29 & 2.37350e-01 \\ \hline
1536 & 199.29 & 4.33445e-02 & 809.51 & 3.35554e-01 \\ \hline
1792 & 313.27 & 5.37376e-02 & 1264.1 & 4.51582e-01 \\ \hline
2048 & 621.26 & 6.57196e-02 & 1765.2 & 5.85376e-01 \\ \hline
\end{tabular}
These numbers are plotted in \Fig{secgb}. The top plot shows that {\tt conviqt} is considerably faster than {\tt totalconvolver} for $\lmax<2048$; however, because both codes scale as $\lmax^3$ as $\lmax\to\infty$, the gap between their total wall clock times will narrow. It is also seen that {\tt conviqt} consumes significantly less memory.
The scaling of {\tt conviqt} timings as a function of the total number of processors is very good. For a $\lmax=4096$ and $\mb=14$ with polarized beam and sky, the log-log plot in \Fig{scaleNode} shows a linear relationship up to a convolution distributed on 128 processors.\\
\begin{tabular}{|c|c|}
\hline
Number of processors & seconds \\ \hline
8 & 965.46 \\ \hline
16 & 486.29 \\ \hline
32 & 247.83 \\ \hline
64 & 128.43 \\ \hline
128 & 69.06 \\ \hline
192 & 52.47 \\ \hline
\end{tabular}\\
This plot was obtained using the single code approach run on the NERSC cluster called {\tt Planck}, a 256 cores cluster of Opteron 2350 2.0GHz processors. To measure the scaling behavior of {\tt conviqt}, no output file was created to avoid skewing the scaling law with the time it takes to write the file (tens of seconds for a 4GB file). As the number of processors increases and the time required to perform the convolution diminishes to less than a minute, the timings become dominated with operations that have nothing to do with the convolution; among these are the reading of the input data sets (which are read in full
by all MPI tasks), the inter-process communication and various calculations which are performed
redundantly on all tasks, because communicating the results would be more expensive. Increasing the number of tasks (while keeping the problem size constant) also means a smaller number of $\beta$ angles per task,
which decreases the achievable quality of load balancing. In addition, different runs with identical inputs show variations of a few seconds in wall clock timings that have an increasing relative impact on the decreasing timings stemming from using larger numbers of processors; the most likely explanation for this are differences in the exact nature of process startup and disk access, which is not exactly reproducible in this kind of computing environment.
\begin{figure}
\resizebox{8.7 cm}{!}{\includegraphics*{scaleNodes.png}}
\caption{Timings of conviqt as the number of processors is increased with $\lmax=4096$ and $\mmax=14$.}\label{scaleNode}
\end{figure}
\section{Summary}
New algorithms for the efficient and accurate calculation of Wigner matrix elements were presented. These algorithms were used in a full sky convolution, massively parallel algorithm called {\tt conviqt} that was shown to be significantly more efficient and much faster than the only other algorithm currently available.\\
\\
{\it \bf Acknowledgements}: GP would like to thank Maura Sandri for the GRASP 8 beams used in these simulations and Charles Lawrence for useful comments on this manuscript. We gratefully acknowledge support by the NASA Science Mission Directorate via the US Planck Project. The research described in this paper was partially carried out at the Jet propulsion Laboratory, California Institute of Technology, under a contract with NASA.
This research used resources of the National Energy Research Scientific
Computing Center, which is supported by the Office of Science of the
U.S. Department of Energy under Contract No. DE-AC02-05CH11231.
MR is supported by the
German Aeronautics Center and Space Agency (DLR), under program 50-OP-0901, funded by the
Federal Ministry of Economics and Technology. Copyright 2010. All rights reserved.
\begin{appendix}
\section{}\label{appa}
\begin{eqnarray}\label{mmsym}
\dlmm &=& (-1)^{m-m^\prime}d^l_{-m-m^\prime}(\beta) \\
\dlmm &=& (-1)^{m-m^\prime}d^l_{m^\prime m}(\beta) \\
\dlmm &=& d^l_{-m^\prime -m}(\beta) \\
d^l_{mm^\prime}(-\beta) &=& d^l_{m^\prime m}(\beta) \\
d^l_{mm^\prime}(-\beta) &=& (-1)^{m-m^\prime}d^l_{mm^\prime }(\beta) \\ \label{pithetasym}
d^l_{mm^\prime}(\pi-\beta) &=& (-1)^{l-m^\prime}d^l_{-mm^\prime }(\beta) \\
d^l_{mm^\prime}(\pi-\beta) &=& (-1)^{l+m^\prime}d^l_{m-m^\prime }(\beta)
\end{eqnarray}
\end{appendix}
|
\section{Introduction}
The magnetohydrodynamics (MHD) equations describe
macroscopic dynamics of electrically conducting fluid that moves in a magnetic field. MHD model is governed by Navier-Stokes equations coupled with Maxwell equations through Ohm's law and Lorentz force. As an example, a resistive MHD system is described by the following equations:
\begin{equation*}
\left\{
\begin{array}{rcl}
\rho(\mathbf{u}_t+\mathbf{u}\cdot \nabla \mathbf{u})+\nabla p &=& \frac{1}{\mu_0}(\nabla\times \mathbf{B})\times \mathbf{B} + \mu\Delta \mathbf{u},\\
\nabla\cdot \mathbf{u} & = & 0,\\
\mathbf{B}_t - \nabla\times(\mathbf{u}\times \mathbf{B}) & = & -\frac{\eta}{\mu_0}(\nabla\times)^2 \mathbf{B} - \frac{d_i}{\mu_0}\nabla\times((\nabla\times \mathbf{B})\times\mathbf{B}) \\
&& - \frac{\eta_2}{\mu_0}(\nabla\times)^4 \mathbf{B}, \\
\nabla\cdot \mathbf{B} & = & 0,
\end{array}
\right.
\end{equation*}
where $\rho$ is the mass density, $\mathbf{u}$ is the velocity, $p$ is the pressure, $\mathbf{B}$ is the magnetic induction field, $\eta$ is the resistivity, $\eta_2$ is the hyper-resistivity, $\mu_0$ is the magnetic permeability of free space, and $\mu$ is the viscosity. The primary variables in MHD equations are fluid velocity $\mathbf{u}$ and magnetic field $\mathbf{B}$.
MHD model has widespread applications in thermonuclear fusion, magnetospheric and solar physics, plasma physics, geophysics, and astrophysics. Mathematical modeling and numerical simulations of MHD have attracted much research effort in the past few decades. Various numerical algorithms have been used in MHD simulations; examples include finite difference methods, finite volume methods, finite element methods, and Fourier-based spectral and pseudo-spectral methods \cite{Toth:1996gd}. In \cite{Jardin:2004qo,Jardin:2005dq, Kang:2008ta,Krzeminski:2000la,Ovtchinnikov:2007kb}, two-dimensional, incompressible MHD problems are studied in terms of finite element approximations of the stream function-vorticity advection formulation. Since MHD flow often develop sharp interfaces, adaptive $h$-refinement techniques have been applied in MHD simulations \cite{Lankalapallia:2007ye, Strauss:1998kc, Ziegler:2003gf}. Finite element computations of MHD problems in three-dimensions have been reported in \cite{Codina:2006zv, Gerbeau:2000lo, Layton:1997hs,Salah:2001fh,Schotzau:2004sw, Wiedmer:1999pr}.
In the existing finite element discreitzations for the above MHD model, a standard pair of stable or stabilized finite element spaces are often used to discretize the velocity and pressure variables in the fluid equations. For the magnetic field variable $\mathbf{B}$, however, at least two approaches are possible when the fourth order term $(\nabla\times)^4\mathbf{B}$ is not presented in the model, namely when electron viscosity $\eta_2=0$. One approach is to use the standard edge element (\cite{Schotzau:2004sw}) and the other approach is to use the Lagrange element after replacing $(\nabla\times)^2\mathbf{B}$ by $-\Delta \mathbf{B}$ (\cite{Gerbeau:2000lo, Salah:2001fh,Wiedmer:1999pr}). Both these approaches will become more difficult when the fourth order term $(\nabla\times)^4\mathbf{B}$ is presented. We may still replace $(\nabla\times)^4 \mathbf{B}$ by a biharmonic operator $\Delta^2 \mathbf{B}$. But discretizing a biharmonic operator in three dimensions is challenging. It requires $220$ degrees of freedom per element if a conforming finite element is used. One possible way to reduce the number of degrees of freedom is to use nonconforming discretizations which allow weaker inter-element smoothness constraints but still provid convergent approximations. Among the class of nonconforming finite elements for fourth order problems, Morley-type elements are special in the sense that they provide approximations with polynomials of minimal degree \cite{ Morley:1968ve, Wang:2006qr}. In \cite{Wang:2006bh}, a systematic construction of Morley-type elements is provided for solving $2m$-th order partial differential equations in $\mathbb{R}^n$.
In particular, we may apply the element in \cite{Wang:2006bh} with $n=3$ and $m=2$ consisting of piecewise quadratic elements to our system of biharmonic equations. This amounts to $30$ degrees of freedom on each element. This element provides a reasonable discretization of MHD equations when it
is appropriate to replace $(\text{curl})^4$ operator by the biharmonic operator. This approach, however, may lead to difficulty for certain boundary conditions in practical applications. Indeed the treatment of boundary conditions is also an issue for the second order problem if $(\nabla\times)^2\mathbf{B}$ is replaced by $-\Delta\mathbf{B}$ \cite{Guermond:2003fk}.
One more natural approach is to discretize the fourth order curl operator by some generalized higher order edge elements. But such type of edge elements are not available in the literature. The construction of such type of edge element is the main goal of this paper.
Another possible approach to deal with the fourth order term $(\nabla\times)^4\mathbf{B}$ is to use operator splitting technique. Namely, one can introduce an intermediate variable $\sigma = (\nabla\times)^2\mathbf{B}$ and then reduce the original problem to a system of second order equations. However, it is known that for some problems, such a technique cannot be applied. For example, when modeling the bending of simply supported plate on non-convex polygonal domains, the original biharmonic problem is not equivalent to the lower order system of two Poisson
equations \cite{Blum:1980ly, Rannacher:1979zr}. In view of this, we consider discretizing the fourth order problem directly.
In this paper, we investigate MHD equations that contain both fourth-order term and second-order term. In the literature, the major tool used for performing MHD simulations involving a fourth order equation has been the pseudo-spectral method \cite{Biskamp:1995it}. By choosing an appropriate formulation, we are able to construct a finite element approximation for this problem. This is a nonconforming finite element
that involves only a small number of degrees of freedom.
The rest of this paper is organized as follows. In Section 2, we
describe a simplified model problem and the corresponding variational formulation. In Section 3, we construct basis functions and provide the convergence analysis. Finally, in Section 4, we give some concluding remarks.
\section{Model Problem}
In the following, we introduce model problems for the fourth-order magnetic induction equations described above. Assume that $\Omega\subset \mathbb{R}^3$ is a bounded polyhedron. By considering a semi-discretization in time and then ignoring the nonlinear terms, we obtain the following equations:
\begin{equation}
\left\{
\begin{aligned}
\alpha(\nabla\times)^4 \mathbf{u} + \beta(\nabla\times)^2 \mathbf{u} +\gamma \mathbf{u}&=\mathbf{f},\;\text{in}\;\Omega,\\
\nabla\cdot\mathbf{u} & = 0,\;\text{in}\;\Omega,
\end{aligned}
\right.\label{four_curl_bvp}
\end{equation}
where $\nabla\cdot\mathbf{f}= 0$, and the parameters $\alpha, \beta, \gamma > 0 $. We consider homogeneous boundary conditions,
\begin{equation}
\mathbf{u}\times \mathbf{n}= 0,\;\nabla\times \mathbf{u}= 0,\;\text{on}\;\partial\Omega.
\end{equation}
The above choice of boundary conditions arise naturally in the variational formulation given below. On the other hand, in the numerical simulations of the problem with pseudo-spectral method, one often uses periodic boundary conditions, e.g., \cite{Biskamp:1995it,Germaschewski:1999yq}.
It is worth pointing out that the parameter $\alpha$ is usually much smaller than either $\beta$ or $\gamma$. This fact imposes some difficulties in designing robust numerical methods, as have been studied in the context of biharmonic problems, e.g., \cite{Nilssen:2001zr,Wang:2006ly}.
The above fourth-order curl equations also arise from an interior transmission problem in the study of inverse scattering problems for inhomogeneous medium, e.g., \cite{Cakoni:2007yq}.
In order to provide an appropriate framework for our analysis, we define the following function spaces:
$$
H(\text{curl};\Omega)=\{\mathbf{v}\in (L^2(\Omega))^3 \; | \;\nabla\times \mathbf{v}\in (L^2(\Omega))^3\},
$$
$$
H_0(\text{curl};\Omega) = \{\mathbf{v}\in H(\text{curl};\Omega)\; | \;\mathbf{v}\times \mathbf{n} = 0, \text{on}\;\partial\Omega\},
$$
$$
V=\{\mathbf{v}\in H_0(\text{curl};\Omega)\; |\; \nabla\times \mathbf{v}\in H_0^1(\Omega)\}.
$$
$V$ is a Hilbert space with scalar product and norm given by
$$
(\mathbf{u},\mathbf{v})_{V}\triangleq (\nabla(\nabla\times \mathbf{u}),\nabla (\nabla\times
\mathbf{v}))+(\nabla\times \mathbf{u},\nabla\times \mathbf{v})+(\mathbf{u},\mathbf{v}),
$$
$$
\norm{\mathbf{u}}_{V}\triangleq \sqrt{(\mathbf{u},\mathbf{u})_{V}}.
$$
The following lemma gives a sufficient condition for a piecewisely defined function to be an element in $V$.
\begin{lemma}
If $\mathbf{v}$ is piecewise smooth, $\mathbf{v}\times \mathbf{n}$ and $\nabla\times \mathbf{v}$ are continuous across element interfaces, then
$\mathbf{v}\in V$.
\end{lemma}
Using the following identity:
\begin{equation*}
(\nabla\times)^2\mathbf{u}=-\Delta\mathbf{u}+
\nabla(\nabla\cdot\mathbf{u})
\end{equation*}
and $\nabla\cdot \mathbf{u}=0$, the first equation in (\ref{four_curl_bvp}) can be rewritten in the following form:
\begin{equation}
-\alpha\nabla\times\Delta(\nabla\times \mathbf{u}) + \beta(\nabla\times)^2 \mathbf{u} +\gamma \mathbf{u}=\mathbf{f}.
\label{delta_curlcurl}
\end{equation}
Multiplying Equation (\ref{delta_curlcurl}) by the test function $\mathbf{v}$ and using integration by parts, we obtain the following variational formulation:
\begin{equation}
\text{Find}\; \mathbf{u}\in V\; \text{such that}\; a(\mathbf{u},\mathbf{v})=(\mathbf{f},\mathbf{v}), \;\forall\;\mathbf{v}\in V,\label{V_1_variational}
\end{equation}
where the bilinear form $a(\cdot,\cdot)$ defined on $V\times
V$ is given by
$$
a(\mathbf{u},\mathbf{v})=\alpha (\nabla(\nabla\times \mathbf{u}),\nabla(\nabla\times \mathbf{v}))+ \beta (\nabla\times
\mathbf{u},\nabla\times \mathbf{v}) + \gamma (\mathbf{u},\mathbf{v}).
$$
The well-posedness of the above variational problem follows from the
Lax-Milgram lemma.
The next lemma indicates that the weak solution satisfies the divergence-free constraint.
\begin{lemma}
Assume $\nabla\cdot \mathbf{f} = 0$, and let $\mathbf{u}$ be the solution of problem (\ref{V_1_variational}). Then $\nabla\cdot \mathbf{u} = 0$.
\end{lemma}
\begin{proof}
Choose test function $\mathbf{v}=\nabla \varphi$ where $\varphi\in C_0^\infty(\Omega)$, then
$$
(\mathbf{u},\nabla\varphi) = (\mathbf{f},\nabla\varphi),
$$
hence, $\nabla\cdot \mathbf{u} =\nabla\cdot \mathbf{f} = 0$.
\end{proof}
\section{A Nonconforming Finite Element}
In this section, we construct a nonconforming finite element to solve the fourth-order equation. One of the advantages for using a nonconforming element is that the number of degrees of freedom is small compared to that for conforming elements. The following construction is based on N\'{e}d\'{e}lec elements of the first family that consist of incomplete vector polynomials \cite{Nedelec:1980xe}. The advantage of using incomplete vector polynomial space is that it provides the same order of convergence in terms of energy norms as the one given by corresponding complete polynomial space. In the following, we define the degrees of freedom in a special way to ensure that the consistency error estimate holds.
\begin{definition}
The finite element triple
$(K,P_K,\Sigma_K)$ is defined by
\begin{itemize}
\item $K$ is a tetrahedron;
\item $\mathcal{P}_K=R_2(K)=\mathbf{P}_{1}\oplus \{\mathbf{p}\in
(\widetilde{P}_2)^3\;|\;\mathbf{p}\cdot \mathbf{x} = 0\}$, where $\widetilde{P}_2$ is the space of homogeneous multivariate polynomials of degree $2$;
\item $\Sigma_K$ is the set of degrees of freedom, see Figure \ref{element},
\begin{itemize}
\item edge degrees of freedom:
\begin{equation}
M_e(\mathbf{u})=\bigg\{\int_e \mathbf{u}\cdot \tau\;q\;d s\;|\;\forall
\;q\in P_1(e),\;\forall\;e\subset K\bigg\},
\label{edge d.o.f}
\end{equation}
where $\tau$ is the unit tangential vector along the edge $e$,
\item face degrees of freedom:
\begin{equation}
M_f(\mathbf{u})=\bigg\{\frac{1}{|f|^2}\int_f (\nabla\times \mathbf{u})\times \mathbf{n} \cdot
q \;d A\;|\;\forall\;q\in
(P_0(f))^2,\;\forall\;f\subset K\bigg\},
\label{face d.o.f}
\end{equation}
where $\mathbf{n}$ is the unit normal vector to the face $f$,
\end{itemize}
$\Sigma_K=M_e(\mathbf{u})\cup M_f(\mathbf{u})$.
\end{itemize}
\label{fem_triple_def}
\end{definition}
\begin{figure}[h]
\begin{center}
\includegraphics[width=60mm]{element.png}
\caption{Degrees of freedom of the finite element}
\label{element}
\end{center}
\end{figure}
In the above finite element triple, the space $\mathcal{P}_K$ is the same as the second order N\'{e}d\'{e}lec element of the first family for $H(\text{curl})$ problem. The difference is the definition of the second set of degrees of freedom. It is designed specifically to ensure consistency for the fourth-order problems. The total number of the degrees of freedom for this element is $20$, which is the same as the dimension of the polynomial space $R_2(K)$.
It should be pointed out that the scaling factor $1/|f|^2$ in the definition of the second set of degrees of freedom is associated with the construction of basis functions to be given later.
The next lemma given in \cite{Nedelec:1980xe} describes a relation between edge integrals and face integrals which will be useful in the error analysis.
\begin{lemma}
If $u\in R_2(K)$ is such that the edge degrees of freedom (\ref{edge d.o.f}) vanish, then
\begin{equation}
\int_f(\nabla\times \mathbf{u})\cdot \mathbf{n}\;dA=0,\;\forall\;\text{face}\;f\subset K.
\label{curl_u_face_integral}
\end{equation}
\label{relation-edge-face}
\end{lemma}
\begin{proof}
Given $\mathbf{u}\in R_2(K)$ satisfies
$$
\int_e \mathbf{u}\cdot\tau \;q\;ds=0,\;\forall \;\text{edge}\; e\subset K.
$$
By Stokes' Theorem,
$$
\int_f (\nabla_f\times \mathbf{u}_T)\cdot q \;dA- \int_f (\vec{\nabla}_f\times q)\cdot \mathbf{u}_T\;dA = \int_{\partial f} \mathbf{u}\cdot \tau\; q\;ds,
$$
where $\mathbf{u}_T$ is the tangential part of $\mathbf{u}$, and $\vec{\nabla}_f\times$ and $\nabla_f\times$ are surface vector curl and surface scalar curl, respectively. Let $q$ be a constant. Notice that
$$
\nabla_f \times \mathbf{u}_T = (\nabla\times \mathbf{u})\cdot \mathbf{n},
$$
we conclude,
$$
\int_f(\nabla\times \mathbf{u})\cdot \mathbf{n}\;dA=0.
$$
\end{proof}
As a direct consequence of Lemma \ref{relation-edge-face}, if both the edge degrees of freedom (\ref{edge d.o.f}) and face degrees of freedom (\ref{face d.o.f}) vanish, then
$$
\int_f (\nabla\times \mathbf{u}) \;dA= 0.
$$
The polynomial space $R_2(K)$ has the following property \cite{Girault:1986fk}.
\begin{lemma}
If $\mathbf{u}\in R_2(K)$ satisfies $\nabla\times \mathbf{u} = 0$, then
$$
\mathbf{u}=\nabla p,\;\text{with}\;p\in P_2.
$$
\label{R2lemma}
\end{lemma}
We recall that the finite element $(K,\mathcal{P}_K,\Sigma_K)$ is said to be unisolvent if a function in $\mathcal{P}_K$ can be uniquely determined by specifying values for degrees of freedom in $\Sigma_K$.
\begin{proposition}
The finite element defined by Definition \ref{fem_triple_def} is unisolvent.
\end{proposition}
\begin{proof}
It is sufficient to prove that, given $\mathbf{u}\in R_2(K)$,
$$
M_e(\mathbf{u}) = M_f(\mathbf{u}) = 0,\;\forall e\subset K, f\subset K\Rightarrow \mathbf{u} = 0.
$$
Obviously, $
\nabla (\nabla\times \mathbf{u})$ is a constant vector.
Then using (\ref{curl_u_face_integral}) and integration by parts, we obtain
$$
\nabla(\nabla\times \mathbf{u}) = \frac{1}{|K|}\int_K \nabla(\nabla\times \mathbf{u}) dx = \frac{1}{|K|}\int_{\partial K} (\nabla\times \mathbf{u})\mathbf{n}^T dA = 0.
$$
This implies that
$$
\nabla(\nabla\times \mathbf{u}) = 0 \Rightarrow \nabla\times \mathbf{u} = \text{const}.
$$
Using again (\ref{curl_u_face_integral}), we have
\begin{equation*}
\nabla\times \mathbf{u} = 0.
\end{equation*}
By Lemma \ref{R2lemma}, we have
$$
\mathbf{u}=\nabla p,\;\text{with}\;p\in P_2(K).
$$
Since $M_e(\mathbf{u}) = 0$, we have
$$
\int_e \frac{\partial p}{\partial \tau}\; q\; ds = 0, \forall q\in P_1(e).
$$
This implies $\partial p/\partial \tau = 0$ on each edge $e$. Hence, $p$ is constant and $\mathbf{u}=0$.
\end{proof}
In the following, we construct the basis functions. The explicit form of these basis functions not only is useful for implementation, but also instrumental for the interpolation error estimate.
\subsection{Basis functions}
The main idea of the construction is to consider linear combinations of basis functions of a related N\'{e}d\'{e}lec element. Let $K$ be an arbitrary tetrahedron with four vertices $a_i$, $a_j$, $a_k$ and $a_l$, see Figure \ref{element2}. The corresponding barycentric coordinates are given by
$\lambda_i$, $\lambda_j$, $\lambda_k$, and $\lambda_l$, respectively.
\begin{figure}[h]
\begin{center}
\includegraphics[width=60mm]{element2.png}
\caption{A tetrahedron with vertices $a_i$, $a_j$, $a_k$, $a_l$. $\mathbf{q}_{ij}$ and $\mathbf{q}_{ik}$} are two tangential vectors on the face $F_l$.
\label{element2}
\end{center}
\end{figure}
On each of the four faces, say face $l$ (with vertices $a_i, a_j, a_k$), we choose the following two tangential direction vectors:
$$
\mathbf{q}_{ij} = \overrightarrow{a_j a_i} = 6|K|(\nabla\lambda_l\times \nabla\lambda_k),
$$
$$
\mathbf{q}_{ik} = \overrightarrow{a_k a_i} =
6|K|(\nabla\lambda_j\times \nabla\lambda_l).
$$
The edge degrees of freedom on edge $e_{ij}$ (with vertices $a_i$ and $a_j$) are defined explicitly by:
$$
M_{ij}^{(1)}(\mathbf{u}) = \int_{e_{ij}} \mathbf{u}\cdot\tau \;ds,
$$
$$M_{ij}^{(2)}(\mathbf{u}) = \int_{e_{ij}} \mathbf{u}\cdot
\tau\left(3-\frac{6}{|e_{ij}|}s\right)\;ds,
$$
where $\tau$ is the unit direction vector of edge $e_{ij}$, $s$ is an arc length parameter. The face degrees of freedom are defined as:
$$
M_{lij}(\mathbf{u}) = \frac{1}{|f_l|^2} \int_{f_{l}}(\nabla \times \mathbf{u})\times \mathbf{n}_l\cdot \mathbf{q}_{ij}\;dA,
$$
$$
M_{lik}(\mathbf{u}) = \frac{1}{|f_l|^2} \int_{f_{l}}(\nabla \times \mathbf{u})\times \mathbf{n}_l\cdot \mathbf{q}_{ik}\;dA,
$$
where $\mathbf{n}_l$ is the unit outward normal vector of the face $f_l$.
We recall that the basis functions of the second order N\'{e}d\'{e}lec element of the first family in barycentric coordinates are (see, e.g., \cite{Gopalakrishnan:2005eu}, \cite{Sun:2008yq}, \cite{Webb:1999sh}):
(1) Two basis functions on each edge $e_{ij}$:
$$
\mathbf{L}_{ij} = \lambda_i\nabla\lambda_j-\lambda_j\nabla\lambda_i,
$$
$$
\mathbf{L}_{ji} = \lambda_i\nabla\lambda_j+\lambda_j\nabla\lambda_i.
$$
(2) Two basis functions on each face $f_l$:
$$
\mathbf{L}_{ijk} = \lambda_i(\lambda_j\nabla\lambda_k - \lambda_k\nabla\lambda_j),
$$
$$
\mathbf{L}_{jik} = \lambda_j(\lambda_i\nabla\lambda_k - \lambda_k\nabla\lambda_i).
$$
In the following, we list a few useful facts about the geometry of a tetrahedron.
(1) The unit outward normal vector of face $f_l$ is given by
$$
-\frac{\nabla \lambda_l}{\|\nabla\lambda_l\|}.
$$
(2) The two tangential vectors of face $f_l$ are given by $\mathbf{q}_{ij}$ and $\mathbf{q}_{ik}$.
(3) Let $h_l$ be the height of the tetrahedron corresponding to the face $f_l$, then
$$
\nabla\lambda_l=\frac{1}{6|K|}\mathbf{q}_{ik}\times \mathbf{q}_{jk},
$$
$$
|\nabla\lambda_l| = \frac{1}{h_l}.
$$
(4) Let $|K|$ be the volume of the tetrahedron $K$, then
$$
6|K|=|\mathbf{q}_{il}\cdot(\mathbf{q}_{jl}\times \mathbf{q}_{kl})| = \frac{-1}{(\nabla\lambda_i\times \nabla\lambda_j)\cdot \nabla\lambda_k }.
$$
Next, we construct basis functions in barycentric coordinates. They provide a set of dual basis functions with respect to the prescribed degrees of freedom.
{\bf Step 1}. Construct eight basis functions $\{\phi_{lij}\}$ corresponding to the face degrees of freedom such that
\begin{equation}
M_{mn}^{(t)}(\phi_{lij})=0,\label{face_function_edge}
\end{equation}
and
\begin{equation}
M_{mnp}(\phi_{lij})=\delta_{ml}\delta_{ni}
\delta_{pj}.\label{face_function_face}
\end{equation}
We use the basis functions of the second order N\'{e}d\'{e}lec element as building blocks as they automatically satisfy the first condition (\ref{face_function_edge}). Using the facts listed above, we find that the basis functions corresponding to the facial degrees of freedom on face $f_l$ are given by the following:
$$
\phi_{lij} = 3|K|(\mathbf{L}_{lij}-\mathbf{L}_{ljk}),
$$
$$
\phi_{lik} = 3|K|(\mathbf{L}_{lik}-\mathbf{L}_{ljk}).
$$
By direct calculation, we have
\begin{eqnarray*}
&&\int_{f_l}(\nabla\times \mathbf{L}_{lij})\times \nabla\lambda_l\cdot(\nabla\lambda_l\times\nabla\lambda_k)\;dA\\
& = & \int_{f_l}\left[2\lambda_l(\nabla\lambda_i\times\nabla\lambda_j)
+\lambda_i(\nabla\lambda_l\times\nabla\lambda_j)-\lambda_j
(\nabla\lambda_l\times\nabla\lambda_i)\right]\\
&& \cdot\left[\nabla\lambda_l(\nabla_l\cdot\nabla_l)
-\nabla\lambda_k(\nabla\lambda_l\cdot\nabla\lambda_k)
\right]\;dA\\
& = & -\left(\int_{f_l}\lambda_i \;dA\right)[(\nabla\lambda_l\times\nabla\lambda_j)
\cdot \nabla\lambda_k](\nabla\lambda_l\cdot\nabla\lambda_l)\\
&& + \left(\int_{f_l}\lambda_j \;dA\right)[(\nabla\lambda_l\times\nabla\lambda_i)
\cdot \nabla\lambda_k](\nabla\lambda_l\cdot\nabla\lambda_l)\\
& = & -\frac{2}{3}|f_l|\frac{1}{6|K|h_l^2},
\end{eqnarray*}
and
\begin{eqnarray*}
\int_{f_l}(\nabla\times \mathbf{L}_{ljk})\times \nabla\lambda_l\cdot(\nabla\lambda_l\times\nabla\lambda_k)\;dA = \frac{1}{3}|f_l|\frac{1}{6|K|h_l^2}.
\end{eqnarray*}
Hence,
\begin{eqnarray*}
M_{lij}(\phi_{lij})&=&\frac{1}{|f_l|^2}\int_{f_l}
(\nabla\times \phi_{lij})\times n_l\cdot \mathbf{q}_{ij}\;dA\\
&=& \frac{1}{|f_l|^2}\int_{f_l}\nabla\times\left[
3|K|(\mathbf{L}_{lij}-\mathbf{L}_{ljk})\right]
\times -\frac{\nabla\lambda_l}{\|\nabla\lambda_l\|}
\cdot(6|K|\nabla\lambda_l\times
\nabla\lambda_k)dA
\\
&=& -\frac{18|K|^2h_l}{|f_l|^2}\int_{f_l}\nabla\times
(\mathbf{L}_{lij}
-\mathbf{L}_{ljk})\times\nabla\lambda_l\cdot (\nabla\lambda_l
\times\nabla\lambda_k)\;dA\\
&=& -\frac{18|K|^2h_l}{|f_l|^2}
\left(-\frac{2}{3}|f_l|\frac{1}{6|K|h_l^2}
-\frac{1}{3}|f_l|\frac{1}{6|K|h_l^2}\right) = 1.
\end{eqnarray*}
Similarly, $M_{lik}(\phi_{lij}) = 0$, and $M_{f_{l'}}(\phi_{lij})=0$ where $l'\neq l$.
{\bf Step 2}. Construct twelve basis functions $\{\psi_{ij}^{(t)}|1\leq i<j\leq 4, t=1,2\}$ corresponding to the edge degrees of freedom such that
\begin{equation}
M_{mn}^{(t')}(\psi_{ij}^{(t)}) = \delta_{t't}\delta_{mn,ij},\label{edge_function_edge}
\end{equation}
\begin{equation}
M_{mnp}(\psi_{ij}) = 0.\label{edge_function_face}
\end{equation}
Here, we use the edge basis functions of the second order N\'{e}d\'{e}lec element as building blocks since they satisfy condition (\ref{edge_function_edge}). Since $\nabla\times \mathbf{L}_{ji}=0$, $\mathbf{L}_{ji}$ automatically satisfy condition (\ref{edge_function_face}).
For functions $\mathbf{L}_{ij}$, we need to subtract from them a linear combination of face basis functions so that (\ref{edge_function_edge}) and (\ref{edge_function_face}) hold. This can be done because by construction, our face basis functions have no edge moments. This strategy for constructing basis functions can be found in \cite{Gopalakrishnan:2005eu,Sun:2008yq}.
Finally, we can write the basis functions of the new element as the following:
(1) Two basis functions on each face $l$ ($1\leq l\leq 4$):
$$
\phi_{lij} = 3|K|(\mathbf{L}_{lij}-\mathbf{L}_{ljk}),
$$
$$
\phi_{lik} = 3|K|(\mathbf{L}_{lik}-\mathbf{L}_{ljk}),
$$
where $\mathbf{L}_{lij}= \lambda_i(\lambda_j\nabla\lambda_k - \lambda_k\nabla\lambda_j)$.
(2) Two basis functions on each edge $e_{ij}$ ($1\leq i<j\leq 4$):
$$
\psi_{ij}^{(1)}=\mathbf{L}_{ji},
$$
$$
\psi_{ij}^{(2)}=\mathbf{L}_{ij}-\sum M_{mnp} (\mathbf{L}_{ij})\phi_{mnp}.
$$
\subsection{Convergence analysis}
Let $\mathcal{T}_h=\{K_i\}_{i=1}^{N_h}$ be a triangulation
of the domain $\Omega$. On this triangulation we introduce the finite element space $V_h$ and define the discrete norm $\|\cdot\|_h$ by
$$
\| \mathbf{v}\|_h=\left[\sum_{K\in\mathcal{T}_h}
\left(\|\mathbf{v}\|_{0,K}^2+\|\nabla\times \mathbf{v}\|_{0,K}^2+\|\nabla(\nabla\times \mathbf{v})\|_{0,K}^2\right)\right]^{1/2}.
$$
Consider the following discrete bilinear form:
\begin{equation*}
\begin{split}
a_h(\mathbf{u}_h,\mathbf{v}_h)
=\sum_{K\in\mathcal{T}_h} \alpha(\nabla(\nabla\times \mathbf{u}_h),\nabla(\nabla\times
\mathbf{v}_h))_{L^2(K)}&+\beta (\nabla\times \mathbf{u}_h,\nabla\times
\mathbf{v}_h)_{L^2(K)}\\
&+ \gamma (\mathbf{u}_h,\mathbf{v}_h)_{L^2(K)}.
\end{split}
\end{equation*}
It is straightforward to verify that the bilinear
form $a_h$ satisfies
$$
a_h(\mathbf{v},\mathbf{v})\gtrsim \|\mathbf{v}\|_h^2,\;\forall\; \mathbf{v}\in V_h,
$$
$$
|a_h(\mathbf{u},\mathbf{v})|\lesssim
\|\mathbf{u}\|_h\|\mathbf{v}\|_h,\;
\forall\;\mathbf{u}\in V+V_h, \mathbf{v}\in V_h.
$$
The nonconforming finite element discretization of problem
(3.10) is:
Find $\mathbf{u}_h\in V_h$, such that for all $\mathbf{v}_h\in V_h$,
\begin{equation}
a_h(\mathbf{u}_h,\mathbf{v}_h)=(\mathbf{f},\mathbf{v}_h).\label{discrete var prob1}
\end{equation}
The convergence of the above finite element approximation can be analyzed through the following second Strang lemma \cite{Ciarlet:1978qf}.
\begin{lemma}
$$
\|\mathbf{u}-\mathbf{u}_h\|_h\lesssim \inf_{\mathbf{v}_h\in V_h}\|\mathbf{u}-\mathbf{v}_h\|_h +
\sup_{\mathbf{w}_h\in V_h}\frac{|a_h(\mathbf{u},\mathbf{w}_h)-
(\mathbf{f},\mathbf{w}_h)|}{\|\mathbf{w}_h\|_h},
$$
where the first term on the right-hand side is called the interpolation error and the second term is called the consistency error.
\end{lemma}
In order to estimate the consistency error we first define an average operator $P_{f}$ on a face $f$ by
$$
P_{f} \mathbf{w} = \frac{1}{|f|}\int_{f} \mathbf{w} \;dA.
$$
Since for any $\mathbf{v}_h\in V_h$, the quantity $\int_f \nabla\times \mathbf{v}_h\;dA$ is continuous, we know that $P_f$ is well-defined for $\nabla\times \mathbf{v}_h$.
The following two lemmas are standard results.
\begin{lemma} Given any face $f\subset K$ and $\mathbf{w}\in (H^1(K))^3$,
$$
\int_{f} |\mathbf{w}-P_{f}\mathbf{w}|^2 \;dA \lesssim h_K|\mathbf{w}|^2_{1,K}.
$$\label{face_estimate_lemma}
\end{lemma}
\begin{lemma}
$$
\int_{\partial K}|\mathbf{w}|^2 \;dA \lesssim
h_K^{-1}\norm{\mathbf{w}}_{0,K}^2+h_K|
\mathbf{w}|_{1,K}^2.
$$\label{trace_lemma}
\end{lemma}
Next, we estimate the interpolation error and consistency error separately.
\subsubsection{Interpolation error estimate}
Let $K$ and $K'_f$ be the two tetrahedra sharing a common face $f$, $r_K$ be the local interpolation operator for the second order N\'{e}d\'{e}lec element of the first family, namely, given $\mathbf{u}\in V$, define $r_K \mathbf{u}$ such that
$$
\int_e r_K \mathbf{u}\cdot \tau\;ds = \int_e \mathbf{u}\cdot\tau\;ds,\;\forall\;\text{edge}\;
e\subset K,
$$
and
$$
\int_f (r_K \mathbf{u}\times \mathbf{n})
\cdot q\;dA = \int_f (\mathbf{u}\times \mathbf{n})\cdot \mathbf{q}\;dA,\;\forall\;\mathbf{q}\in (P_0(f))^2,\;\forall\;\text{face}\;f\subset K.
$$
Define $\mathbf{u}_I\in V_h$ such that
$$
M_e(\mathbf{u}_I)=M_e(r_K \mathbf{u})=M_e(\mathbf{u}),
$$
$$
M_f(\mathbf{u}_I)=[M_f(r_{K} \mathbf{u})+M_f(r_{K'_f} \mathbf{u})]/2.
$$
If $f\subset \partial \Omega$, we set $M_f(\mathbf{u}_I)=M_f(r_{K} \mathbf{u})$.
\begin{lemma} Given $\mathbf{u}\in V$, let $\mathbf{u}_I$ be defined as above, then
$$
\|\mathbf{u}-\mathbf{u}_I\|_h \lesssim h(|\mathbf{u}|_2+|\nabla\times \mathbf{u}|_2).
$$
\end{lemma}
\begin{proof}
Let $r_h \mathbf{u}$ be the global interpolation operator defined by $r_h \mathbf{u}|_K = r_K \mathbf{u}$.
By triangle inequality,
$$
\|\mathbf{u}-\mathbf{u}_I\|_h\leq \|\mathbf{u}-r_h \mathbf{u}\|_h+\|r_h \mathbf{u} - \mathbf{u}_I\|_h.
$$
By the interpolation error estimate of N\'{e}d\'{e}lec element, we have
$$
\|\mathbf{u} - r_h \mathbf{u}\|_h \lesssim h(|\mathbf{u}|_2+|\nabla\times \mathbf{u}|_2).
$$
Notice that on each tetrahedron $K$,
$$
r_K \mathbf{u} - \mathbf{u}_I = \sum_{f\subset K}\sum M_{mnp} (r_K \mathbf{u} - \mathbf{u}_I)\phi_{mnp},
$$
where $\{\phi_{mnp}\}$ are basis functions on face $f$, and $\{M_{mnp}(\cdot)\}$ are degrees of freedom on face $f$.
Using Lemma \ref{trace_lemma} and $\|q_{np}\|_{L^2(f)} = O(h^2)$, we get
\begin{eqnarray*}
&&2|M_{mnp}(r_K \mathbf{u} - \mathbf{u}_I)|=|M_{mnp}(r_{K'_f} \mathbf{u})-M_{mnp}(r_{K} \mathbf{u})|\\
&&=\frac{1}{|f|^2}\bigg|\int_f(\nabla\times r_{K'_f} \mathbf{u} - \nabla\times r_{K} \mathbf{u} )\times \mathbf{n}\cdot \mathbf{q}_{np}\;dA\bigg|\\
&&\leq \frac{1}{|f|^2}\bigg|\int_f (\nabla\times(r_{K'_f}\mathbf{u} - \mathbf{u})\times \mathbf{n} \cdot \mathbf{q}_{np}\;dA\bigg| + \frac{1}{|f|^2}\bigg|\int_f (\nabla\times(r_{K}\mathbf{u} - \mathbf{u})\times \mathbf{n} \cdot \mathbf{q}_{np}\;dA\bigg| \\
&&\lesssim \frac{\|\mathbf{q}_{np}\|_{L^2(f)}}{|f|^2}
(\|\nabla\times(r_{K'_f}\mathbf{u} - \mathbf{u})\times \mathbf{n}\|_{L^2(f)} +\|\nabla\times (r_{K}\mathbf{u} - \mathbf{u})\times \mathbf{n}\|_{L^2(f)})\\
&&\lesssim h^{-2}(h^{-1/2}\|\nabla\times(r_{K'_f}\mathbf{u} - \mathbf{u})\|_{0, K\cup K'} + h^{1/2}| \nabla\times (r_{K}\mathbf{u} - \mathbf{u})|_{1,K\cup K'})\\
&&\lesssim h^{-1/2}|\nabla\times \mathbf{u}|_{2,K\cup K'}.
\end{eqnarray*}
Notice $\nabla\lambda_i=O(h^{-1})$, and $\|\phi_{mnp}\|_{0,K}^2=O(h^7)$, by Cauchy-Schwarz inequality, we have
\begin{eqnarray*}
\|r_K \mathbf{u} - \mathbf{u}_I\|_{0,K} &\leq &\left(\sum_{f\subset K}\sum |M_{mnp}(r_K \mathbf{u} - r_I \mathbf{u})|^2\right)^{1/2}\left( \sum_{f\subset K}\sum \|\phi_{mnp}\|^2_{0,K}\right)^{1/2}\\
& \lesssim & h^3|\nabla\times \mathbf{u}|_{2,S(K)},
\end{eqnarray*}
where $S(K)=\cup_{K'\in\mathcal{T}_h, K'\cap K\neq \emptyset} K'$.
Hence,
\begin{equation}
\|r_h \mathbf{u} - \mathbf{u}_I\|_{0,\Omega} = (\sum_K \|r_K \mathbf{u} - \mathbf{u}_I\|^2_{0,K})^{1/2}\lesssim h^3|\nabla\times \mathbf{u}|_{2,\Omega}.
\label{ruuI}
\end{equation}
By inverse inequality, we have
$$
\|\nabla\times (r_h \mathbf{u} - \mathbf{u}_I)\|_{0,\Omega}\lesssim h^2|\nabla\times \mathbf{u}|_{2,\Omega},
$$
$$
\|\nabla(\nabla\times (r_h \mathbf{u} - \mathbf{u}_I))\|_{0,\Omega}\lesssim h|\nabla\times \mathbf{u}|_{2,\Omega}.
$$
Combining these estimates, we get
$$
\|r_h \mathbf{u} - \mathbf{u}_I\|_h\lesssim h|\nabla\times \mathbf{u}|_{2,\Omega},
$$
and the desired estimate follows.
\end{proof}
Remark: We note that the error estimate (\ref{ruuI}) indicates that $r_h \mathbf{u}$ and $\mathbf{u}_I$ are super-close. Such type of estimate can not usually be obtained by the standard scaling argument (using Bramble-Hilbert lemma). In our proof, we made use of the detailed information of the basis functions constructed in the previous section.
\subsubsection{Consistency error estimate}
Given a tetrahedron $K$, in addition to the local interpolation operator $r_K$, we introduce another local interpolation operator $\tilde{r}_K$ corresponding to the first order N\'{e}d\'{e}lec element of the second family, namely, $\tilde{r}_K \mathbf{u}\in (P_1(K))^3\subset R_2(K)$, and
$$
\int_e ((\tilde{r}_K \mathbf{u})\cdot \tau ) \;q\;ds = \int_e
(\mathbf{u}\cdot\tau)\; q\;ds,\;\forall\;q\in P_1(e),\;\forall\;\text{edge}\;e\subset K.
$$
Consider two tetrahedra $K$ and $K'_f$ that share a common face $f$. Given $\mathbf{v}_h\in V_h$, define $\mathbf{v}_K=\mathbf{v}_h|_K$. By definition,
$$
\tilde{r}_{K}\mathbf{v}_{K} = \tilde{r}_{K'_f}\mathbf{v}_{K'_f},\;\text{on face}\;f.
$$
Hence,
\begin{equation}
\sum_K \int_{\partial K} \varphi\cdot [(\tilde{r}_K\mathbf{v}_K)\times \mathbf{n}] \;dA = 0,\;\forall\;\varphi\in H(\text{curl};\Omega),
\label{nedelec face equality}
\end{equation}
where $\mathbf{n}$ is the unit outward normal vector of $\partial K$.
Consider the decomposition (see \cite{Girault:1986fk}):
$$
\mathbf{v}_K = \nabla p_K + \mathbf{w}_K,
$$
where $\text{div}\;\mathbf{w}_K = 0$, $\mathbf{w}_K\cdot \mathbf{n}|_{\partial K}=0$, and $p_K\in P_2(K)$.
The following Lemma \ref{hu-zou-lemma} can be found in \cite{Hu:2004fp}:
\begin{lemma}\
$$
\|\tilde{r}_K \mathbf{w}_K - \mathbf{w}_K\|_{0,K}\lesssim h\|\nabla\times \mathbf{v}_K\|_{0,K}.
$$
\label{hu-zou-lemma}
\end{lemma}
As a consequence of Lemma \ref{hu-zou-lemma}, we have the following estimate.
\begin{lemma}
$$
\|\tilde{r}_K \mathbf{v}_K - \mathbf{v}_K\|_{0,K}\lesssim h\|\nabla\times \mathbf{v}_K\|_{0,K}.
$$
\label{wk lemma}
\end{lemma}
\begin{proof}
Using the interpolation operators defined above, we obtain
$$
\tilde{r}_K\mathbf{v}_K = \tilde{r}_K\nabla p_K + \tilde{r}_K\mathbf{w}_K = \nabla p_K +\tilde{r}_K\mathbf{w}_K.
$$
Hence,
$$
\tilde{r}_K \mathbf{v}_K- \mathbf{v}_K = \tilde{r}_K \mathbf{w}_K - \mathbf{w}_K.
$$
By Lemma \ref{hu-zou-lemma}, we obtain
$$
\|\tilde{r}_K \mathbf{v}_K - \mathbf{v}_K\|_{0,K}\lesssim h\|\nabla\times \mathbf{v}_K\|_{0,K}.
$$
\end{proof}
Now, we can show the following lemma, which is critical for the consistency error estimate.
\begin{lemma} For $\varphi\in H(\text{curl};\Omega)$,
$$
|\sum_{K}\int_{\partial K}\varphi \cdot (\mathbf{v}_h\times \mathbf{n})\;dA
\lesssim h(\|\varphi\|_{0,\Omega}+\|\nabla\times \varphi\|_{0,\Omega})\left(\sum_K\|\nabla\times \mathbf{v}_h\|_{1,K}^2\right)^{1/2}.
$$
\label{critical lemma for consistency}
\end{lemma}
\begin{proof}
By the interpolation error estimates of the N\'{e}d\'{e}lec elements
$$
\|\nabla\times (\tilde{r}_K\mathbf{v}_K -\mathbf{v}_K)\|_{0,K}\lesssim h \|\nabla\times \mathbf{v}_K\|_{1,K},
$$
Lemma \ref{wk lemma}, and Equation (\ref{nedelec face equality}), we have
\begin{eqnarray*}
&&\bigg|\sum_K \int_{\partial K}\varphi\cdot (\mathbf{v}_K\times \mathbf{n})\;dA\bigg|
= \bigg|\sum_K\int_{\partial K}\varphi\cdot [(\tilde{r}_K \mathbf{v}_K-\mathbf{v}_K)\times \mathbf{n}]\;dA\bigg|\\
&&= \bigg|\sum_K\int_K (\nabla\times \varphi)\cdot (\tilde{r}_K\mathbf{v}_K-\mathbf{v}_K)\;dx
- \varphi\cdot [\nabla\times (\tilde{r}_K\mathbf{v}_K-\mathbf{v}_K)]\;dx\bigg| \\
&&\leq \sum_K(\|\nabla\times\varphi\|_{0,K}\|\tilde{r}_K
\mathbf{v}_K-\mathbf{v}_K\|_{0,K}
+\|\varphi\|_{0,K}\|\nabla\times(\tilde{r}_K
\mathbf{v}_K-\mathbf{v}_K)\|_{0,K})\\
&&\lesssim h(\|\varphi\|_{0,\Omega}+
\|\nabla\times\varphi\|_{0,\Omega})
\left(\sum_K\|\nabla\times \mathbf{v}_h\|_{1,K}^2\right)^{1/2}.
\end{eqnarray*}
\end{proof}
Next, we show the consistency error estimate for the nonconforming finite element approximation defined above.
\begin{theorem}
Assume that $\mathbf{u}\in V$ is sufficiently smooth and $\mathbf{v}_h\in V_h$, then
\begin{multline*}
|a_h(\mathbf{u},\mathbf{v}_h)-(\mathbf{f},\mathbf{v}_h)|
\lesssim h(\| \nabla\times \Delta(\nabla\times \mathbf{u})\|+|\nabla(\nabla\times \mathbf{u})|_1\\+\|(\nabla\times)^2 \mathbf{u}\|+\|\nabla\times \mathbf{u}\|)\left(\sum_K\|\nabla\times
\mathbf{v}_h\|_{1,K}^2\right)^{1/2}.
\end{multline*}
\end{theorem}
\begin{proof}
By applying integration by parts, we get
\begin{eqnarray*}
&&\;\;\;\;\left(\nabla(\nabla\times \mathbf{u}),\nabla(\nabla\times \mathbf{v}_h)\right)_K \\&=& -(\Delta(\nabla\times \mathbf{u}),\nabla\times \mathbf{v}_h)_K+(\nabla(\nabla\times \mathbf{u})\cdot \mathbf{n},\nabla\times \mathbf{v}_h)_{\partial K}\\
&=&-(\nabla\times\Delta(\nabla\times \mathbf{u}), \mathbf{v}_h)_K
+(\Delta (\nabla\times \mathbf{u}),\mathbf{v}_h\times \mathbf{n})_{\partial K}\\
&&+(\nabla(\nabla\times \mathbf{u})\cdot \mathbf{n},\nabla\times \mathbf{v}_h)_{\partial K},
\end{eqnarray*}
and
$$
(\nabla\times \mathbf{u},\nabla\times \mathbf{v}_h)_K
=((\nabla\times)^2 \mathbf{u},\mathbf{v}_h)_K-(\nabla\times \mathbf{u}, \mathbf{v}_h\times \mathbf{n})_{\partial K}.
$$
Hence,
\begin{eqnarray*}
&&\;\;\;\; a_h(\mathbf{u},\mathbf{v}_h)-(\mathbf{f},\mathbf{v}_h)\\
&&=\sum_{K\in\mathcal{T}_h}[\alpha(\Delta(\nabla\times \mathbf{u}), \mathbf{v}_h \times
\mathbf{n})_{\partial K}+\alpha(\nabla(\nabla\times \mathbf{u})\cdot \mathbf{n},\nabla\times
\mathbf{v}_h)_{\partial K}\\
&& \quad\quad\quad\quad- \beta(\nabla\times \mathbf{u},\mathbf{v}_h\times \mathbf{n})_{\partial K}]\\
&&=\sum_{K\in\mathcal{T}_h}\left[(\alpha \Delta (\nabla\times \mathbf{u})-\beta\nabla\times \mathbf{u},\mathbf{v}_h\times \mathbf{n})_{\partial K}\right]\\
&&\quad\quad+\sum_{K\in\mathcal{T}_h}\left[
\alpha(\nabla(\nabla\times \mathbf{u})\cdot \mathbf{n}, \nabla\times \mathbf{v}_h)_{\partial K}\right].\label{int-by-parts}
\end{eqnarray*}
By Lemma \ref{critical lemma for consistency}, we have
\begin{eqnarray*}
&&\sum_{K\in\mathcal{T}_h}\left[(\alpha \Delta (\nabla\times \mathbf{u})-\beta\nabla\times \mathbf{u},\mathbf{v}_h\times \mathbf{n})_{\partial K}\right]\\
&& \lesssim h(\|\Delta(\nabla\times \mathbf{u})\|_{0,\Omega}+\|\nabla\times \Delta(\nabla\times \mathbf{u})\|_{0,\Omega}+\|\nabla\times \mathbf{u}\|_{0,\Omega}+\|(\nabla\times)^2 \mathbf{u}\|_{0,\Omega}
)\\
&&\hskip 3mm \left(\sum_K\|\nabla\times \mathbf{v}_h\|_{1,K}^2\right)^{1/2}.
\end{eqnarray*}
By Lemma \ref{face_estimate_lemma} and the inter-element continuity of $\nabla\times \mathbf{v}_h$, we get
\begin{eqnarray*}
&&\sum_{K\in\mathcal{T}_h}\left[
\alpha(\nabla(\nabla\times \mathbf{u})\cdot \mathbf{n}, \nabla\times \mathbf{v}_h)_{\partial K}\right]\\
&&\leq \alpha\left| \sum_{K\in\mathcal{T}_h}\sum_{f\subset\partial K}(
\nabla(\nabla\times \mathbf{u})\cdot \mathbf{n}-P_f(\nabla(\nabla\times \mathbf{u})\cdot \mathbf{n}),\nabla\times \mathbf{v}_h-P_f(\nabla\times \mathbf{v}_h))_f\right|\\
&&\lesssim h|\nabla(\nabla\times \mathbf{u})|_{1,\Omega}\left(
|\sum_{K\in\mathcal{T}_h}|\nabla\times \mathbf{v}_h|_{1,K}^2\right)^{1/2}.
\end{eqnarray*}
The theorem follows by combining the above estimates of the two boundary integrals.
\end{proof}
Finally, we have the following convergence result.
\begin{theorem}
Let $\mathbf{u}$ and $\mathbf{u}_h$ be the solutions of the problems (3.10) and (3.16) respectively, then
$$
\norm{\mathbf{u}-\mathbf{u}_h}_{0,h}+\norm{\nabla\times
(\mathbf{u}-\mathbf{u}_h)}_{0,h}+\norm{\nabla(\nabla\times (\mathbf{u}-\mathbf{u}_h))}_{0,h}\lesssim
h \norm{\mathbf{u}}_{4,\Omega}
$$
when $\mathbf{u}\in (H^4(\Omega))^3$.\label{convergence-theorem}
\end{theorem}
\begin{proof}
Using the second Strang lemma,
\begin{eqnarray*}
&&\;\;\;\;\norm{\mathbf{u}-\mathbf{u}_h}_{0,h}+
\norm{\nabla\times
(\mathbf{u}-\mathbf{u}_h)}_{0,h}+
\norm{\nabla(\nabla\times (\mathbf{u}-\mathbf{u}_h))}_{0,h}\\
&&\lesssim \inf_{\mathbf{w}_h\in V_{h}}(\norm{\mathbf{u}-\mathbf{w}_h}_{0,h}+\norm{\nabla\times
(\mathbf{u}-\mathbf{w}_h)}_{0,h}+\norm{\nabla(\nabla\times (\mathbf{u}-\mathbf{w}_h))}_{0,h})
\\
&&\;\;\;\;+ \sup_{\mathbf{w}_h\in V_{h},\mathbf{w}_h\neq
0}\frac{a_h(\mathbf{u},\mathbf{w}_h)-(\mathbf{f},
\mathbf{w}_h)}{\norm{\nabla\times
\mathbf{w}_h}_{1,h}},
\end{eqnarray*}
and previous lemmas, the desired inequality follows.
\end{proof}
\section*{Acknowledgement}
The authors would like to thank Prof. Ludmil Zikatanov and Dr. Luis Chac\'{o}n for many helpful discussions.
\bibliographystyle{amsplain}
|
\section{\textbf{Introduction}}}
\vspace{.5cm} Any description of Quantum Gravity suggests a
smallest (but finite) length scale $l$ (or a finite upper bound of
energy $\kappa$), which of course should be observer independent.
The natural candidate for this is the Planck Length (or the Planck
Energy). But this proposition obviously contradicts the principles
of Einstein's Special Relativity (SR) Theory, as in SR, the length
or the mass (or energy) of an object varies for different
observers. Thus we need an extension of SR theory where along with
the velocity of light, another observer-independent quantity, a
fundamental length-scale exists. As a consequence, there must be
some modifications of SR theory in the high energy (Planck energy) regime.\\
As a possible solution, a new theory (DSR Theory) was first
proposed by Amelino-Camelia \cite{ac1}. Another model, perhaps
simpler, was given by Amelino-Camelia \cite{ac2} and by Magueijo
and Smolin \cite{mag1} (for discussion and review, see
\cite{acr,kowglik1} and references therein). As said earlier, in
these theories, there are two invariant quantities, $c$, the
velocity of light and $\ka$, an upper limit of energy. But for
consistent inclusion of this second invariant quantity along with
the other principles of SR theory, the well known dispersion
relation (or mass-shell condition) for a particle \be E^2 - p^2 =
m^2 \ee has to be modified as: \be E^2 - p^2 = m^2 \left(1 -
\frac{E}{\ka}\right)^2. \label{mdr} \ee Here $E$ and $p$ are
respectively the energy and the magnitude of the three-momentum of
the particle, $m$ is the mass of the particle and we have taken
$c=1$. We refer this model as the Magueijo-Smolin (MS) model.\\
In earlier work \cite{dasghosh}, we considered a particular
dispersion relation as in \cite{mag1}. Then we derived an
expression for the energy-momentum tensor for a perfect fluid and
studied dynamics of the perfect fluid with this modified
expression. Due to the presence of the invariant energy scale, our
derivation of the energy-momentum tensor was subtle where
nonlinear representation of Lorentz transformations played an
essential role. In this work, we adopt the same scheme and
consider (\ref{mdr}) as our fundamental equation. Then we go on to
study the thermodynamic properties of an ideal photon gas using
the methods of conventional statistical mechanics, but generalized
to be applicable in a theory where
an invariant energy scale is present.\\
We have arranged this paper as follows:~~In section 2, we discuss
about the modified dispersion relation. In the next section we
derive the expression for the density of states and the important
expression of the partition function. The derivation of the
expression for partition function is the most crucial result of
our work. In section 4, we go on to study the thermodynamic
properties of photon gas using this partition function. In
particular, we evaluated analytic expressions for the pressure,
equation of state, internal energy, entropy and specific heat of
the photon gas. We also show the comparisons between the
thermodynamic variables in the MS model and in the usual SR
scenario. Further, we see that the density of states as well as
the entropy decreases in the MS model as compared to that in the
SR framework. This happens due to the deformation in Lorentz
symmetry in the theory where an invariant energy scale is present.
It is another major result of our work. Finally, we conclude
summarizing our results and discuss some of the future prospects
in this regard.
\vspace{1cm} \n {\section{\textbf{Modified Dispersion Relation}}
\vspace{.5cm} We choose a particular modified dispersion relation
as given in \cite{mag1,dasghosh} $$ E^2 - p^2 = m^2 \left(1 -
\frac{E}{\ka}\right)^2. $$ Thermodynamic properties for photon gas
with a different dispersion relation has been studied in
\cite{camacho1}. Also, thermodynamics of bosons and fermions with
another modified dispersion relation and its cosmological and
astrophysical implications has been observed in \cite{magcos,
bertolami}. But these two modified dispersion relations appear
from a phenomenological point of view whereas the dispersion
relation (\ref{mdr}) has a more theoretical motivation which we
discuss below in some details. \\
It was shown in \cite{sghosh} that existence of an invariant
length scale in the theory is consistent with a non-commutative
(NC) phase space ($\ka$-Minkowski spacetime) such that the usual
canonical Poisson brackets between the phase space variables are
modified. Also, the linear Lorentz transformations (LT) are
replaced by non-linear $\ka$-Lorentz transformations ($\ka$-LT)
\cite{bruno, sghosh}. But still Lorentz algebra is intact in the
theory. As a result, we have the $\ka$-LT invariant modified
dispersion relation (\ref{mdr}) as: \be \{J_{\mu \nu} ,
\frac{p^2}{\left(1 - \frac{E}{\ka} \right)^2}\} = 0.
\label{invariant} \ee The angular momentum $J_{\mu \nu}$ is
defined as in \cite{sghosh} $$ J_{\mu \nu} = x_\mu p_\nu - x_\nu
p_\mu$$ where $x$ and $p$ are the phase space variables. Due to
the nontrivial expression for the dispersion relation (\ref{mdr}),
firstly it was supposed that the velocity of photon $ c = \frac{d
E}{d p} $ have to be energy dependent. But it was shown in
\cite{hoss} that a modified dispersion relation does not
necessarily imply a varying (energy dependent) velocity of light.
Thus, though the above two models (\cite{camacho1} and
\cite{magcos, bertolami}) admit a varying speed of light, in case
of MS model, for photons ($ m = 0$) the dispersion relation
(\ref{mdr}) is the same as in SR theory. Also the speed of light
$c$ is an invariant quantity in the MS model \cite{mag1, sghosh,
dasghosh}. Thus the MS model considered in \cite{mag1, sghosh} has
a more theoretical motivation and it can be developed starting
from the NC phase space variables \cite{sghosh} whereas the models
considered in \cite{camacho1, magcos, bertolami} are
phenomenological in nature and as far as we know, there is no
fundamental phase space structures to describe these models.\\
Another interesting fact is that both the models described in
\cite{camacho1} and in \cite{magcos, bertolami} have no finite
upper bound of energy of the photons though they have a momentum
upper bound. But, as stated earlier, in the MS case, though the
dispersion relation for the photons is unchanged, there is a
finite upper bound of photon energy which is the Planck energy
$\ka$. One can readily check that this is an invariant quantity by
using the $\ka$-Lorentz transformation law
for the energy \cite{bruno, sghosh}. \\
One more thing must be clarified here. In case of the models
(\cite{camacho1} and \cite{magcos, bertolami}), clearly the
Lorentz symmetry was broken and as a result, the number of
microstates and hence the entropy increases as compared to the
Lorentz symmetric SR theory. On the other hand, we are dealing
with a different scenario where the Lorentz symmetry is not broken
as Lorentz algebra between the phase space variables is intact. In
fact, the framework we describe here still satisfies the basic
postulates of Einstein's SR theory; moreover it possesses another
observer-independent quantity \cite{acr}. Thus it seems that
Lorentz symmetry is further restricted in this MS model. As a
result of this, we expect to have a less number of microstates and
less entropy in the Ms model. As we will show later in our
explicit calculations, this expected result is correct.\\
As we have said earlier, the modified dispersion relation
(\ref{mdr}) in case of the photons (massless particles) does not
change from the usual SR scenario. Thus, for the photons, the
dispersion relation now becomes \be p = E.
\label{photondispersion} \ee
\vspace{1cm} \n {\section{\textbf{Partition Function for Photon
Gas}}
\vspace{.5cm} To study the thermodynamic behavior of photon gas,
we have to find out an expression for the partition function
first, as it relates the microscopic properties with the
thermodynamic (macroscopic) behavior of a physical system
\cite{pathria, greiner}, which we do in this section.\\
\n {\subsection{Number of states:}} \vspace{.2cm}
We consider a box containing photon gas. Following the standard
procedure as given in \cite{pathria, greiner}, we consider a
continuous spectrum of momentum instead of quantizing it. The
number of microstates available to the system ($\sum$) in the
position range from $r$ to $r+dr$ and in the momentum range from
$p$ to $p+dp$ is given by \cite{pathria, greiner} \be \sum =
\frac{1}{h^3} \int \int d^3 \vec{r} d^3 \vec{p} \label{state1} \ee
where $h$ is the phase space volume of a single lattice and $$\int
\int d^3 \vec{r} d^3 \vec{p}$$ is the total phase space volume
available to the system. \\
It should be mentioned here that in case of SR theory, the
quantities $E d^3 x$ and $\frac{d^3 p}{E}$ are invariant under the
Lorentz transformations and hence the phase space volume element
$d^3 x d^3 p$ is a Lorentz invariant quantity \cite{misner}. The
nonlinear $\ka$-Lorentz transformations \cite{bruno, sghosh} are
explicitly given by: \be t' = \alpha \gamma (t - v x)~~,~~x' =
\alpha \gamma (x - v t)~~,~~y' = \alpha y~~,~~z' = \alpha z $$$$
E' = \frac{\gamma (E - v p_x)}{\alpha}~~,~~p'_x = \frac{\gamma
(p_x - v E)}{\alpha}~~,~~p'_y = \frac{p_y}{\alpha}~~,~~p'_z =
\frac{p_z}{\alpha}. \label{kalt} \ee The prime over a quantity
denotes the corresponding quantity in the boosted frame and
$\alpha = 1 + \frac{1}{\kappa}((\gamma - 1) E - v \gamma p_x)$. We
have considered the three-momentum to be of the usual form: $
\vec{p} = (v E, 0, 0)$ and $\gamma = \frac{1}{\sqrt{1-v^2}}$ where
$v$ is the velocity of the boosted frame. In case of our model,
the phase space volume element $d^3 x d^3 p$ is invariant under
the $\kappa$-Lorentz transformations (\ref{kalt}) as: \be d^3 x'
d^3 p' = \alpha^3 \gamma~d^3 x~ \frac{\gamma}{\alpha^3} \left(1 -
\frac{v p_x}{E}\right)d^3 p = d^3 x d^3 p. \label{invphsp} \ee It
is interesting to note that the factor $\alpha$ arising from the
nonlinear $\kappa$-Lorentz transformation finally cancels out in
(\ref{invphsp}). Strictly speaking, to derive (\ref{invphsp}) we
should consider the effect coming from variation of $\alpha$. But
we have omitted the term $d \alpha$ in (\ref{invphsp}) as it is a
dynamical effect and may not be relevant for the free particle
case as considered here.
If the volume of the box is considered to be $V$, the number of
microstates can be written in the following form using the
spherical polar coordinates \cite{pathria, greiner} \be \sum =
\frac{4 \pi V}{h^3} \int_0^{\infty} E^2 d E \label{state2}. \ee We
used the dispersion relation $ p=E$ to change the integration
variable to $E$. Then considering the fact that we have an finite
upper limit of energy ($\ka$), we obtain the number of
microstates: \be {\tilde{\sum}} = \frac{4 \pi V}{h^3} \int_0^{\ka}
E^2 d E \label{state}, \ee where $ \sim $ on a quantity represents
the corresponding quantity in the model we have considered. It is
obvious from the expressions (\ref{state2}) and (\ref{state}) that
the available number of microstates to the system is less than
that in the SR theory, as the energy spectrum of a particle in SR
theory can go all the way up till infinity.
This result agrees with our expectation stated earlier. \\
\n {\subsection{Partition function:}} \vspace{.2cm}
It is very crucial to get an expression for the partition function
as all the thermodynamic properties can be thoroughly studied
using the knowledge about the partition function. The single
particle partition function $Z_1(T, V)$ is defined as
\cite{greiner} \be Z_1(T, V) = \frac{4 \pi V}{h^3} \int_0^{\infty}
p^2 e^{-\beta E} d p, \label{partition1} \ee where
$\beta=\frac{1}{k_B T}$, $k_B$ is the
Boltzman constant and $T$ is the temperature of the particle.\\
For the MS model, the single particle partition function
$\tilde{Z}_1(T, V)$ is defined as \be \tilde{Z}_1(T, V) = \frac{4
\pi V}{h^3} \int_0^{\ka} p^2 e^{-\beta E} d p.
\label{partition1dsr} \ee In the limit $\ka \rightarrow \infty$,
we should get back normal SR theory results. \\
It should be noted that in the MS model which we have considered,
the photon dispersion relation is not modified at all. But still
there is modification in the partition function
(\ref{partition1dsr}) due to the presence of an energy upper bound
of particles ($\ka$) in the theory. So the upper limit of
integration is $\ka$ in (\ref{partition1dsr}) whereas in the
normal SR theory expression (\ref{partition1}), the upper limit of
integration is $\infty$ since there is no upper bound of energy in
the SR theory. In all the models \cite{camacho1, magcos,
bertolami}, though the upper limit of energy is infinity as in SR
theory, these models differ
due to the different dispersion relations. \\
Using the dispersion relation for photons ($E = p$) and using the
standard table and formulae for integrals \cite{gradstein}, we
finally have an analytical expression of the single particle
partition function \be \tilde{Z}_1(T, V) = \frac{4 \pi V}{h^3}
\int_0^{\ka} E^2 e^{-\beta E} d E = \frac{4 \pi V}{h^3}
\left[\frac{2}{\beta^3} - \frac{e^{-\beta \ka}}{\beta^3}(2+\beta
\ka (2+\beta \ka))\right]. \label{part1} \ee Thus the partition
function for a $N$-particle system $\tilde{Z}_N(T, V)$ is given by
\be \tilde{Z}_N(T, V) = \frac{1}{N!} [\tilde{Z}_1(T, V)]^N
\label{partn} \ee where we have considered classical
(Maxwell-Boltzman) statistics along with the Gibb's factor. As we
get the expression for the partition function, now we go on to
study various thermodynamic properties of the photon gas in our
model. It should be noted that as $\ka \rightarrow \infty$, this
partition function coincides with the partition function in SR
theory and thus all of our results coincides with the usual SR
case in this limit. \vspace{1cm}
\n {\section{\textbf{Thermodynamic Properties of Photon Gas}}}
\vspace{.5cm}
With the expression for the partition function in our hand, now we
go on to study various thermodynamic properties of photon gas in a
theory where an observer-independent fundamental energy scale is present.\\
\n {\subsection{Free energy:}} \vspace{.2cm}
We use Stirling's formula for $ ln [N!]$ \cite{greiner} $$ ln [N!]
\approx N~ln[N] - N$$ in the expression for partition function
(\ref{partn}) to obtain the free energy $\tilde{F}$ of the system
\be \tilde{F} = -k_B T~ln [\tilde{Z}_N(T, V)] $$$$ = - N k_B T
\left[ 1 + ln \left[\frac{4 \pi V}{N}\left(\frac{k_B
T}{h}\right)^3 \left\{2 - e^{- \frac{\ka}{k_B T}}\left(2 +
\frac{\ka}{k_B T}\left(2 + \frac{\ka}{k_B
T}\right)\right)\right\}\right] \right]. \label{free} \ee In the
limit $\ka \rightarrow \infty$, the terms containing $\ka$
vanishes and we get back normal SR theory result:
$$ F = N k_B T. $$ \vspace{.2cm}
\n {\subsection{Pressure:}} \vspace{.2cm}
From the expression for free energy (\ref{free}) we can readily
obtain the pressure $\tilde{P}$ of photon gas in our considered
model as \cite{greiner} \be \tilde{P} = - \left(\frac{\partial
F}{\partial V}\right)_{T, N} = \frac{N k_B T}{V}. \label{pressure}
\ee Thus, we have the same equation of state
$$ P V = N k_B T$$ as in SR theory. \vspace{.2cm}
\n {\subsection{Entropy:}} \vspace{.2cm}
As we have the expression for free energy (\ref{free}), also we
can evaluate the entropy $\tilde{S}$ of the system from the
following relation \cite{greiner} $$ \tilde{S} =
-\left(\frac{\partial F}{\partial T}\right)_{V, N}. $$ The
expression for entropy takes the following form \be \tilde{S} = N
k \left[4 + ln \left[\frac{4 \pi V}{N}\left(\frac{k_B
T}{h}\right)^3 \left\{2 - e^{- \frac{\ka}{k_B T}}\left(2 +
\frac{\ka}{k_B T}\left(2 + \frac{\ka}{k_B T}\right) \right)
\right\}\right] \right. $$$$ \left. - \frac{\ka^3}{2 k_B^3 T^3
e^{\frac{\ka}{k_B T}} - \left(2 k_B^3 T^3 + 2 k_B^2 T^2 \ka + k_B
T \ka^2\right)}\right]. \label{entropy} \ee The terms containing
$\ka$ are the modifications from the SR theory expression of
entropy \cite{greiner}. As in the earlier expressions, in the
limit $\ka \rightarrow \infty$ the terms containing $\ka$ vanish
and we get back the SR theory result: \be S = N k_B \left[4 + ln
\left[\frac{8 \pi V}{N}\left(\frac{k_B T}{h}\right)^3 \right]
\right]. \ee We plot the entropy $S$ against $T$ both for the
model we considered and for SR theory to study the deviation of
entropy in the two models.
\begin{figure}[htb]
{\centerline{\includegraphics[width=9cm, height=5cm] {st.eps}}}
\caption{{\it{Plot of entropy of photon $S$ against temperature
$T$ for both in the SR theory and in our case; the dashed line
corresponds to the SR theory result and the thick line represents
the corresponding quantity in our result. We have used the Planck
units and the corresponding parameters take the following values $
\ka = 10000, k_B = 1, N = 10000, V = .01, h = 1 $ in this plot as
well as in all other plots in the paper. In this scale, $T=10000$
is the Planck temperature.}}} \label{fig1}
\end{figure}
In Figure 1, we have plotted entropy against temperature for both
the case of our invariant energy scale scenario and normal SR
theory. It is clearly observable from the plot that the entropy
grows at a much slower rate in case of our result than in the SR
theory and as temperature increases, the entropy in our considered
model deviates more from the entropy in the SR theory. This result
matches with our earlier expectation considering the underlying
symmetry of the theory that the entropy in the MS model should be
less than the entropy in SR theory.\\
It is well known that the total number of microstates available to
a system is a direct measure of the entropy for that system.
Therefore our result merely reflects the fact that due to the
existence of an energy upper bound $\ka$, the number of
microstates gradually saturates to some finite value near Planck
scale. \vspace{.2cm}
\n {\subsection{Internal energy:}} \vspace{.2cm}
We expect modification in the expression of the internal energy
$U$ for photon gas in the MS model as the expression of entropy is
modified and internal energy is related to the entropy as follows:
$$ U = F + T S. $$ In the usual SR scenario, the explicit
expression for internal energy is given by \be U = 3 N k_B T.
\label{internalsr} \ee But in the MS scenario we considered, the
expression for internal energy ($\tilde{U}$) of photon gas takes
the following form \be \tilde{U} = N k_B T \left[3 - \frac{ \ka^3
e^{- \frac{\ka}{k_B T}}}{2 k_B^3 T^3 - e^{- \frac{\ka}{k_B T}}(2
k_B^3 T^3 + 2 \ka k_B^2 T^2 + \ka^2 k_B T)}\right].
\label{internaldsr} \ee It is easy to see from the expression of
internal energy (\ref{internaldsr}) that we get back the usual SR
theory expression in the limit $\ka \rightarrow \infty$. As in the
case of entropy, here we also plot internal energy against
temperature for both the SR and MS case.
\begin{figure}[htb]
{\centerline{\includegraphics[width=9cm, height=5cm] {ut.eps}}}
\caption{{\it{Plot of internal energy of photon $U$ against temperature
$T$ for both in the SR theory and MS scenario; the dashed line
corresponds to the SR theory result and the thick line represents
the quantity in the MS model we considered here.}}} \label{fig2}
\end{figure}
In Figure 2, we plotted internal energy of photon gas against its
temperature for both the case of MS model and SR theory. The
expression in SR theory (\ref{internalsr}) tells us that internal
energy depends linearly on the temperature and this is supported
from the plot. But from the expression of internal energy in the
MS model (\ref{internaldsr}) it is clearly observed that the
relation of internal energy with temperature is not linear at all.
Also one can easily check that the value of internal energy (for a
particular temperature) in the MS model (\ref{internaldsr}) is
always less than its value (for the same temperature) in the SR
theory (\ref{internalsr}). This is very clear from the plot as the
curve for the MS model always lies below the straight line which
corresponds the SR theory result. \\
Since the internal energy $U$ of photon gas becomes saturated
after a certain temperature in case of the MS model, it is
tempting to point out that probably our results are moving towards
the right direction related to the ``Soccer Ball Problem" that
plagues multi-particle description in the framework of DSR. The
problem lies in the fact that if we apply linear addition rule for
momenta/energies of many sub-Planck energy particles we may end up
with a multi-particle state, such as a soccer ball, whose total
energy is greater than the Planck energy which is forbidden in the
DSR theory. For further discussion about the ``Soccer Ball
Problem" see \cite{maghoss}. \vspace{.2cm}
\n {\subsection{Pressure-energy density relation:}} \vspace{.2cm}
Though internal energy $U$ of a physical system is not directly
measurable, still we can detect the effect of it through other
thermodynamic quantities, such as the relation between pressure
$P$ and energy density $\rho$ of that system. Energy density of a
system $\rho$ is defined as $$ \rho = U/V $$ where $U$ is the
internal energy of the system and $V$ is the volume occupied by
the system. As the expression for internal energy $U$ is modified
in the MS model, we also expect modifications in the expression
for the energy density $\rho$. The modified relation between
pressure $\tilde{P}$ and energy density $\tilde{\rho}$ is given
by: \be \tilde{P} = \frac{1}{3} \tilde{\rho} + \frac{1}{3} \frac{
N k_B T \ka^3 e^{-\frac{\ka}{k_B T}}}{2 V k_B^3 T^3 - V
e^{-\frac{\ka}{k_B T}}(2 k_B^3 T^3 + 2 \ka k_B^2 T^2 + \ka^2 k_B
T)}. \label{prenergy} \ee For $\ka \rightarrow \infty$, we get
back the usual pressure-energy density relation in SR theory: $$ P
= \frac{1}{3} \rho. $$ It should be pointed out that in our
earlier work \cite{dasghosh}, it was shown that in the
ultra-relativistic regime (for photons), the relation $P =
\frac{1}{3} \rho$ remains unaffected. But here we have a
modification in this pressure-energy density relation
(\ref{prenergy}). In \cite{dasghosh}, we obtained the result
considering some simplified assumptions. But in this work, we
start with the partition function and apply the methods of
Statistical Mechanics (which naturally deals with multi-particle
systems). So we do not really have to consider any strong
assumptions here. \vspace{.2cm}
\n {\subsection{Specific heat:}} \vspace{.2cm}
There is another thermodynamic parameter, specific heat ($C_V$),
through which we can observe the modifications in the expression
for internal energy. Specific heat $C_V$ is defined as $$ C_V =
\left(\frac{\partial U}{\partial T}\right)_V. $$ For the MS model
we considered here, explicit calculation yields the following
result \be \tilde{C}_V = 3 N k_B - \frac{2 N k_B \ka^3 \left(k_B^2
T^2 \ka(1 + e^{\frac{\ka}{k_B T}}) + 2 k_B^3 T^3 (1 -
e^{\frac{\ka}{k_B T}})\right)}{\left(2 k_B^3 T^3
(e^{\frac{\ka}{k_B T}} - 1)-2 k_B^2 T^2 \ka - k_B T \ka^2
\right)^2}. \label{spheat} \ee After doing a bit of algebra, one
can check from the above expression (\ref{spheat}) that specific
heat calculated from the MS model is always less than the value
calculated from usual SR theory. Also, when $ \ka \rightarrow
\infty$, we obtain $$ C_V = 3 N k_B $$ which is the usual specific
heat for photon as calculated in SR theory.
\begin{figure}[htb]
{\centerline{\includegraphics[width=9cm, height=5cm] {c_vt.eps}}}
\caption{{\it{Plot of specific heat of photon $C_V$ against temperature
$T$ for both in the SR theory and MS scenario; the dashed line
corresponds to the SR theory result and the thick line represents
the quantity in the MS model we considered here.}}} \label{fig3}
\end{figure}
In Figure 3, we have plotted the specific heat $C_V$ against
temperature $T$. It is clear from the plot that in the case of the
MS model, the specific heat $C_V$ asymptotically decreases to zero
suggesting that photon gas has reached its temperature ceiling
which is the Planck temperature.
\n {\section{\textbf{Conclusion and Future Prospects}}}
\vspace{.2cm}
We consider the modified dispersion relation as given in the MS
model \cite{mag1}. We explicitly show that for photons, the number
of microstates available to a macrostate is less in the MS model
than in the usual SR scenario. We stress that it happens since
Lorentz symmetry is not broken in this model. But due do the
presence of an invariant energy upper bound in this theory,
microstates can avail energies only up to a finite cut-off whereas
in SR theory, microstates can attain energies up to infinity.
Thus, quite naturally, the number of microstates in this MS model
is less than that in SR theory. \\
The most significant result of our work is the derivation of
$N$-particle partition function in the MS model. Due to the
presence of the deformed dispersion relation, this task becomes
highly non-trivial. However, for photons, we find out an analytic
expression for the partition function. Once we have the partition
function in our hand, we evaluate other various thermodynamic
parameters of photon gas such as the free energy, pressure,
entropy, internal energy, specific heat for the MS model and
compare them with the known results of SR theory. \\
As a consequence of deformed Lorentz symmetry, the entropy in the
MS model is also less than that in the SR scenario. We show this
behavior analytically and graphically. Also the internal energy is
modified in case of the MS model and as a consequence the
expression for the specific heat is also modified.\\
Though highly non-trivial, one can similarly study the behavior of
an ideal gas using this modified dispersion relation. Also one can
study behavior of fermion gas in this MS model. There might be
some modifications in the Fermi energy level which can modify the
Chandrasekhar mass limit for the white dwarf stars
\cite{bertolami}. Thus astrophysical phenomena in an MS
framework is another issue remains to be addressed.\\
Further, as we have the expression for energy-momentum tensor, one
can study the cosmological aspects of MS model using the Friedmann
equations. But this requires idea about the geometry sector
(precisely the metric $g_{\mu \nu}$ and hence Einstein tensor
$G_{\mu \nu}$) which is till unknown in the context of MS model.
This still remains another open issue to be further studied.
It is noteworthy to mention here that ``bouncing" loop quantum
cosmology theories (for example see \cite{bojowald} and references
therein) entails some modifications to the geometry of spacetime
which in turn effectively puts a bound on the curvature avoiding
the big bang singularity. However, for these ``bouncing" models,
the perturbation technique cannot be done as at the point of
curvature saturation, the energy density of the cosmic fluid
diverges. So it is unclear how to construct the matter part of
Einstein equation. One alternative to avoid the big bang
singularity is the inflation theory where the perturbation method
can also be applied. On the other hand, in our model, the energy
density of the cosmic fluid saturates to the Planck energy which
is a finite real quantity. Possibly a combination of the model
considered in this paper along with the ``bouncing" loop quantum
cosmology can successfully describe a situation where big bang
singularity can be avoided. \vspace{.5cm}
\n {\textbf{Acknowledgements}} \vspace{.2cm}
We would like to thank Prof. Subir Ghosh for helpful discussions.
|
\section{\label{sec1}Introduction}
New and precise measurements of the cross section for electron-positron
annihilation into hadrons have been performed during the past years
which were based on the method of "Radiative Return"
\cite{Zerwas,Binner:1999bt}. Exclusive reactions, specifically two-body final states like
$\pi^+\pi^-$ \cite{:2008en,:2009fg},
$p\bar p$ \cite{Aubert:2005cb} or $\Lambda\bar \Lambda$ \cite{Aubert:2007uf}
and three- \cite{Aubert:2004kj,Aubert:2007ym}
and four-meson final states \cite{Aubert:2005eg,Aubert:2007ur}
have been explored.
An important ingredient in these analyzes was and is the simulation of
all these reactions through a Monte Carlo generator. In a first step, the
generator EVA was developed \cite{Binner:1999bt,Czyz:2000wh},
which is based on leading
order matrix elements combined with structure function methods for an
improved treatment of initial state radiation.
Subsequently the complete next-to-leading
order (NLO) QED corrections were evaluated
\cite{Rodrigo:2001jr,Kuhn:2002xg}
and
implemented into the generator PHOKHARA
\cite{Rodrigo:2001kf,Czyz:2002np,Czyz:PH03,Nowak,Czyz:PH04,Czyz:2004nq,Czyz:2005as,Czyz:2007wi,Czyz:2008kw},
which is now
available for a variety of exclusive final states.
(For a recent review of theoretical and experimental results
see e.g. \cite{Actis:2009gg}.)
B-meson factories,
operating at energies around 10~GeV and with high luminosity, allow to
explore hadronic final states with relatively large invariant masses, up
to 3~GeV and beyond. Therefore, the narrow resonances $J/\psi$ and
$\Psi(2S)$ can be studied through the radiative return, in particular in
decay channels of low multiplicity, leptonic ones like $\mu^+\mu^-$
\cite{Aubert:2003sv},
or
two-body hadronic modes like $\pi^+\pi^-$, $K^+K^-$, $K^0 \bar K^0$ or
$p\bar p$ \cite{Aubert:2005cb}. The signal is identified with the help of a very good mass
resolution and particle identification in the resonance region.
For an analysis exploiting the large statistics, the inclusion of
radiative corrections from initial- and final-state radiation (ISR and
FSR) is mandatory, since it affects the cross section and the line
shape of the resonance. For the simulation of hadronic final states both
the electromagnetic contribution, i.e. a parameterization of the form
factor, and the strength of the direct coupling of the resonance to the
hadrons are required. The latter is absent for final states with
positive G-parity ($2\pi$, $4\pi$, ... ) but non-vanishing e.g. for
$K\bar K$, $3\pi$ or final states with baryons. On the other hand, a careful
analysis of the resonance line shape in the various channels would allow
a model-independent determination of the direct coupling and of the form
factors close to resonance \cite{Czyz:2009vj,Seth:Jphi,Yuan:2003hj,Rosner:1999zm,Suzuki:1999nb,LopezCastro:1994xw,Milana:1993wk}.
With this motivation in mind we reanalyze the pion and kaon form factors
with emphasis on the region above the $\rho$-resonance. The basic
ingredients are very similar to those employed in an earlier study
\cite{Bruch}. However, additional assumptions are required to properly describe
the different resonance-like structures in the energy region between
1~GeV and 3~GeV. The details of this model and its parameters are
described in sections II and III for pions and kaons, respectively. The
new implementation of these modes into PHOKHARA, which includes, as
before, NLO ISR and FSR, is presented in Section IV. Section V is
concerned with the implementation of the narrow resonances in the
channels $\mu^+\mu^-$, $\pi^+\pi^-$, $K^+K^-$ and $K^0 \bar K^0$.
Hadronically and electromagnetically induced amplitudes are included,
together with the radiative corrections from ISR and FSR. Section VI
contains a brief summary and our conclusions.
\section{\label{pionff} The pion form factor}
For a realistic
generation a model for the electromagnetic form factor is required.
The ansatz presented in \cite{Bruch} was published before
the CLEO-c measurement
of the form factor in the vicinity of the $\psi(2S)$ resonance
\cite{Pedlar:2005sj} and underestimates the experimental result
significantly.
The same applies to the model predictions at $J/\psi$
as compared to the pion form factor calculated in \cite{Milana:1993wk}
from $B(J/\psi\to\pi^+\pi^-)$ and $B(J/\psi\to e^+e^-)$ decay rates.
\begin{figure}[h]
\vspace{0.5 cm}
\begin{center}
\includegraphics[width=8.cm,height=8.cm]{v_plus_pik.ps}
\includegraphics[width=8.cm,height=8.cm]{v_plus_4.ps}
\caption{(color online)
The experimental data
\cite{Bisello:1988hq,:2009fg,Akhmetshin:2003zn,Akhmetshin:2006bx,Akhmetshin:2006wh,Achasov:2005rg,
Aulchenko:2006na,:2008en,Pedlar:2005sj}
compared to the model fits results (see text for details).
The form factor at $J/\psi$ comes from its theoretical extraction
\cite{Milana:1993wk,Amsler:2008zz} from the data.\label{pipi}
}
\end{center}
\vspace{0.5 cm}
\end{figure}
To accommodate the new data,
the updated
model ansatz for the pion form factor is taken
similarly to \cite{Bruch}
\bea
F_{\pi}(s) &=& \left[\sum\limits_{n=0}^N c_{\rho_n}^\pi
BW_{\rho_n}(s)\right]_{fit} \nonumber \\
&&+ \left[\sum\limits_{n=(N+1)}^{\infty}c_{\rho_n}^\pi BW_{\rho_n}(s)
\right]_{dQCD} \ ,
\label{piformf}
\end{eqnarray}
however, with different set of parameters.
Those of the first $N+1$
$\rho$ radial excitations are fitted and the rest is taken from the
``dual QCD model'' \cite{Dominguez:2001zu}.
It is necessary to take $N=5$ to
fit the data. For the precise treatment of $\rho_4$ and $\rho_5$ see below.
For the Breit-Wigner function we adopt
the Gounaris-Sakurai \cite{Gounaris:1968mw} version
with pion loop corrections included:
\bea
BW_{\rho_n}(s)=
\frac{m_{\rho_n}^2+ H(0)}{m_{\rho_n}^2-s +H(s)
-i\sqrt{s} \ \Gamma_{\rho_n}(s)} \ ,
\label{GS1}
\end{eqnarray}
where
\bea
H(s)= \hat{H}(s)-\hat{H}(m_{\rho_n}^2)-(s-m_{\rho_n}^2)
\frac{d}{ds}\hat{H}(m_{\rho_n}^2)\,,\label{hh}
\end{eqnarray}
\bea
\hat{H}(s)=\left(\frac{m_{\rho_n}^2\Gamma_{\rho_n}}{2\pi[p(m_{\rho_n})]^3}\right) \left(\frac{s}4-m_\pi^2\right)\nonumber \\
\times v(s)\log\frac{1+v(s)}{1-v(s)}\,,
\label{hhhat}
\end{eqnarray}
\bea
p(s) =\frac12 (s-4m_\pi^2)^{1/2} \,, \ \
v(s) = \sqrt{ 1 - \frac{4 m_{\pi}^2} {s} }\, .
\end{eqnarray}
Correspondingly we use the $s$- dependent widths
\bea
&&\Gamma_{\rho_n}(s) = \frac{ m^2_{\rho_n}}{s}
\biggl(\frac{p(s)}{p(m^2_{\rho_n})}\biggr)^{3}
\ \Gamma_{\rho_n} \ \theta({s-4m_\pi^2}).
\label{gamrho}
\end{eqnarray}
which are taken from two--body $P$- wave final states and for simplicity
(and lack of experimental information)
also used for the rest of decay channels
\cite{Bruch}. In Eqs. (\ref{hhhat}) and (\ref{gamrho}) we have used
$\Gamma_{\rho_n}\equiv\Gamma_{\rho_n}(s=m_{\rho_n}^2)$, which is
the total width of the $\rho_n $ meson.
The constraint $\sum_{n=0}^{n=\infty} c_{\rho_n}^\pi =1$ together with
$BW_{\rho_n}(0)=1$ enforces the proper normalization of the form factor
$F_\pi(0)=1$.
For the ground state $\rho(770)$ isospin violation
from $\rho-\omega$ mixing
is taken into account by substituting
\bea
c_{\rho_0}^\pi BW_{\rho_0}(s)
\to \frac{c_{\rho_0}^\pi BW_{\rho_0}(s)}
{1+c_{\omega}^\pi}(1+c_{\omega}^\pi BW_{\omega} )\, .
\label{ffrho0}
\end{eqnarray}
A Breit-Wigner function with constant width
\bea
BW_{\omega} = \frac{ m^2_{\omega} }{m^2_{\omega} - s
- i m_{\omega} \Gamma_{\omega}} \ ,
\end{eqnarray}
is used for description of the $\omega$ resonance.
As discussed in the Introduction,
the couplings $c_{\rho_{n}}^{\pi}$ are based on the ansatz predicted in the
dual-$QCD_{N_c=\infty}$ model
\cite{Dominguez:2001zu}
\bea
c_{\rho_{n}}^{\pi} = \frac{ (-1)^n \Gamma(\beta-1/2)}
{\alpha' m^2_{\rho_n} \sqrt{\pi}\Gamma(n+1)\Gamma(\beta-1-n)}\ ,
\label{couplrho}
\end{eqnarray}
where $\alpha'=1/(2m_{\rho_0}^2)$ is the slope of the Regge trajectory
$\alpha_\rho(s)= 1+\alpha'(s- m_{\rho_0}^2)$. The model postulates an
equidistant mass spectrum $ m^2_{\rho_n} = m^2_{\rho_0} \left(1+2n\right)$
and a linear relation between mass and width of a given resonance
$\Gamma_{\rho_n} = \gamma m_{\rho_n}$, with $\gamma$ derived from the
lowest resonance.
The parameters $\beta$ and $m_{\rho_0}$
are to be taken from the fit.
We fit the data in the time-like region which provides detailed
information about the structure of the resonances and coincides
with the region relevant for the
PHOKHARA Monte Carlo generator.
We have used new data
\cite{Akhmetshin:2003zn,:2009fg,Akhmetshin:2006bx,Akhmetshin:2006wh,Achasov:2005rg,
Aulchenko:2006na,:2008en,Pedlar:2005sj},
whenever possible. They are more accurate and the treatment of
radiative corrections is well documented.
Furthermore, we adopt the theoretical extraction
of the pion form factor at $J/ \psi$ using
\cite{Milana:1993wk}
\bea
|F_\pi|^2 = \frac{4 B(J/ \psi\to \pi^+\pi^-)}{\beta^3_\pi B(J/ \psi\to e^+e^-)}
\ ,
\label{ffpsi}
\end{eqnarray}
$\left(\beta_\pi = \sqrt{1-4m^2_\pi/M^2_{J/ \psi}} \ \right)$
and recent experimental data \cite{Amsler:2008zz}.
If one would assume independent point to point
statistical and systematic
errors of the new data
\cite{Akhmetshin:2003zn,Akhmetshin:2006bx,Akhmetshin:2006wh,Achasov:2005rg,
Aulchenko:2006na,:2008en} and combine these in quadrature
the results would be inconsistent and no fit could be made.
Summing
linearly the statistical and systematic experimental errors
for each experimental data point
one finds very good agreement between the experimental data. This
approach will be
adopted below. The new BaBar data \cite{:2009fg} become available
only after our
analysis was finished and we include here only their part (above 1.2
GeV). The BaBar data below 1.2 GeV are in conflict with KLOE data
and further investigations would be required how to merge these
conflicting data samples.
In \cite{Akhmetshin:2003zn,Akhmetshin:2006bx,Akhmetshin:2006wh,Achasov:2005rg,Aulchenko:2006na,:2008en}
the form factor including vacuum polarization was measured. We prefer
to parameterize
the 'bare' form factor $F_\pi$
(see \cite{Czyz:2009vj} for definition),
which is used
throughout this paper and for example directly obtained in
Eq.(\ref{ffpsi}).
The vacuum polarization corrections
are taken from \cite{Jeg_web,Czyz:2005as}.
For the extraction of the form factor from the cross section,
the CLEO-c collaboration \cite{Pedlar:2005sj}
has corrected for the leptonic part of the vacuum polarization effects.
Hence their result has still to be corrected only for the hadronic part,
which corresponds to a 1.5\% shift of $|F_\pi|^2$ only and
is irrelevant at the present experimental precision.
\begin{table}[ht]
\begin{center}
\vskip0.3cm
\begin{tabular}{|c|c|c|c|}
\hline
Parameter & model(fit) &
PDG value & model \\
&& \cite{Amsler:2008zz}&\\
\hline
$m_{\rho_0}$ & 773.37\,$\pm$\,0.19&
775.49\,$\pm$\,0.34 & input\\
$\Gamma_{\rho_0}$ & 147.1 \,$\pm$\,1.0 &
149.4\,$\pm$\,1.0 & input\\
\hline
$m_{\omega}$ & 782.4\,$\pm$\,0.5 &
782.41\,$\pm$\,0.12 & - \\
$\Gamma_{\omega}$ & 8.33\,$\pm$\,0.27
& 8.49\,$\pm$\,0.08 & - \\
\hline
$m_{\rho_1}$ & 1490\,$\pm$\,11 &
1465\,$\pm$\,25 & 1340 \\
$\Gamma_{\rho_1}$ & 429\,$\pm$\,27 & 400\,$\pm$\,60
& 256\\
\hline
$m_{\rho_2}$ & 1870$\,\pm$\,25 &
1720$\,\pm$\,20 & 1730 \\
$\Gamma_{\rho_2}$ & 357\,$\pm$\,46 & 250\,$\pm$\,100
& 330 \\
\hline
$m_{\rho_3}$ & 2120\,\cite{Czyz:2008kw} & - & 2047\\
$\Gamma_{\rho_3}$ & 300\,\cite{Czyz:2008kw} & - & 391\\
\hline
$m_{\rho_4}$ & model & -& 2321\\
$\Gamma_{\rho_4}$ & model & -& 444\\
\hline
$m_{\rho_5}$ & model & -& 2567\\
$\Gamma_{\rho_5}$ & model & -& 491\\
\hline
\hline
$\beta$&
2.148$\pm$0.003 & - & input\\
\hline
$|c_\omega^\pi|$
& (18.7$\pm$0.5)$\cdot$10$^{-4}$ & -&-\\
$Arg(c_\omega^\pi)$
& 0.106\,$\pm$\,0.020 & -&-\\
\hline
$|F_2|$ & 0.59\,$\pm$\,0.10 &-&
- \\
$Arg(F_2)$ & -2.20\,$\pm$\,0.16 &- &
- \\
$|F_3|$ & 0.048\,$\pm$\,0.056 &-&
- \\
$Arg(F_3)$ & -2.\,$\pm$\,1.4 &- &
- \\
$|F_4|$ & 0.40\,$\pm$\,0.07 &-&
- \\
$Arg(F_4)$ & -2.9\,$\pm$\,0.3 &- &
- \\
$|F_5|$ & 0.43\,$\pm$\,0.05 &-&
- \\
$Arg(F_5)$ & 1.19\,$\pm$\,0.18 &- &
- \\
\hline
$\chi^2/d.o.f.$ & 271/270 &-&-\\
\hline
\end{tabular}
\caption{{\it Parameters of the pion form factor (Eq.(\ref{piformf})
and Eq.(\ref{newcoef}))
and results of the fit to the data.}}
\label{tab:pion}
\end{center}
\end{table}
\begin{figure}[h]
\vspace{0.5 cm}
\begin{center}
\includegraphics[width=8.cm,height=8.cm]{tailvsnotail.ps}
\caption{(color online)
The relative difference between the form modulus square of the pion
form factor and the form factor calculated with first
six ($N=0,...,5$) resonances.\label{pipitail}
}
\end{center}
\vspace{0.5 cm}
\end{figure}
We have attempted to fit the experimental data keeping the coupling
constants $c_{\rho_{n}}^{\pi}$ fixed to the model values
(one fit parameter $\beta$ for all of them) and fitting only
the masses and the widths of the first few resonances (up to $n=5$).
This parameterization is satisfactory
up to $\sqrt{s}\sim 1.3-1.4$ GeV, where the details
of the model for resonances, with $n=2$ and higher, are not important.
However, the model is
definitively too simple for a description of the details of
higher radial excitations
including issues like
coupled channels in decays of the higher radial $\rho$ excitations.
Hence we adopt a heuristic approach, where we allow
for arbitrary complex couplings $f_n$ of the $\rho_n$ ($n=1,2,3,4,5$)
\bea
f_n = F_n \left(\sum_{i=1}^5c_{\rho_{i}}^{\pi}\right) / \left(
{\sum_{i=1}^5F_i} \right)\, ,
\label{newcoef}
\end{eqnarray}
with $F_1=1$ and four complex constants $F_n \ , \ n=2,3,4,5$
fitted to the experimental data.
The mass and the width of $\rho_3$ are fixed to
their values obtained in the fit
to the four pion production data \cite{Czyz:2008kw}.
For the masses and widths of the higher excitations ($n\geq4$)
we use their model values.
The results are shown in Table \ref{tab:pion}.
The fitted value of $m_{\rho_0}$
is smaller than its
PDG2008 \cite{Amsler:2008zz} value,
a consequence of using the dressed form factor in
\cite{Akhmetshin:2003zn,Akhmetshin:2006bx,Akhmetshin:2006wh,Achasov:2005rg,
Aulchenko:2006na}. This phenomenon was also observed
in \cite{Ghozzi:2003yn}.
The parameters describing the radial $\rho$ excitations
obtained in the fit have to be taken with great care as they are
strongly correlated, while in Table \ref{tab:pion} we give only MINOS
(MINUIT procedure from CERNLIB) parabolic errors.
To ilustrate the numerical importance of the higher radial excitations within
the ``dual QCD model'' in Fig. \ref{pipitail} we show the relative
difference between the full modulus square of the pion form factor and
the result calculated with the first six resonances which were used in
the fit. It is evident that it is impossible to neglect the higher
resonances and even in the $\rho_0$ region they give small, but not
negligible contribution to the form factor.
\section{\label{sec2} The kaon form factor}
The kaon form factors were revisited for the same reasons as the pion
form factor. Compared to the CLEO-c result \cite{Pedlar:2005sj}
the model
presented in \cite{Bruch} underestimates the kaon form factor
in the vicinity of the $\psi(2S)$ resonance.
It is impossible to fit the existing data,
including the CLEO-c result,
with the functional form used in \cite{Bruch} or
adding one or two more radial excitations, unless one would accept
inclusion of a huge wide resonance in the region between $J/\psi$
and $\psi(2S)$. To cure the situation, a model analogous to the
one used for the pion form factor, assuming an infinite tower of resonances,
was adopted. The ansatz reads
\begin{widetext}
\bea
F_{K^+}(s)=\frac{1}{2} \biggl( \left[\sum\limits_{n=0}^{N_\rho} c_{\rho_n}^K
BW_{\rho_n}(s) \right]_{fit}
+ \left[\sum\limits_{n=N_\rho+1}^{\infty}c_{\rho_n}^K BW_{\rho_n}(s)
\right]_{dQCD}
\biggr) \nonumber \\
+ \frac{1}{6}
\biggl( \left[\sum\limits_{n=0}^{N_\omega} c_{\omega_n}^K
BW_{\omega_n}^{c}(s)\right]_{fit}
+ \left[\sum\limits_{n=N_\omega+1}^{\infty}c_{\omega_n}^K BW^c_{\omega_n}(s)
\right]_{dQCD}
\biggr)\nonumber \\
+\frac{1}{3} \biggl( \left[\sum\limits_{n=0}^{N_\phi} c_{\phi_n}^K
BW_{\phi_n}^{K}(s) \right]_{fit}
+ \left[\sum\limits_{n=N_\phi+1}^{\infty}c_{\phi_n}^K BW^{K}_{\phi_n}(s)
\right]_{dQCD}
\biggr) \,,
\label{Kformfpm1}
\end{eqnarray}
\bea
F_{K^0}(s)= -\frac{1}{2}\biggl( \left[\sum\limits_{n=0}^{N_\rho} c_{\rho_n}^K
BW_{\rho_n}(s) \right]_{fit}
+ \left[\sum\limits_{n=N_\rho+1}^{\infty}c_{\rho_n}^K BW_{\rho_n}(s)
\right]_{dQCD} \biggr)
\nonumber\\
+ \frac{1}{6}
\biggl( \left[\sum\limits_{n=0}^{N_\omega} c_{\omega_n}^K
BW_{\omega_n}^{c}(s)\right]_{fit}
+ \left[\sum\limits_{n=N_\omega+1}^{\infty}c_{\omega_n}^K BW^c_{\omega_n}(s)
\right]_{dQCD}
\biggr)
\nonumber \\
+\frac{1}{3}\biggl( \left[ \eta_\phi c_{\phi_0}^K BW_{\phi_0}^{K}(s)
+ \sum\limits_{n=1}^{N_\phi}c_{\phi_n}^K BW^K_{\phi_n}(s)\right]_{fit}
+ \left[\sum\limits_{n=N_\phi+1}^{\infty}c_{\phi_n}^K BW^K_{\phi_n}(s)
\right]_{dQCD}
\biggr)\,.
\label{Kformf01}
\end{eqnarray}
\begin{figure}[ht]
\vspace{0.5 cm}
\begin{center}
\includegraphics[width=8.5cm,height=8.5cm]{v_K+K-_1_1.05.ps}
\includegraphics[width=8.5cm,height=8.5cm]{v_K+K-_1.05_4.ps}
\caption{(color online)The experimental data
\cite{Bisello:1988ez,Dolinsky:1991vq,Akhmetshin:1995vz,Achasov:2001ni,Pedlar:2005sj,Achasov:2007kg}
compared to the model fits results (see text for details).
The form factor at $J/\psi$ comes from its theoretical extraction
\cite{Seth:Jphi} from the data. It was not used in the fit. \label{KpKm}}
\end{center}
\vspace{0.5 cm}
\end{figure}
\begin{figure}[ht]
\vspace{0.5 cm}
\begin{center}
\includegraphics[width=8.5cm,height=8.5cm]{v_K0K0_1_1.05.ps}
\includegraphics[width=8.5cm,height=8.5cm]{v_K0K0_1.05_4.ps}
\caption{(color online )The experimental data
\cite{Mane:1980ep,Achasov:2001ni,Akhmetshin:2002vj,Akhmetshin:2003zn,Achasov:2006bv}
compared to the model fits results (see text for details).
\label{KzKz}}
\end{center}
\vspace{0.5 cm}
\end{figure}
\end{widetext}
The couplings in the part with subscript $fit$ were fitted to
the experimental data as well as the constants $\eta_\phi$ and $c_{\phi_0}^K$.
The values of $N_\rho, N_\omega$ and $N_\phi$ are listed in
Table~\ref{tab:kaon}.
The entry PDG in Table \ref{tab:kaon} implies, that
masses and widths as given in PDG2008 \cite{Amsler:2008zz} were used.
The masses and widths of the radial excitations, which were not measured,
were calculated assuming an equidistant mass spectrum
and a linear relation between the mass and the width of a given resonance
\bea
m_{j_n}^2=m_{j}^2 ( 1 + 2n ), \, \
\Gamma_{j_n} = \gamma_j m_{j_n} , \ \ \ j= \rho,\omega,\phi \ .
\label{mn}
\end{eqnarray}
The value of $\gamma_\rho$ was calculated from Eq.(\ref{mn}) for
$n=0$, the other values
were fitted to the data.
Two versions of the model were investigated: the `unconstrained'
version were the couplings between kaons and $\rho_n$, $\omega_n$ and
$\phi_n$ are not related and the `constrained' version were
$c_{\omega_n}^K=c_{\rho_n}^K \ , \ n=0,\cdots \infty$.
The `constrained' model
is not able to reproduce data as good as the `unconstrained' model,
however as evident from
Tab. {\ref{tab:kaon}}
the corrections to the assumption $c_{\omega_n}^K=c_{\rho_n}^K$ are
small for the lowest two resonances.
Despite this, the two models predict completely different
asymptotic behaviour of the neutral kaon form factor
in the region, where no data are available (Fig. \ref{KzKz}).
The `constrained' model, being closer to the SU(3) symmetric case
where the neutral kaon form factor vanishes, arrives at
significantly smaller
predictions.
The values of the couplings, which were not fitted, were calculated
from the formula
\bea
&&c^K_{j_n}=\frac{(-1)^n \Gamma(\beta^K_j-1/2)}{\alpha'\sqrt{\pi}m_{j_n}^2
\Gamma(n+1)
\Gamma(\beta^K_j-1-n)}, \nonumber \\
&&\alpha'=1/(2m_{j_0}^2), \ \ j = \rho,\omega,\phi
\label{cn}
\end{eqnarray}
\noindent
with the exception of the couplings next to the last fitted, which
were calculated from the normalization requirements
\bea
\sum_{n=0}^{\infty} c^K_{j_n} = 1 \, , \ \ \ \ \ j = \rho,\omega, \phi \, .
\label{c3}
\end{eqnarray}
\noindent
Breit-Wigner propagators
\bea
BW_{\alpha}^c = \frac{ m^2_{\alpha} }{m^2_{\alpha} - s
- i m_{\alpha} \Gamma_{\alpha}} , \ \
\alpha = \omega_n
\end{eqnarray}
with constant widths were used for all $\omega_n$,
Breit-Wigner propagators with s-dependant widths
\bea
&&\kern-10pt BW^{j}_{\phi_n} = \frac{ m^2_{\phi_n} }{m^2_{\phi_n} - s
- i m_{\phi_n} \Gamma_{\phi_n}^{j}}, \ \nonumber \\
&& \Gamma_{\phi_n}^{j} = \frac{ m^2_{\phi_n}}{s}
\biggl(\frac{s - 4 m_j^2}{ m^2_{\phi_n} - 4 m_j^2}\biggr)^{\frac{3}{2}}
\Gamma_{\phi_n},\nonumber \\
\ && j = K^+, K^0
\end{eqnarray}
\noindent
were used for $\phi_n$, the
radial excitations of $\phi$,
and the $GS$ Breit-Wigner functions (Eq.(\ref{GS1}))
were used for
$\rho_n$.
The parameters $\beta^K_j,\ \ j=\rho,\omega,\phi$,
were calculated from Eq.(\ref{cn})
using fitted $c^K_{j_0}$ parameter.
The results of the fits are summarized in Tab.~{\ref{tab:kaon}} and in
Figs. \ref{KpKm} and \ref{KzKz}. The high energy behaviour of
both form factors is completely driven by the CLEO \cite{Pedlar:2005sj}
measurement.
Following \cite{Bruch}, i.e. assuming isospin symmetry, one arrives
at the following predictions for the branching ratio of the
$\tau$-lepton decay into $K^-K^0\nu_\tau$
\bea
Br(\tau^- \to K^-K^0\nu_\tau) &=& (0.135/0.190)\cdot10^{-3}
\end{eqnarray}
for `unconstrained' and `constrained' models, respectively.
The model dependence is characterized by the spread between the two
results and is far larger than the errors
resulting from the fits of the parameters within one model.
These results can
be
compared with the PDG value \cite{Amsler:2008zz}
\bea
Br(\tau^- \to K^-K^0\nu_\tau) &=& (0.158\pm 0.16)\cdot10^{-3}
\end{eqnarray}
and are found to be reasonably consistent.
Within the same assumptions one can predict the $K^-K^0$ invariant mass
distribution and compare it (see Fig. \ref{taudistr}) with existing CLEO data
\cite{Coan:1996iu}. As evident from Fig. \ref{taudistr} both models
give very similar predictions and both agree with the data. Thus we
conclude that within the current experimental accuracy, which is,
however, very poor, isospin symmetry works well and the details
of the models do not play any role in its tests.
\begin{widetext}
\begin{table}
\begin{center}
\begin{tabular}{|c|r|c|c|r|c|c|}
\hline
Parameter & Input & Fit(1)& Fit(2) & PDG value& model(1) & model(2)\\
&&&&&&\\
\hline
$m_{\phi_0}$
& -
& 1019.415\,$\pm$\,0.004
& 1019.415\,$\pm$\,0.003
& 1019.455\,$\pm$\,0.020\,
& input
&input\\
$\Gamma_{\phi_0}$
& -
& 4.34\,$\pm$\,0.01
& 4.22\,$\pm$\,0.04
& 4.26\,$\pm$\,0.05
& input
& input \\
\hline
$m_{\phi_1}$
& 1680
&-
&-
&1680\,$\pm$\,20
& 1766
&1766\\
$\Gamma_{\phi_1}$
& 150
&-
&-
& 150\,$\pm$\,50
& 353
& 353\\
\hline
\hline
$m_{\rho_0}$
& 775.49
&-
&-
& 775.49\,$\pm$\,0.34
&input
&input \\
$\Gamma_{\rho_0}$
& 149.4
&-
&-
&149.4\,$\pm$\,1.0
&input
&input\\
\hline
$m_{\rho_1}$
& 1465
&-
&-
& 1465\,$\pm$\,25
& 1345
& 1345 \\
$\Gamma_{\rho_1}$
& 400
&-
&-
& 400\,$\pm$\,60
&259
&259 \\
\hline
$m_{\rho_2}$
& -
&1680\,$\pm$\,4
& PDG
& 1720$\,\pm$\,20
&1734
&1734 \\
$\Gamma_{\rho_2}$
& -
& 365\,$\pm$\,59
& PDG
& 250\,$\pm$\,100
&334
&334 \\
\hline
$m_{\omega_0}$
& 782.65
&-
&-
& 782.65\,$\pm$\,0.12
&input
&input\\
$\Gamma_{\omega_0}$
& \,\,\,\,\,\,8.49
&-
&-
& 8.49\,$\pm$\,0.08
& input
& input\\
\hline
$m_{\omega_1}$
& 1425
&-
&-
& 1400-1450
&1356
&1356 \\
$\Gamma_{\omega_1}$
&-
&145\,$\pm$\,9
&PDG
& 180-250
&678
&678\\
\hline
$m_{\omega_2}$
&-
&1729\,$\pm$\,76
&PDG
& 1670\,$\pm$ 30\,
&1750
&1750 \\
$\Gamma_{\omega_2}$
&-
&245\,$\pm$\,9
& PDG
& 315\,$\pm$\,35
&875
&875\\
\hline
\hline
$\eta_{\phi}$
& -
& 1.040\,$\pm$\,0.007
& 1.055\,$\pm$\,0.010
&-
&-
&- \\
\hline
$\beta_{\phi}$
& $c^K_{\phi_0}$ (\ref{cn})
& 1.97\,$\pm$\,0.02
& 1.91\,$\pm$\,0.02
& -
& -
& - \\
$\gamma_{\phi}$
& -
& 0.2
& 0.2
& -
& input
& input \\
$c^K_{\phi_0}$
& -
& 0.985\,$\pm$\,0.006
& 0.947\,$\pm$\,0.009
& -
& input
& input \\
$c^K_{\phi_1}$
& -
& 0.0042\,$\pm$\,0.0015
& 0.0136\,$\pm$\,0.0024
&-
& 0.0084
& 0.0271\\
$c^K_{\phi_2}$
&-
& 0.0039\,$\pm$\,\,0.0026
&(\ref{c3})0.0214\,$\pm$\,0093
&-
& 0.0026
& 0.0088\\
$c^K_{\phi_3}$
&-
&(\ref{c3})0.0033\,$\pm$\,0.0067
&-
&-
&0.0012
&- \\
$\sum\limits_{n=N_\phi+1}^{\infty} c^K_{\phi_n}$
& model
& 0.0036
& 0.0180
&-
& 0.0036
& 0.0180\\
\hline
$\beta_{\rho}$
& $c^K_{\rho_0}$ (\ref{cn})
& 2.23\,$\pm$\,0.06
& 2.21\,$\pm$\,0.05
& -
& -
& - \\
$\gamma_{\rho}$
& -
& 0.193 (\ref{mn}) (= $\Gamma_{\rho}$/$m_{\rho}$)
& 0.193 (\ref{mn}) (= $\Gamma_{\rho}$/$m_{\rho}$)
& -
& input
& input \\
$c_{\rho_0}^K$
&-
&1.138\,$\pm$\,0.011
&1.120\,$\pm$\,0.007
&-
& input
& input \\
$c_{\rho_1}^K$
& -
& -0.043\,$\pm$\,0.014
& -0.107\,$\pm$\,0.010
&-
& -0.087
& -0.078\\
$ c_{\rho_2}^K$
& -
&-0.144\,$\pm$\,0.015
&-0.028\,$\pm$\,0.012
&-
& -0.020
& -0.019\\
$ c_{\rho_3}^K$
&-
&-0.004\,$\pm$\,0.007
&(\ref{c3})0.032\,$\pm$\,0.017
&-
& -0.0084
& -0.0079\\
$ c_{\rho^{_4}}^K$
&-
&(\ref{c3})0.0662\,$\pm$\,0.0243
&-
&-
&-0.0045
&-\\
$\sum\limits_{n=N_\rho+1}^{\infty} c^K_{\rho_n}$
& model
&-0.0132
&-0.0170
&-
&-0.0132
&-0.0170\\
\hline
$\beta_{\omega}$
& $c^K_{\omega_0}$ (\ref{cn})
& $\beta_\rho$
& 2.75\,$\pm$\,0.06
& -
& -
& - \\
$\gamma_{\omega}$
& -
& 0.5
& 0.5
& -
& input
& input\\
$c_{\omega_0}^K$
& -
& $c^K_{\rho_0}$
& 1.37\,$\pm$\,0.03
&-
& input
& input\\
$c_{\omega_1}^K$
& -
& $c^K_{\rho_1}$
&-0.173\,$\pm$\,0.003
&-
& -0.087
&-0.345\\
$ c_{\omega_2}^K$
& -
&$c^K_{\rho_2}$
&-0.621\,$\pm$\,0.020
&-
& -0.020
& -0.026\\
$ c_{\omega_3}^K$
&-
& $c^K_{\rho_3}$
& (\ref{c3})0.43\,$\pm$\,0.04
&-
& -0.0084
& -0.0079\\
$ c_{\omega_4}^K$
& -
&$c^K_{\rho_4}$
&-
&-
&-0.0045
&-\\
$\sum\limits_{n=N_\omega+1}^{\infty} c^K_{\omega_n}$
& model
&$\sum\limits_{n=N_\rho+1}^{\infty} c^K_{\rho_n}$
&-0.0096
&-
&-0.0132
&-0.0096\\
\hline
$\chi^2/d.o.f.$&- & 277/256 &221/260&-&-&-\\
\hline
\end{tabular}
\caption{{\it Parameters of the kaon form factors
and results of the fit to the data. Masses and widths
are given in MeV. The column 'Fit(1)' (Fit(2)) contains the values
of the constrained (unconstrained) fits.}}.
\label{tab:kaon}
\end{center}
\end{table}
\end{widetext}
\begin{figure}[ht]
\vspace{0.5 cm}
\begin{center}
\includegraphics[width=8.5cm,height=8.5cm]{normalized_distributions.ps}
\caption{(color online). Normalized distributions
$\frac{d\Gamma(\tau\to K^-K^0\nu_\tau)/d\sqrt{Q^2}}{\Gamma(\tau\to
K^-K^0\nu_\tau)}$ of the kaon pair invariant mass predicted within
two models, described in the text.
CLEO data \cite{Coan:1996iu} normalized to the
total number of events are also shown.
\label{taudistr}}
\end{center}
\vspace{0.5 cm}
\end{figure}
\section{Monte Carlo implementation of $K^+ K^-$ and $K^0 \bar{K}^0$}
\begin{figure}
\begin{center}
\includegraphics[width=8.5cm,height=7.cm]{v_ifs_tot_dok1.ps}
\includegraphics[width=8.5cm,height=7.5cm]{sigma_10_K0.ps}
\caption{(color online) Differential cross section for $\sqrt{s} = 10.52 {\rm GeV}$
of the processes
$e^+ e^- \to K^+K^-\gamma (\gamma)$ and
$e^+ e^- \to K^0\bar{K}^0\gamma (\gamma)$
with angular cuts. \label{cs10IFSNLO} }
\end{center}
\end{figure}
The event
generator PHOKHARA has been extended to generate
$K^+ K^-$ and $K^0 \bar{K}^0$ final states.
In this section we present the implementation and results for
the region below the narrow resonances $J/\psi$ and $\psi(2S)$.
Charged kaons have been implemented
in the same way
as the $\pi^+ \pi^-$ channel \cite{Czyz:2002np,Czyz:PH03},
with the kaon form factor described in Section \ref{sec2}.
The NLO FSR corrections have been implemented as well.
For the neutral kaons the corrections are limited
to ISR.
With the enormous luminosity of B factories one expects
hundreds of events even for $Q^2$ between 3 and 4 $GeV^2$ and large statistic
around the $\phi$ resonance (Fig. \ref{cs10IFSNLO}).
The next-to-leading FSR corrections are relevant
for a measurement
in the neighbourhood of
the $\phi$ resonance if an accuracy better then 10\% is aimed
(Fig. \ref{diff10_I_F}).
\begin{figure}
\begin{center}
\includegraphics[width=7.5cm,height=6.5cm]{v_isr_ifs_10_cie1.ps}
\caption{(color online) Comparison of IFSNLO and ISRNLO distributions with angular cuts.
}
\label{diff10_I_F}
\end{center}
\end{figure}
\section{ \label{sec3}
Narrow resonances and the radiative return
}
\begin{figure}[ht]
\vspace{0.5 cm}
\begin{center}
\includegraphics[width=8.5cm,height=12.5cm]{diagram_ifs_10_n.ps}
\caption{(color online). The contributions to the radiative return
cross section included in PHOKHARA. Label '1' at a photon line means
that the photon is hard.
\label{rrdiag}}
\end{center}
\vspace{0.5 cm}
\end{figure}
The radiative return, as implemented now in PHOKHARA,
receives contributions from a multitude of amplitudes
shown in Fig. \ref{rrdiag}.
The notation introduced in this figure also applies to the narrow
resonance amplitudes.
Below we list the formulae
for $\mu^+\mu^-$, $\pi^+\pi^-$, $K^+K^-$ and $\bar K^0K^0$.
For the $\bar K^0K^0$ amplitude $FSR$ emission is not present.
We
explicitly indicate the vacuum polarization contributions
as well as the $\phi$ contribution.
We assume that in the vacuum polarization $J/\psi$, $\psi(2S)$
and $\phi$ were not accounted for.
The differential cross section given by PHOKHARA
reads
\bea
&&d\sigma=\nonumber \\
&& |M_{\gamma_1,LOISR} \cdot C^{VP}_{R,P}(Q^2)
+ M_{\gamma_1,LOFSR} \cdot C^{VP}_{R,P}(s)|^2 d\Phi_1
\nonumber \\
&&+|M_{2\gamma,ISR} \cdot C^{VP}_{R,P}(Q^2)|^2 d\Phi_2\nonumber \\
&& +2\ {\rm Re}(M_{\gamma_1,NLOISR}\times M^\dagger_{\gamma_1,LOISR})
\cdot |C^{VP}_{R,P}(Q^2)|^2 d\Phi_1\nonumber \\
&&+|M_{\gamma_1,ISR \ ;\ \gamma,FSR}\cdot C^{VP}_{R,P}((Q+k_\gamma)^2)|^2
d\Phi_2 \nonumber \\
&&+2\ {\rm Re}(M_{\gamma_1,LOISR}^{NLOFSR}\times M^\dagger_{\gamma_1,LOISR}
) \cdot |C^{VP}_{R,P}(Q^2)|^2 d\Phi_1
\nonumber \\
&&+|M_{\gamma,ISR \ ;\ \gamma_1,FSR}
\cdot C^{VP}_{R,P}((Q+k_{\gamma_1})^2)|^2 d\Phi_2\nonumber \\
&&+2\ {\rm Re}(M_{\gamma_1,LOFSR}^{NLOISR}\times M^\dagger_{\gamma_1,LOFSR})
\cdot |C^{VP}_{R,P}(s)|^2 d\Phi_1\ ,
\nonumber \\
\end{eqnarray}
\noindent
where
\bea
C^{VP}_{R,P}(s) &=& \frac{1}{1-\Delta\alpha(s)}- \frac{3\Gamma_e^\phi}
{\alpha m_\phi} \ BW_\phi(s) \delta_P
\nonumber \\
&&+ C_{J/\psi,P}(s)
+ C_{\psi(2S),P}(s)\ ,
\end{eqnarray}
\bea
C_{R,P}(s) = \frac{3\sqrt{s}}{\alpha}
\frac{\Gamma_e^R (1+c_P^R)}{s-M_R^2+i \Gamma_R M_R} \ .
\end{eqnarray}
\noindent
and $d\Phi_1$ ($d\Phi_2$) denote the phase space with one (two) photon(s)
in the final state with all statistical
factors included.
For $P=\mu$ and $P=\pi$, $c_P^R=0$ (no direct decay of the narrow
resonances into $\mu^+\mu^-$ and $\pi^+\pi^-$),
while $\delta_P=0$ for $P=K$
and $\delta_P=1$ for $P=\mu$ and $P=\pi$. The $\phi$ contributions
to the kaon pair production are included in the kaon form factor,
hence $\delta_K=0$.
The notation and the detailed description of the narrow resonance
contribution to the amplitude
can be found in \cite{Czyz:2009vj} (see \cite{Yuan:2003hj} for similar studies).
From \cite{Czyz:2009vj} we also take
$|c_{K^+}^{J/\psi}|=1.27\pm 0.32$ and $|c_{K^+}^{\psi(2S)}|=2.94\pm 0.99$.
The information on the neutral kaon couplings to the narrow resonances
is almost nonexisting and we use the lower limits of
$|c_{K^0}^{J/\psi}|=2.81$, $|c_{K^0}^{\psi(2S)}|=5.35$, which
correspond to the upper limit on the neutral kaon form factor
(see \cite{Czyz:2009vj} for details).
The phases are essentially not known and we use $100^{\circ}$ to obtain
the numerical values in the next section.
\section{ The implementation of narrow resonances
into the Monte Carlo event generator PHOKHARA}
The tests of the ISR part of the implementation of the narrow resonances
where straightforward and followed the standard tests we perform
for each new channel
\cite{Rodrigo:2001kf,Czyz:2002np,Czyz:PH03,Nowak,Czyz:PH04,Czyz:2004nq,Czyz:2005as,Czyz:2008kw}.
The comparisons were made with the analytic formulae of \cite{Berends:1987ab}
separately for one and two photon emission. The precision of the comparisons
was at the level of a small fraction of a per mill, proving
the technical precision of the program at that level. The independence of the
results on the separation parameter between soft photon, calculated
analytically and hard photon, generated by means of the Monte Carlo method,
was also tested with that precision.
\begin{figure}
\begin{center}
\includegraphics[width=7.5cm,height=6.5cm]{v_diff_mc-7-4.ps}
\caption{(color online) Comparison between invariant mass distributions obtained
with $w=10^{-4}$ and $w=10^{-7}$.
}
\label{FSRww}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=7.5cm,height=6.5cm]{v_-7-4_xx.ps}
\caption{(color online) Comparison between invariant mass distributions obtained
with $w=10^{-4}$ and $w=10^{-7}$ smeared with Gaussian with the standard
deviation of 14.5 MeV.
}
\label{FSRww_smeared}
\end{center}
\end{figure}
The implementation of the NLO FSR part is more tricky. The analytic
formulae used
in \cite{Czyz:PH03,Czyz:PH04} for soft photon contributions are still valid,
which we have checked numerically with the precision of 0.02\%.
However if one chooses the separation parameter between soft and hard part
at the usual value $w=10^{-4}$, which corresponds to the photon energy
$E_\gamma = 1$~MeV for $\sqrt{s}=10$~GeV,
the `soft' integral
receives contributions from the whole resonance region,
as a consequence of the small width ( $\Gamma_{J/\psi} = 93.4$~keV).
For a cutoff of $10^{-4}$ the part of the matrix element,
which multiplies the soft emission factor is rapidly varying and
the basic assumption underlying the whole approach, that the soft
emission can be integrated analytically with the multiplicative
remainder being constant, is not longer valid. Pushing the value
of the cutoff to an extremely small value, say $10^{-7}$, solves
this problem. However, single-photon emission is not an adequate
description for such soft photons and in principle one should
use exponentiation. From the technical side this is reflected
in the appearance of negative weights. Inclusion of YFS-like
multi-photon production would allow to cure this problem. However
since this would amount to completely restructuring our Monte
Carlo generator,
we have adopted a simpler approach,
which gives correct distributions, when
convoluted with an energy resolution typical
for a detector at a $\phi$- or $B$-meson factory.
\begin{figure}
\begin{center}
\includegraphics[width=7.5cm,height=6.5cm]{v_qq_all_a.ps}
\caption{(color online) Comparisons of the invariant mass distributions obtained
with $w=10^{-7}$ taking into account only ISRNLO contributions
and the complete
(ISR+FSR)NLO result.
}
\label{FSRNLOvsISR}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=7.5cm,height=6.5cm]{v_qq_xx_IF_m.ps}
\caption{(color online) Comparisons of the invariant mass distributions obtained
with $w=10^{-7}$ taking into account only ISRNLO contributions
and the complete
(ISR+FSR)NLO result. Detector smearing effect are taken into account.
}
\label{FSRNLOvsISRxx}
\end{center}
\end{figure}
Due to the finite
detector resolution one never observes
the true distribution of the events, but the one convoluted with the
detector resolution function. This increase of the effective width
by about a factor hundred is sufficient to cure the problem.
For a cutoff of $10^{-4}$ the distribution remains smooth and we
can produce the unweighted events sample.
The result will, as expected, depend
on the resolution of the detector. To check if this
is true for the realistic energy resolution of the
BaBar detector \cite{Aubert:2003sv} of 14.5 MeV we have compared
the muon invariant mass distributions obtained with $w=10^{-4}$ and
$w=10^{-7}$, smeared with a Gaussian distribution with a standard
deviation of 14.5 MeV. Even if the non-smeared distributions are
completely different, as shown in Fig. \ref{FSRww} the smeared
distributions agree within 2 per mill as shown in Fig. \ref{FSRww_smeared}.
This 2 per mill is the intrinsic error coming from the method we use,
but the generator should be accurate enough for any practical purposes.
It is interesting to observe (Fig. \ref{FSRNLOvsISR} )
that the FSRNLO contributions
fill completely the interference dip, still visible if only ISR corrections
are taken into account. Thus the absence of the dip in the
observed invariant mass distribution is not only the effect
of the detector smearing.
\begin{figure}
\begin{center}
\includegraphics[width=7.5cm,height=6.5cm]{v_w-4_xx_IF_m.ps}
\caption{(color online) Relative ratio of the invariant mass distributions
taking into account only ISRNLO contributions
and the complete
(ISR+FSR)NLO result. Detector smearing effect are taken into account.
}
\label{rFSRNLOvsISRxx}
\end{center}
\end{figure}
The huge FSRNLO corrections seen in Fig.~\ref{FSRNLOvsISR} are washed out
if one looks at the detector smeared distributions shown
in Fig. \ref{FSRNLOvsISRxx}. The corrections are
seen more accurately in Fig. \ref{rFSRNLOvsISRxx},
where the relative difference is shown.
The FSRNLO corrections cannot be neglected if one aims at
a precision better then 10\%, unless one considers only the integral
over the whole resonance region (together with the side bands as shown in
Fig. \ref{FSRNLOvsISRxx}).
In the integrated cross section a large part of
the corrections cancel
($\sigma_{ISRNLO}=6.901$~pb, $\sigma_{IFSRNLO}= 6.954 $~pb).
Identical tests were performed for the pion pair production with identical
conclusions, so we do not present them here.
\section{\label{sec5} Summary}
New parametrizations of the pion and kaon form factors, based on the
"dual QCD model", are presented which are derived from a fit to a
combination of old measurments and more recent experimental results in
the energy region above the $\rho$-resonance. These form factors and the
results of a recent analysis of the direct hadronic coupling of
$K\bar K$ to $J/\psi$ and $\psi(2S)$ are incorporated in the Monte
Carlo generator PHOKHARA which is now also adapted to the simulation of
narrow resonances, including the effects of ISR and FSR in NLO.
\begin{acknowledgments}
Henryk Czy\.z is grateful for the support and the kind hospitality
of the Institut f{\"u}r Theoretische Teilchenphysik
of the Karlsruhe Institute of Technology.
\end{acknowledgments}
|
\section{Introduction}
We call $(M,\LL,g)$ a compact $C^2$ lamination if
$\LL$ is an $n$ dimensional lamination of class $C^2$ on a compact
metric space $M$ and if $g$ is a leafwise Riemannian metric of class
$C^2$.
(For the precise definition, see Sect.\ 2.)\ \
Then the leafwise Laplacian $\Delta f$ is defined for any continuous
leafwise $C^2$ function on $M$. A probability measure $m$ on $M$ is called
{\em harmonic} if for any such $f$, we have $m(\Delta f)
=0$.
A harmonic measure always exists for any compact $C^2$ lamination.
Given a harmonic measure $m$, there is a saturated conull set
$M^*$ such that a positive harmonic function $h$, called the
characteristic harmonic function, is defined
on the universal cover of each leaf in $M^*$ up to a constant
multiple.
This function is obtained in the way of
describing the measure $m$ on each local chart.
We first show (Theorem \ref{12}) that if $m$ is
ergodic and not completely invariant, then for any leaf in a saturated
conull set,
the characteristic harmonic function $h$ is unbounded.
A compact $C^2$ lamination $(M,\LL,g)$ of dimension $d$+1
is called {\em hyperbolic} if the metric $g$ has
curvature -1 on each leaf. The universal cover of each leaf is isometric
to the hyperbolic space $\DD$,
and the characteristic harmonic
function $h$ corresponds to a probability
measure
$\mu$ on
the boundary at infinity $\boundary$. It depends upon
the choice of a base point in $\DD$, but its equivalence class is
uniquely determined by the leaf.
We show (Theorem \ref{13}) that if $m$ is ergodic harmonic and not
completely invariant, then for any leaf in a saturated conull set,
the measure $\mu$ is
singular to the Lebesgue measure of $\boundary$.
\begin{definition}
A harmonic measure $m$ is called of {\em Type I} \ if the measure
$\mu$ of $\boundary$ is a point mass for
any leaf in a saturated conull set, and of {\em Type II} \ if the support of
$\mu$ is the whole $\boundary$.
\end{definition}
The main theorem of this paper is the following.
\begin{theorem} \label{1}
An ergodic harmonic measure is either of Type I or of Type II.
\end{theorem}
In Sect.\ 2, we prepare some necessary prerequisites about
harmonic measures. Especially the characteristic harmonic
function $h$ is defined. In Sect.\ 3, after a brief
description of the
leafwise Brownian motion, we study its reverse process.
The reverse process plays a crucial role
in the proof of the unboundedness of $h$ (Sect.\ 3)
and the singularity of $\mu$ (Sect.\ 4).
In Sect.\ 5, we study the leafwise unit tangent
bundle $N$ of a compact hyperbolic lamination $\LL$.
There is a naturally defined lamination $\HH$ on $N$
of the same dimension as $\FF$.
Generalizing a result in [BM],
we discuss one to one correspondance between harmonic measures
on $\LL$ and pointed harmonic measures on $\HH$,
the latter being defined in Sect.\ 5.
Finally the proof of Theorem \ref{1}, as well as
some examples, is given in Sect.\ 6.
The author is grateful to Masahiko Kanai for supplying
him necessary knowledge about positive harmonic functions,
and to the referee for valuable comments which are helpful
for the improvement of the paper.
\section{Harmonic measure}
Let $M$ be a compact metric space, covered by a finite number
of open sets $E_i$. Assume there is a homeomorphism
$\varphi_i:E_i\to U_i\times Z_i$, where $U_i$ is an open ball in
$\R^n$ and $Z_i$ is a locally compact metric space.
If $E_i\cap E_j\neq\emptyset$, then the transition function
$\psi_{ji}=\varphi_j\circ\varphi_i^{-1}$ is defined as a homeomorphism
from $\varphi_i(E_i\cap E_j)$ onto $\varphi_j(E_i\cap E_j)$.
Assume that the transition function is of the form
$$
\psi_{ji}(u,z)=(\alpha(u,z),\beta(z)),
$$
where $\alpha$ and $\beta$ are continuous, and $\alpha$ is
of class $C^2$ with respect to the first coordinate $u$ and its
first and second derivatives are continuous in $z$.
A subset of $M$ is called a {\em plaque}
if it is of the form $\varphi_i^{-1}(U_i\times z)$,
and a {\em transversal} if $\varphi_i^{-1}(u\times Z_i)$.
A maximal pathwise connected countable union of
plaques is called a {\em leaf}. This gives birth to a decompositon
$\LL$ of $M$ into leaves, which is called a lamination of class $C^2$.
A leaf naturally has a structure of $n$ dimensional $C^2$ manifold.
A field of leafwise metric tensors is called a
{\em leafwise Riemannian metric of class $C^2$} if
its leafwise derivatives up
to order 2 (including order 0) are continuous on $M$.
In this paper the triplet $(M,\LL,g)$ is simply refered to
as a {\em compact $C^2$ lamination}.
By the compactness of $M$, each leaf of $\LL$ is complete
and of bounded geometry.
The leafwise volume defined by $g$ is denoted by ${\rm vol}$.
Henceforce we depress the homeomorphism $\varphi_i$ and
consider $U_i\times Z_i$ as an open subset of $M$,
which is called a {\em local chart}.
A function $f:M\to\R$ is said to be {\em continuous leafwise $C^2$}
if it is of class $C^2$ in each leaf and its derivative up to
order 2 is continuous on $M$. Then the leafwise Laplacian
$\Delta f$ with respect to $g$ is defined, and is a continuous
function on $M$.
\begin{definition}
A probability measure $m$ on $M$ is called {\em harmonic}
if $m(\Delta f)=0$ for any continuous leafwise $C^2$
function.
\end{definition}
\begin{remark}
A harmonic measure always exits for any compact
$C^2$ lamination $(M,\LL,g)$.
\end{remark}
See [C] Theorem 3.5 for a simple proof using the Hahn-Banach theorem.
Here is a structure theorem of a harmonic measure on a local chart.
\begin{theorem} \label{2}
Assume $m$ is a harmonic measure on a compact $C^2$ lamination.
For any local chart $U\times Z$, there are a measure $\nu$
on $Z$ and a
function $h:U\times Z\to\R$
with the following properties.
{\rm (1)} $h$ is positive and $m$-measurable.
{\rm (2)} For $\nu$-a.a.\ $z$, the restriction of
$h$ to the plaque $U\times z$
is harmonic and $h$vol is a probablity measure of the plaque.
{\rm (3)} For any continuous function with
support in $U\times Z$, we have
$$
m(f)=\int_Z\int_{U\times{z}}f(u,z)h(u,z)d{\rm vol}(u)d\nu(z).
$$
Furthermore if a probability measure $m$ on $M$ is represented
in this way in any local chart, then $m$ is harmonic.
\end{theorem}
\begin{notation}
The theorem says that the measure $m$ restricted to $U\times Z$
is disintegrated in such a way that the conditional probability measure on each fiber
$U\times z$ is $h(\cdot,z){\rm vol}$ and the push forward measure on the base
$Z$ is $\nu$. Henceforth this is denoted as
\begin{equation} \label{e}
m\vert_{U\times Z}=\int_Zh{\rm vol}\,d\nu.
\end{equation}
\end{notation}
\bd
By the disintegration theorem
we have
$$m\vert_{U\times Z}=\int_{Z} m_z\,d\nu(z),
$$
where $m_z$ is a probability measure
on $U\times z$ and the assignment $z\mapsto m_z$ is measurable.
The measure $\nu$ is
the push forward of $m$ by the projection $p_2:U\times Z\to Z$,
and is not necessarily a probability measure.
Denote the other projection by $p_1:U\times Z\to U$.
The leafwise Riemannian metric on each plaque $U\times z$ is
transfered to a Riemannian metric on $U$,
and the corresponding Laplacian on $U$ is denoted by $\Delta_z$.
Consider any function $f$ from the space $C^2_c(U)$ of
the $C^2$ functions with compact support,
and any continuous
function $g$ on $Z$ with compact support.
Then the product $f\circ p_1\,g\circ p_2$
is a continuous leafwise $C^2$ function
whose support is contained in $U\times Z$ and we have
$$\Delta(g\circ p_2\,f\circ p_1)=g\circ p_2\,\, \Delta(f\circ p_1)
\ \mbox{ and }\
m_z(\Delta(f\circ p_1))=m_z(\Delta_zf).
$$
Since $m(\Delta(g\circ p_2\,f\circ p_1))=0$, we have
$$\int_{Z}m_z(\Delta_z f)g(z)d\nu(z)=0.
$$
By the measurablility of the assignment $z\mapsto m_z$
and the boundedness of $\Delta_zf$, $m_z(\Delta_z f)$ is an integrable function
on $Z$ and thus $m_z(\Delta_z f)\nu$ is a signed measure on $Z$,
for which an arbitrary compactly supported continuous
function $g$ integrates to 0.
This implies that for $\nu$-a.a.\ $z$, $m_z(\Delta_z f)=0$.
Since $C^2_c(U)$ has a countable dense subset $S$, there is
a $\nu$-conull set $Z^*$ of $Z$ such that if $z\in Z^*$,
$m_z(\Delta_z f)=0$ for
any $f\in S$, and therefore for any $f\in C^2_c(U)$.
But as is well known ([N]), $m_z(\Delta_z f)=0$ for any
$f\in C^2_c(U)$ if and only if $m_z=h_z\,{\rm vol}$ for a harmonic
function $h_z$ on $U$ with respect to the Laplacian $\Delta_z$.
Setting $h(u,z)=h_z(u)$, we obtain (\ref{e}).
Next we are going to show that the function $h$
is measurable.
Consider another measure $m_0$ on $U\times Z$, given
by $\int_Z {\rm vol}/{\rm vol}(U\times z)d\nu(z)$. Clearly $m$ and
$m_0$ are mutually equivalent and thus we have
$$
m=km_0$$ for some $m_0$-measurable (equivalently $m$-measurable) function $k$.
But the uniqueness of the disintegration implies that for $\nu$-a.a.\ $z$,
$$
h(u,z)=k(u,z)/{\rm vol}(U\times z),
$$
showing that $h$ is measurable.
Finally the converse statement is easy to show.
\qed
\medskip
As an immediate corollary, we have:
\begin{corollary} \label{2A}
If a function $f$ on $M$ is $C^2$ on each leaf and $\Delta f$
is $m$-integrable, then $m(\Delta f)=0$.
\end{corollary}
\begin{remark} In [G], harmonic measures are defined
by the property in Corollary \ref{2A}, and the structure
theorem is obtained. Our proof of Theorem \ref{2} shows the equivalence of
the two definitions.
\end{remark}
\smallskip
Suppose two local charts $U\times Z$ and $U'\times Z'$ intersect
and the harmonic measure $m$ is decomposed on each local chart
as
$$m\vert_{U\times Z}=\int_Zh\,{\rm vol}\,d\nu\ \mbox{ and }\
m\vert_{U'\times Z'}=\int_{Z'}h'\,{\rm vol}\,d\nu',
$$
then
in the intersection of $\nu$-a.a.\ plaque $U\times z$
and $\nu'$-a.a.\ plaque $U'\times z'$, we have
\begin{equation} \label{e1}
h'/h=d\nu/d\beta\nu',
\end{equation}
where $\beta$ is the holonomy map from (a part of) $Z'$ to $Z$.
On one hand this shows that $\nu$ and $\nu'$ are equivalent via
the holonomy map. More generally we have:
\begin{proposition} \label{3}
A harmonic measure $m$ is leafwise smooth, i.e.,
{\rm (1)} If a Borel set $B\subset M$ satisfies ${\rm vol}(B\cap L)=0$ for any leaf
$L$, then $m(B)=0$.
{\rm (2)} If a Borel set $B$ is $m$-null, then the set $\widehat B$ is also
$m$-null, where $\widehat B$ is the union of the leaves $L$ such
that ${\rm vol}(B\cap L)>0$.
\end{proposition}
On the other hand the equality (\ref{e1})
shows that on the intersection of two plaques,
the function $h'$ is a positive constant multiple of $h$.
Dividing $h'$ by that constant,
one can continuate $h$ along a chain of plaques. Of course
this does not yield a function on a leaf, since
there will be a monodromy for $h$. However we will get
a function on the holonomy cover of a leaf.
In what follows, when we say ``for $m$-a.a.\ leaf $L$'',
this means `` there exists a saturated conull set
$M^*$ and for any leaf $L$ in $M^*$''.
\begin{proposition} \label{4}
{\rm (1)} For $m$-a.a.\ leaf $L$, the function $h$ has
a well defined prolongation as a positive harmonic function
on the holonomy cover $\hat L$. On $\hat L$ two such functions
which start from different plaques are unique up to
a positive constant multiple.
{\rm (2)} Given a path in $L$ the ratio
of $h$ at the initial point and the terminal point of
any lift of the path to $\hat L$ is the same.
\end{proposition}
\noindent
{\sc Proof}.
To see (1), cover $M$ with a finite number
of local charts $U_i\times Z_i$. Then there is a $\nu$-conull set
$Z_i^*$ of each $Z_i$ such that the harmonic function $h$ is
defined on $U_i\times Z_i^*$. The saturation of the union of
$Z_i\setminus Z_i^*$ is $m$-null by Proposition \ref{3},
and for any leaf $L$ in the complement $M^*$,
the function $h$ has a prolongation on its holonomy cover
$\hat L$.
The uniqueness part in (1), as well as
(2), follows immediately from the construction.
\qed
\bigskip
Of course the harmonic function $h$ has a lift to the universal
covering space $\tilde L$ of $m$-a.a.\ leaf $L$, which will be
denoted by the same letter $h$.
The above statement (2) holds also for lifts of paths to the universal
covering
space. Let $\Gamma$ be the deck transformation group of the covering
map $\tilde L\to L$. Then we have:
\begin{corollary} \label{4+}
For any $\gamma\in\Gamma$, $h\circ\gamma$ is a constant multiple
of $h$.
\end{corollary}
\noindent
{\sc Proof}. Join two points $x\in\tilde L$ and $y\in\tilde L$
by an arc $c$.
Then $\gamma x$ and $\gamma y$ are joined by $\gamma c$.
Two arcs $c$ and $\gamma c$ are lifts of the same arc in $L$.
Therefore we have
$$
h(y)/h(x)=h(\gamma x)/h(\gamma y).$$
Since $x$ and $y$ are arbitrary, this shows that the function
$h(\gamma x)/h(x)$ is independent of $x$.
\qed
\medskip
\begin{definition} The function $h$ in Proposition \ref{4} is called
the {\em characteristic harmonic function of $m$}.
\end{definition}
Notice that the characteristic harmonic function is defined only
up to a positive constant multiple.
\medskip
A harmonic measure $m$ is called {\em completely invariant} if
the characteristic harmonic functions are constant on
(the holonomy cover of) $m$-a.a.\ leaves. In this case
$m$ corresponds to a transverse invariant measure, i.e.,
an assignment of a finite measure to each transversal
which is invariant by the holonomy maps.
Conversely a transverse invariant measure gives rise to
a harmonic measure $m$ whose characteristic harmonic function is constant on
$m$-a.a.\ leaf.
Only a special class of laminations admit completely invariant
measures.
\section{Brownian motion and its reverse process}
Let $(M,\LL, g)$ be a compact $C^2$ lamination. Denote by
$\Omega$ the space of all the continuous leafwise paths
$\omega:[0,\infty)\to M$ and for $t\geq0$, define a map
$X_t:\Omega\to M$ by $X_t(\omega)=\omega(t)$.
Let $\mathcal B$ be the $\sigma$-algebra of the Borel subsets
of $\Omega$ with respect to the compact open topology. It is well known,
easy to show, that $\mathcal B$ coincides with the minimal
$\sigma$-algebra for which $X_t$ ($t\geq 0$) is Borel.
In other words $\mathcal B$ is generated by the {\em cylinder sets}
$\{X_{t_1}\in B_1,\ldots,X_{t_r}\in B_r\}$ ($0\leq t_1<\cdots< t_r$,
$B_i$; a Borel subset of $M$).
The leafwise Riemannian metric $g$ gives the heat kernel
$p_t(x,y)$ ($t>0$) on each leaf. Define $p_t(x,y)$
for any two points $x,y\in M$ by setting $p_t(x,y)=0$
unless $x$ and $y$ lie on the same leaf. The heat kernel defines the
Wiener probability measure $W^x$ on $\Omega$ ($x\in M$).
For a cylinder set $\{X_{t_1}\in B_1,\ldots,X_{t_r}\in B_r\}$
($t_1>0$), we define
\begin{eqnarray}\label{eqn}
& W^x\{X_{t_1}\in B_1,\ldots, X_{t_r}\in B_r\}
\\
&=\int_{B_1}\cdots\int_{B_r}p_{t_1}(x,y_1)p_{t_2-t_1}(y_1,y_2)
\cdots p_{t_{r}-t_{r-1}}(y_{r-1},y_r)\,d{\rm vol}(y_r)\cdots d{\rm vol}(y_1).
\nonumber
\end{eqnarray}
Then $W^x$ satisfies the dropping condition, and
therefore it is defined not only for a cylinder set but also for any
set in $\mathcal B$. That is, $W^x$ is a probability measure on
$(\Omega,\mathcal B)$. It is concentrated on the subset
$\Omega_x=X_0^{-1}(x)$, since the probability measure
$p_t(x,\cdot){\rm vol}$ tends to the Dirac mass at $x$ as $t\to0$.
\begin{lemma} \label{24}
The system of measures $\{W^x\}_{x\in M}$ is Borel in the sense
that for any $S\in\mathcal B$, the assignment $x\mapsto W^x(S)$
is Borel.
\end{lemma}
\bd
Let $\mathcal C$ be the family of the subsets $S$ in $\Omega$
for which $M\ni x \mapsto W^x(S)$ is Borel,
and let $\mathcal A_0$ be the finite algebra formed by
finite disjoint unions of cylinder sets.
Then
$\mathcal A_0$ is contained in $\mathcal C$.
For $\{X_{t_1}\in B_{t_1}\}\in\mathcal A_0$, see [CC] Lemma 2.3.1.
General case follows easily from this.
For an isolated ordinal $\alpha>0$, define
$\mathcal A_\alpha$ to be the family of a subset which is
obtained from subsets of $\mathcal
A_{\alpha-1}$,
by a finite sequence of two operations; one,
taking a countable increasing union
and the other, countable decreasing intersection.
Then it is easy to show that $\mathcal A_\alpha$ forms a finite algebra.
Moreover $\mathcal A_\alpha$ is contained in $\mathcal C$,
since a pointwise limit of Borel functions is Borel.
For a limit ordinal $\alpha$, let
$\mathcal A_\alpha=\bigcup_{\beta<\alpha}A_\beta$.
Then again $\mathcal A_\alpha$ is a finite algebra contained in $\mathcal C$.
The increasing sequence $\{\mathcal A_\alpha\}$ stabilizes.
Define $\mathcal A=\mathcal A_{\alpha_0}$, where
$A_\beta=A_{\alpha_0}$ for any
ordinal $\beta\geq \alpha_0$. Then $\mathcal A$ is contained in $\mathcal C$.
On the other hand $\mathcal A$
is clearly a $\sigma$-algebra.
Therefore any Borel set, an element of the minimal $\sigma$-algebra
which contains $\mathcal A_0$, belongs to $\mathcal A$,
and hence to $\mathcal C$.
\qed
\bigskip
The expectation
of $W^x$ is denoted by $E^x$.
Applying Lemma \ref{24}, one can show that for any bounded Borel
function $f:M\to\R$, its diffusion
$D_tf$ is bounded Borel, where
$$
(D_tf)(x)=E^x[f(X_t)]=\int_M p_t(x,y)f(y)d{\rm vol}(y).
$$
More generally the diffusion operator $D_t$ defines a semigroup of contractions
on the space $L^p(M,m)$ ($1\leq p<\infty$) for
a harmonic measure $m$ and on $C(M)$,
the space of continuous functions ([C]).
Since $\{W^x\}$ is a Borel system of measures,
by integrating $W^x$ over any probability measure $m$ on $M$,
we get a probability measure $\PP_m$ on $\Omega$, i.e.,
$$\PP_m=\int_MW^xdm(x).
$$ Precisely
for any bounded Borel function $F:\Omega\to\R$,
$$
\PP_m(F)=\int_ME^x[F]dm(x).
$$
For $t\geq0$ let $\theta_t:\Omega\to\Omega$ denote the shift map by $t$,
i.e.,
$(\theta_t\omega)(t')=\omega(t+t')$.
\begin{theorem} \label{4A}
The probability measure $m$ is harmonic if and only if the probability measure
$\PP_m$ is $\theta_t$-invariant for any $t\geq0$.
\end{theorem}
For the proof, see [CC] Theorem 2.3.7.
A harmonic measure $m$ is called {\em ergodic} if whenever it is written as
a nontrivial linear combination of two harmonic measures $m_1$
and $m_2$, we have $m=m_1=m_2$.
\begin{theorem} \label{5}
Let $m$ be a harmonic measure. Then the following conditions
are equivalent.
{\rm (1)} $m$ is ergodic.
{\rm (2)} For any saturated Borel set $M'$ in $M$, we have either $m(M')=0$ or
$m(M')=1$.
{\rm (3)} If $f\in L^1(M,m)$ satisfies $D_tf=f$ for any $t\geq0$, then $f$ is a
constant.
{\rm (4)} $\PP_m$ is ergodic with respect to the semiflow
defined by the shift map $\theta_t$, i.e.,
if a Borel subset $S$ satisfies $\theta_t^{-1}(S)=S$
for any $t\geq0$, then either $\PP_m(S)=0$ or $\PP_m(S)=1$.
\end{theorem}
That (1) $\Rightarrow$ (2) follows from Corollary \ref{2A},
and that (4) $\Rightarrow$ (1) is immediate. The other
implications (2) $\Rightarrow$ (3) $\Rightarrow$ (4)
can be shown in exacly the same way as the proof of
Theorem 3.1 of [F].
\bigskip
The diffusion operator $D_t:L^2(M,m)\to L^2(M,m)$ ($m$ a harmonic measure)
is not self-adjoint unless $m$ is completely invariant.
Its adjoint $D_t^*$ is first considered in [K].
Let $h$ be the characteristic harmonic function defined on the
holonomy cover $\hat L$ of $m$-a.a.\ leaf $L$.
Denote by $\hat p_t$ the heat kernel on $\hat L$. We have
$$
p_t(x,y)=\sum_{\hat y}\hat p_t(\hat x, \hat y),
$$
where the sum is taken for all the points $\hat y$
over $y$, and independent of the choice of $\hat x$ over $x$.
We shall summerize well known properties of the heat kernel $\hat p_t$
on $\hat L$ which follows from the bounded geometry of $\hat L$.
\begin{lemma}\label{last}
For any harmonic function $g:\hat L\to\R$, we have
\begin{eqnarray*}
&g(\hat x)=\int_{\hat L}g(\hat y)\hat p_t(\hat x,\hat y)d{\rm vol}(\hat y)
\ \mbox{ and }\\
&\hat p_{t+t'}(\hat x,\hat z)=\int_{\hat L}
\hat p_t(\hat x,\hat y)\hat p_{t'}(\hat y,\hat z)d{\rm vol}(\hat y).
\end{eqnarray*}
\end{lemma}
Now define a new heat kernel on $\hat L$ by
$$
\hat q_t(\hat x,\hat y)
=\frac{h(\hat y)}{h(\hat x)}\hat p_t(
\hat x,\hat y).
$$
The following lemma follows immediately from Lemma \ref{last}.
\begin{lemma} \label{6}
We have
\begin{eqnarray} \label{e3}
&\int_{\hat L}\hat q_t(\hat x,\hat y)\,d{\rm vol}(\hat y)=1
\ \mbox{ and }\\
&\hat q_{t+t'}(\hat x,\hat z)=\int_{\hat L}
\hat q_t(\hat x,\hat y)\hat q_{t'}(\hat y,\hat z) \label{e4}
\,d{\rm vol}(\hat y).
\end{eqnarray}
\end{lemma}
Define a heat kernel $q_t$ on the leaf $L$ by
$$
q_t(x,y)=\sum_{\hat y}\hat q_t(\hat x, \hat y).
$$
\begin{theorem} \label{7}
The dual operator $D_t^*$ is expressed for any $f\in L^2(M,m)$ as
$$
(D_t^*f)(x)=\int_Lq_t(x,y)f(y)\,d{\rm vol}(y),
$$
where $L$ is the leaf through $x$.
\end{theorem}
Although this theorem is known to Vadim Kaimanovich, we shall include
a proof, since there seems to be none in the literature.
Let $\GG$ denote the holonomy groupoid associated to the lamination
$\LL$, i.e., $\GG$ is the space of leafwise paths modulo same end points
and identical holonomy germs. Denote by $r,\,s:\GG\to M$ the range
and the source maps. The fiber $s^{-1}(x)$ is homeomorphic to
the holonomy cover of the leaf through $x$, and the corresponging volume form
of $s^{-1}(x)$ is denoted by ${\rm vol}_x$. Integrating these forms
(seen as measures) over the harmonic measure $m$ of $M$,
we get a measure $m_\GG$ on $\GG$. That is,
$$m_\GG=\int_M {\rm vol}_x dm(x).$$
Likewise we define a measure ${\rm vol}^y$ on
$r^{-1}(y)$. Define a function $\varphi:\GG\to\R$ by
$\varphi([\gamma])=h(\gamma(1))/h(\gamma(0))$, where $h$ is the
characteristic harmonic function which is defined on the holonomy
cover of $m$-a.a.\ leaf. The function $\varphi$ is well defined by
Proposition \ref{4} (2). Denote by $J:\GG\to\GG$ the inverse map.
\begin{lemma}
We have $Jm_\GG=\varphi\cdot m_\GG$.
\end{lemma}
\bd
For an arbitrary $[\gamma]\in\GG$,
choose a neighbourhood $U\times V\times Z$ of $[\gamma]$ in $\GG$, where
$U\times Z$ is a local chart containing $\gamma(0)$
so that the holonomy along $\gamma$ is defined on $Z$ and $V$
is a leafwise neighbourhood of $\gamma(1)$.
Changing the notations slightly, we consider $U$ (resp.\ $V$)
to be a neighbourhood of $\tilde \gamma(0)$ (resp.\ $\tilde \gamma(1)$)
in the universal cover $\tilde L$ of the leaf $L$, where $\tilde \gamma$
is a lift of $\gamma$
to $\tilde L$.
Choosing $Z$ smaller if necessary, we may assume
that there is a precompact simply
connected open
set $W$ of $\tilde L$ such that $U\cup V\cup\tilde\gamma\subset W$ and that
there is a lamination preserving embedding of $W\times Z$ into $M$.
Then by Theorem \ref{2},
$$m\vert_{W\times Z}=\int_Zh\,{\rm vol}\,d\nu
$$
for a leafwise
harmonic function $h$ and a measure $\nu$ on $Z$.
For $(u,v,z)\in U\times V\times Z\subset\GG$,
denote $s(u,v,z)=(u,z)=x$ and $r(u,v,z)=(v,z)=y$.
Restricted to $U\times V\times Z$, ${\rm vol}_x$ is
the volume form on $u\times V\times z$ and ${\rm vol}^y$
on $U\times v\times z$.
On $U\times V\times Z$ we have
$$
m_\GG=\int_Z {\rm vol}_x\cdot h(x){\rm vol}^y\,d\nu.$$
On the other hand on $V\times U\times Z$,
$$
Jm_\GG=\int_Z {\rm vol}_y\cdot h(y){\rm vol}^x\,d\nu
=\int_Z \varphi\cdot {\rm vol}_y\cdot h(x){\rm vol}^x\,d\nu
=\varphi\cdot m_{\GG},$$
showing the lemma.
\qed
\smallskip
\begin{remark}
The measure $m_\GG$ is defined not only for a harmonic measure, but
also for any probability measure $m$ on $M$. It is interesting
to remark that the leafwise smoothness (Proposition \ref{3}) of $m$ is
equivalent to a basic notion in measured groupoids,
the equivalence of $Jm_\GG$ with $m_\GG$ [AR].
\end{remark}
\smallskip
\noindent
{\sc Proof of Theorem \ref{7}.}
The Riemannian heat kernel on the holonomy cover of the leaf yields
a function $\check p_t$ on $\GG$ by
$$\check p_t([\gamma])=\hat p_t(\gamma(0),\gamma(1)).$$
Notice that $\check p_t\circ J=\check p_t$.
Likewise
a function $\check q_t$ is defined from $\hat q_t$. They satisfy
$\check q_t=\varphi\check p_t$.
Clearly we have
$$
(D_tf)(x)=\int_{s^{-1}(x)}\check p_t\, f\circ r \, d{\rm vol}_x.$$
Thus
\begin{eqnarray*}
&\langle D_tf,g\rangle=\int_M(\int_{s^{-1}(x)}
\check p_t\, f\circ r \, d{\rm vol}_x)g(x)dm(x)
=
\int_\GG \check p_t\,f\circ r\,g\circ s\, dm_\GG\\
&=\int_\GG \check p_t\, f\circ s\, g\circ r\, \varphi\,dm_\GG
=\int_\GG \check q_t\, g\circ r\, f\circ s\, dm_\GG\\
&=\int_M(\int_{s^{-1}(x)}\check q_t\, g\circ r\, d{\rm vol}_x)
f(x)dm(x)
=\langle f,D_t^*g\rangle.
\end{eqnarray*}
Therefore we have
$$
(D_t^*g)(x)=\int_{s^{-1}(x)}\check q_t\, g\circ r\, d{\rm vol}_x
=\int_L q_t(x,y)g(y)\, d{\rm vol}(y),$$
completing the proof.
\qed
\medskip
Now let us define the reverse process.
First of all extend the new heat kernel $q_t$ to $M\times M$,
by putting $q_t(x,y)=0$ unless $x$ and $y$ lie on the
same leaf.
Let $\Omega_-$ be the
space of continuous leafwise paths $\omega$ from $(-\infty,0]$ to $M$,
with the random variable $X_{-t}:\Omega_-\to M$ defined
by $X_{-t}(\omega)=\omega(-t)$ ($t\geq0$).
For $x\in M$, define the Wiener measure $W^x_-$ on $\Omega_-$
using the kernel $q_t$, that is, for example for
$0<t_1< t_2$ and for any Borel sets $B_1$ and $B_2$ of $M$,
$$
W^x_-\{X_{-t_2}\in B_2, X_{-t_1}\in B_1\}
=\int_{B_2}\int_{B_1} q_{t_1}(x,y)q_{t_2-t_1}(y,z)\,d{\rm vol}(y)\,d{\rm vol}(z).$$
Lemma \ref{6} implies that $W^x_-$ is a probability
measure, a probability because of
(\ref{e3}), the dropping condition guaranteed by (\ref{e4}).
The kernel $q_t$ clearly satisfy the normal estimate of Cheng, Li
and Yau ([CLY]) since the ratio to the Riemannian heat kernel is
controlled by the Harnack inequality;
the logarithm of any positive harmonic
function defined on the holonomy cover of any leaf of $\LL$ is uniformly
Lipschitz (due to the uniform boundedness of geometry of leaves).
Therefore the reverse Wiener measure
$W^x_-$ is concentrated on the set of continuous paths.
Moreover it is concentrated on the subspace $\Omega_{-,x}=X_{0}^{-1}(x)$.
Now let $\overline \Omega$ be the space of biinfinite continuous leafwise paths
$\omega:\R\to M$. Denote the like defined random variable
by the same letter $X_t:\overline\Omega\to M$ for $t\in\R$.
Also denote $\overline\Omega_x=X_0^{-1}(x)$.
Then by the natural identification of $\Omega_{-,x}\times\Omega_{x}$
with $\overline\Omega_x$,
the product measure $W_-^x\times W^x$ is considered to be
a measure on $\overline \Omega_x$, or on $\overline \Omega$.
Define a probability measure
$\overline\PP_m$ on
$\overline \Omega$
by
$$\overline \PP_m=\int_MW^x_-\times W^x\, dm(x).
$$
Denote its
expectation by $\overline\EE_m$.
Let $\theta_t:\overline\Omega\to\overline\Omega$ be the shift map.
\begin{proposition} \label{8}
The shift map $\theta_t:\overline\Omega\to\overline\Omega$ preserves the measure
$\overline\PP_m$.
\end{proposition}
\bd
We shall raise one example of computation.
$$
\overline\PP_m\{X_{-t}\in B, X_{t'}\in B'\}
=\int_M dm(x)\int_B q_t(x,y)dy\int_{B'}p_{t'}(x,z)dz
$$
$$
=\langle D_t^*\chi_B, D_{t'}\chi_{B'}\rangle_m
=\langle \chi_B, D_tD_{t'}\chi_{B'}\rangle
=\overline \PP_m\{X_0\in B, X_{t+t'}\in B'\}.$$
\qed
\begin{theorem} \label{9}
If $m$ is an ergodic harmonic measure, then $\overline \PP_m$ is
ergodic with respect to the flow $\{\theta_t\}$.
\end{theorem}
Before starting the proof, we
recall the definition of conditional expectations.
Denote by $\overline\FF$ the $\sigma$-algebra formed by
the $\overline\PP_m$-measurable subsets.
For $t\in\R$, let $\overline\FF_t$ be the minimal complete $\sigma$-algebra
for which the map $X_s$ is measurable for any $s\geq t$.
For example, in order to understand $\overline\FF_0$,
consider the measurable partition of $\overline\Omega$
defined by the natural projection
$\pi:\overline\Omega\to\Omega$.
Then $\overline\FF_0$ consists of measurable subsets saturated
by this partition.
A function $F$ is $\overline\FF_0$-measurable if
and only if there is a measurable function $H$ on $\Omega$
such that $F=H\circ\pi$.
For any integrable function $F:\overline\Omega\to\R$,
denote by $\overline\EE_m\,[F\mid\overline\FF_t]$ the conditional expectation
with respect to $\overline\FF_t$. This is a unique $\overline\FF_t$-measurable
function on $\overline\Omega$ such that for any
bounded $\overline\FF_t$-measurable
function $G$,
$$\overline\EE_m\,[G\,\overline\EE_m\,[F\mid\overline\FF_t]\,]=
\overline\EE_m\,[GF].
$$
One word of explanation for the geometer readers.
$\overline\FF_t$ defines a measurable partition of $\overline\Omega$:
almost all classes of the partition admit the conditional probability measure.
Integrating $F$ by the conditional probability measure we obtain a measurable function
on the quotient space. But it is customary, more convenient, to
consider it to be a
$\overline\FF_t$-measurable function $\overline\EE_m\,[F\mid\overline\FF_t]$
defined on the total space $\overline \Omega$.
\smallskip
\noindent
{\sc Proof of Theorem \ref{9}.}
For an integrable function $F$ on $\overline\Omega$ define the
Birkhoff average $\BB F$ by
$$
\BB F=\lim_{t\to\infty}\frac{1}{t}\int_0^tF\circ \theta_s ds.
$$
By the ergodic theorem, the operator $\BB$ is a well defined contraction
on $L^1(\overline\Omega,\overline\PP_m)$, which is $\theta_t$-invariant.
Since by Theorem \ref{5}, $\theta_t$ is ergodic in $(\Omega,\PP_m)$,
the Birkhoff average $\BB F$ is constant if $F$ is $\overline\FF_0$-measurable.
Moreover this holds for any $\overline\FF_{-t}$-measurable function $F$
for any $t$, since then $F\circ \theta_{t}$ is $\overline\FF_0$-measurable
and $\BB F=\BB(F\circ\theta_{t})$.
For any bounded $\overline \FF$-measurable function $F$,
the $\overline\FF_{-n}$-measurable function
$F_{-n}=\overline\EE_m\,[F\mid \overline\FF_{-n}]$ converges to $F$ pointwise,
by the martingale convergence theorem ([O] Appendix C).
Thus we have $\BB F_{-n}\to \BB F$, and since $\BB F_{-n}$ is constant,
the function $\BB F$ is also constant, showing the ergodicity.
\qed
\bigskip
Applying the Birkhoff theorem to $f\circ X_0:\overline \Omega\to\R$
for a continuous function $f:M\to \R$ by virtue of Theorem \ref{9},
we have $\overline\PP_m$-almost surely
$$
\lim_{t\to\infty}\frac{1}{t}\int_0^tf(X_s) ds
=\lim_{t\to\infty}\frac{1}{t}\int_{-t}^0f(X_s) ds
=m(f).
$$
Equivalently, denoting the Dirac mass by $\delta_.$, we have
$\overline\PP_m$-almost surely
\begin{equation} \label{e21}
\lim_{t\to\infty}\frac{1}{t}\int_0^t\delta_{X_s} ds
=\lim_{t\to\infty}\frac{1}{t}\int_{-t}^0\delta_{X_s} ds
=m,
\end{equation}
where the limit is taken in the space of the probability
measures on $M$ with the ${\rm weak}^*$ topology.
\bigskip
Finally let us define an exponent for the biinfinite Brownian motion.
Assume $m$ is an ergodic harmonic measure of $(M,\LL, g)$ and $h$
the characteristic harmonic function of $m$. Given $\omega\in \overline\Omega$ and a positive number $t$,
the ratio $h(X_t(\omega))/h(X_0(\omega))$
is well defined by Proposition \ref{4}, since
a path from $X_0(\omega)$ to $X_t(\omega)$ is specified by $\omega$.
Define a random variable $A_t:\overline\Omega\to\R$ by
$$A_t=\log h(X_t)-\log h(X_0).$$
Let us show that $A_t\in L^1(\overline \Omega,\overline\PP_m)$.
Denote the expectation of $W^x\times W_-^x$ by $\overline E^x$. Since
$A_t$ is $\FF_0$-measurable, we have $\overline E^x[A_t]=E^x[A_t]$, where $E^x$
is the expectation of $W^x$ defined before.
By the Harnack inequality
$$
E^x[\abs{A_t}]\leq C_1 E^x[d(X_0,X_t)] \leq C_2 t,
$$
where $d$ is the leafwise distance on the
universal cover of the leaf induced from $g$. The last inequality
follows from the bounded geometry of the leaf.
Thus we have
$$
\overline\EE_m\,[\abs{A_t}]=\int_ME^x[\abs{A_t}]dm(x)\leq
C_2t,
$$
showing that $A_t\in L^1(\overline \Omega, \overline \PP_m)$.
Now $A_t$ satisfies
\begin{equation} \label{e2}
A_{t+t'}=A_t+A_{t'}\circ \theta_t.
\end{equation}
This shows that $\overline\EE_m\,[A_t]$ is additive in $t$. Moreover it is
continuous in $t$ at $t=0$, since $E^x[d(X_0,X_t)]\to0$ as $t\to 0$.
That is, $\overline\EE_m[A_t]=-\lambda t$ for some number $\lambda$.
\begin{proposition} \label{10}
We have
$\lim_{t\to\infty}(1/t)A_t=-\lambda$ almost surely, and
$\lambda\geq0$; furthermore $\lambda>0$ unless
$m$ is completely invariant.
\end{proposition}
\bd
The first statement follows from (\ref{e2}) by the Birkhoff ergodic
theorem.
To show $\lambda\geq0$ notice that
$$\int_ME^x[A_t]dm(x)=\overline\EE_m\,[A_t]=-\lambda t.$$
The expectation $E^x[A_t]$ can be computed
upstairs on the holonomy cover. Let $\hat x$ be a lift of $x$
and
$\hat X_t(\omega)$ the lift of $X_t(\omega)$ starting at $\hat x$
for $\omega\in\Omega_x$.
Then
$$
E^x[A_t]=E^{x}[\log h(\hat X_t)]-\log h(\hat x)
$$
$$
\leq \log E^{x}[h(\hat X_t)]-\log h(\hat x)
= \log (\hat D_th)(\hat x)-\log h(\hat x)=0,$$
where $\hat D_t$ is the diffusion operator on the holonomy cover.
The inequality follows from the concavity of $\log$,
and the last equality from the harmonicity of $h$,
showing $\lambda\geq 0$.
For the last statement, notice that $\lambda=0$
implies that for fixed $t$, $h(\hat X_t)$ is constant
$W^{ x}$-almost surely. This shows that $h$ is constant for the
holonomy cover of $m$-a.a.\ leaf, completing the proof.
\qed
\medskip
For $-t<0$ define a random variable $A_{-t}:\overline \Omega\to\R$ by
$$A_{-t}=\log h(X_{-t})-\log h(X_0).$$
It satisfies
\begin{equation} \label{e5}
A_{-t-t'}=A_{-t}+A_{-t'}\circ \theta_{-t}.
\end{equation}
Clearly $\overline\EE_m[A_{-t}]=\lambda$, and again by the
Birkhoff ergodic theorem we have from (\ref{e5}):
\begin{proposition} \label{11}
$\overline\PP_m$-almost surely,
$\lim_{t\to\infty}(1/t)A_{-t}=\lambda$.
\end{proposition}
Propositions \ref{10} and \ref{11} implies
that for $m$-a.a. point $x$, we have
$W^x\times W^x_-$-almost surely
$$\lim_{t\to\infty}(1/t)A_{t}=-\lambda\ \ {\rm and }\
\lim_{t\to\infty}(1/t)A_{-t}=\lambda,
$$
showing:
\begin{theorem} \label{12}
For a non completely invariant ergodic harmonic measure,
the characteristic harmonic function is unbounded on
the holonomy cover of $m$-a.a.\ leaf.
\end{theorem}
\section{Hyperbolic laminations}
Henceforth in this paper
we only consider a compact hyperbolic
$C^2$ lamination $(M,\LL,g)$, i.e.,
we assume throughout that the leafwise metric $g$ has constant curvature $-1$,
and denote the dimension of leaves by $d$+$1$.
Let $m$ be an ergodic harmonic measure for $\LL$. The universal cover
of $m$-a.a.\ leaf $L$ is identified with the simply connected complete hyperbolic
space $\DD$, and the characteristic harmonic function $h$
of $m$ is defined on $\DD$. Choose a base point $\tilde x\in\DD$
and assume $h(\tilde x)=1$.
For any point
$\xi$ of the ideal boundary $\boundary$, let $k_\xi$
denote the minimal positive harmonic function on $\DD$
corresponding to $\xi$
normalized to take value 1 at $\tilde x$.
In other words, $k_\xi=\exp(-dB_\xi)$, where
$B_\xi$ is the Buseman function corresponding to $\xi$ such
that $B_\xi(\tilde x)=0$.
Then there is
a unique probability measure $\mu_{\tilde x}$ on $\boundary$ such that
\begin{equation} \label{e8}
h=\int_{\boundary} k_{\xi} d\mu_{\tilde x}(\xi).
\end{equation}
See [AS] for details and related topics.
Although the measure $\mu_{\tilde x}$ depends on the choice of
the point $\tilde x$, its equivalence class $[\mu_L]$ is an invariant
of the leaf $L$. Here two measures $\mu_1$ and $\mu_2$ on $\boundary$
are said to be equivalent if for any Borel subset $B$ of $\boundary$,
$\mu_1(B)=0$ if and only if $\mu_2(B)=0$.
In fact, for another point $\tilde y\in\DD$, we have
\begin{equation}\label{e8+}
h/h(\tilde y)=\int_{\boundary} k_{\xi}/k_{\xi}(\tilde y) d\mu_{\tilde y}(\xi).
\end{equation}
The uniqueness of the measure $\mu_{\tilde x}$ implies by (\ref{e8})
and (\ref{e8+}) that
$$\mu_{\tilde x}=(h(\tilde y)/k_\xi(\tilde y))\mu_{\tilde y},
$$
showing that $\mu_{\tilde x}$ and $\mu_{\tilde y}$ differ
by a multiple of a bounded positive function, that is,
they are equivalent.
\begin{theorem} \label{13}
For a non completely invariant
ergodic harmonic measure $m$ on a compact hyperbolic
lamination $(M,\LL,g)$ and for $m$-a.a.\ leaf $L$,
the measure class $[\mu_L]$ is singular to
the Lebesgue measure of $\boundary$.
\end{theorem}
Before starting the proof, we need to study connections
among the probability measures on $\boundary$, positive
harmonic functions on $\DD$ and the Wiener measures.
Denote by $\PPP(\boundary)$ the space of probability measures on
$\boundary$, a compact metrizable convex
space by the weak* topology. Denote by $\PPP\HH$ the space
of the positive harmonic function on $\DD$ taking value 1 at $\tilde x$,
also a compact metrizable convex space
by the compact open topology, (compact thanks to the Harnack inequality).
The map $\varphi_1:\PPP(\boundary)\to\PPP\HH$ defined by
$$
\varphi_1(\mu)=\int_{\boundary}k_\xi\,d\mu(\xi)$$
is an affine homeomorphism.
For any $f\in\PPP\HH$, define a heat kernel $q_t$ on $\DD$ by
$$
q_t(u,v)=\frac{f(v)}{f(u)}p_t(u,v),$$
where $p_t$ is the Riemannian heat kernel and $u$ and $v$
are points of $\DD$. The heat kernel defines a Wiener measure
$W^u_f$ for each point $u\in\DD$. Denote by $\Omega_{\tilde x}$
the space of continuous paths $\omega:[0,\infty)\to \DD$ such
that $\omega(0)=\tilde x$
and by $\PPP(\Omega_{\tilde x})$ the space of probability
measures on $\Omega_{\tilde x}$. Then easy calculation shows that
the map $\varphi_2:\PPP\HH\to\PPP(\Omega_{\tilde x})$ defined by
$$
\varphi_2(f)=W^{\tilde x}_f
$$
is affine. (This is just for the base point $\tilde x$ where
$\PPP\HH$ is normalized.)
Now let $\Omega_{\tilde x}^\infty$ denotes the subspace of
$\Omega_{\tilde x}$ consisting of those paths $\omega$
in $\Omega_{\tilde x}$
such that $\lim_{t\to\infty}\omega(t)$ exists
in $\boundary$. Let us show that for any $f\in\PPP\HH$,
the set $\Omega_{\tilde x}^\infty$ is
$W^{\tilde x}_f$-conull. As is well known, this is
true for $f=k_\xi$ for any $\xi\in\boundary$, but any
measure $W^{\tilde x}_f$ is written as the convex
integration
$$
W^{\tilde x}_f=\int_{\boundary}W^{\tilde x}_{k_\xi}d\mu(\xi)$$
for some $\mu\in\PPP(\boundary)$ since
$\varphi_1$ and $\varphi_2$ are affine, showing the claim in
the general case.
Denoting by $X_\infty:\Omega_{\tilde x}^\infty\to\boundary$
the hitting map, we define an affine map
$\varphi_3:\varphi_2(\PPP\HH)\to\PPP(\boundary)$
by $\varphi_3(W^{\tilde x}_f)=X_\infty W^{\tilde x}_f$.
Then the composite $\varphi_3\circ\varphi_2\circ\varphi_1$ is the identity
on $\PPP(\boundary)$, since this is true for the point masses,
the map $\varphi_3\circ\varphi_2\circ\varphi_1$ is affine,
and any measure in $\PPP(\boundary)$ is a convex integral
of the point masses.
\bigskip
\noindent
{\sc Proof of Theorem \ref{13}.}\ \ Let $m$ be a non
completely invariant ergodic harmonic measure of a compact
hyperbolic lamination $(M,\LL,g)$, and let $\DD$
be the universal cover of $m$-a.a.\ leaf $L$.
A base point $\tilde x\in\DD$ is chosen and the
characteristic harmonic function $h$ normalized at
$\tilde x$ is written as (\ref{e8}) using a probability
measure $\mu_{\tilde x}$. The Wiener measure
$W^{\tilde x}_h$ defined by the characteristic harmonic
function $h$ corresponds to the measure $W^{\tilde x}_-$
of the reverse process in Sect.\ 3. As before denote by
$W^{\tilde x}$ the usual Riemannian Wiener measure.
Then by Propositions \ref{10} and \ref{11}, for an appropriate
choice of $\tilde x$ we
have $W^{\tilde x}$-almost surely
\begin{equation}\label{ee1}
\lim_{t\to\infty}(1/t)\log(h(X_t))=-\lambda,
\end{equation}
while $W^{\tilde x}_h$-almost surely
\begin{equation}\label{ee2}
\lim_{t\to\infty}(1/t)\log(h(X_t))=\lambda,
\end{equation}
where $\lambda$ is the characteristic exponent, positive
in our case.
On one hand, the hitting measure $X_\infty W^{\tilde x}$
of the Riemannian Wiener measure $W^{\tilde x}$ coincides with
the visible measure $\mu_0$ at $\tilde x$, which
is equivalent to the Lebesgue measure. On the other hand, the other hitting
measure $X_\infty W^{\tilde x}_h$ is the measure $\mu_{\tilde x}$.
Thus we have
$$
W^{\tilde x}=\int_{\boundary}W^{\tilde x}_{k_\xi}d\mu_0(\xi)\
\mbox{ and }\
W^{\tilde x}_h=\int_{\boundary}W^{\tilde x}_{k_\xi}d\mu_{\tilde
x}(\xi).
$$
That is, for $\mu_0$-a.a.\ point $\xi$,
$W^{\tilde x}_{k_\xi}$-a.a.\ path satisfies
(\ref{ee1}), while
for $\mu_{\tilde x}$-a.a.\ point $\xi$,
$W^{\tilde x}_{k_\xi}$-a.a.\ path satisfies
(\ref{ee2}),
showing that the two measures $\mu_0$ and $\mu_{\tilde x}$
are mutually singular.
\qed
\section{The leafwise unit tangent bundle of a hyperbolic lamination}
Associated with a compact hyperbolic lamination $(M,\LL,g)$,
there is defined
the leafwise unit tangent bundle $N$ of $\LL$ and the
stable foliation $\HH$ on $N$. The space $M$ is covered by open sets $E_i$
on which the local charts $\varphi_i:E_i\to U_i\times Z_i$
are defined. For a hyperbolic lamination, we can assume that each
$U_i$ is an open (precompact) ball in the hyperbolic space $\DD$
and the transition function $\psi_{ji}=\varphi_j\circ\varphi_i^{-1}$
wherever defined is of the form
\begin{equation} \label{e6}
\psi_{ji}(u,z)=(g(z)u, \beta(z)),
\end{equation}
where $g(z)$ is an element of the Lie group $G$
of the orientation preserving isometries of $\DD$.
The leafwise unit tangent bunde $N$ of $\LL$
is defined from the collection of spaces $T^1(U_i)\times Z_i$
by glueing them using the transition function $\psi_{ji}$
defined by the same expression as (\ref{e6}),
where $T^1(U_i)$ is the unit tangent bundle of $U_i$.
Notice that the tangent bundle $T^1(\DD)$ is $G$-equivariantly identified
with $\DD\times\boundary$ by assigning
to a unit tangent vector $v$ the couple $(\pi(v),v_\infty)$,
where $\pi:T^1(\DD)\rightarrow \DD$ is the bundle projection
and $v_\infty\in\boundary$ is the hitting point of the
geodesic ray whose initial vector is $v$.
Thus a local chart $T^1(U_i)\times Z_i$ is identified with
$U_i\times\boundary \times Z_i$. Then the transition function becomes
$$
\psi_{ji}(u,\xi,z)=(g(z)u, g(z)\xi, \beta(z)).
$$
The plaques of the form $U_i\times\xi\times z$ are incorporated
to a lamination
$\HH$ of $N$, called the {\em stable foliation} of $\LL$.
The canonical projection $p:N\to M$ yields a submersion of a leaf
of $\HH$ onto a leaf of $\LL$, and thus
the leafwise Riemannian metric $g$ of $\LL$ can be pulled up
to a leafwise Riemannian metric $\check g$ of $\HH$,
the triplet $(N,\HH,\check g)$ being a compact hyperbolic
lamination. The leafwise volume form of $\HH$ is again denoted
by ${\rm vol}$.
As before $k_\xi$ denotes the minimal positive harmonic function
associated to the point $\xi\in\boundary$ normalized at the point
$\tilde x$.
\begin{definition}
A harmonic measure $\lambda$ on $N$ is called {\em pointed
harmonic} if for each local chart $U\times\boundary\times Z$,
$\lambda$ disintegrates on a plaque $U\times\xi\times z$
to a probabality measure which is a constant times $k_\xi {\rm vol}$.
\end{definition}
The purpose of this section is to establish a one to one
correspondence between harmonic measures of $\LL$
and pointed harmonic measures of $\HH$.
We begin with a harmonic measure $m$ of $\LL$, and
associate it to a pointed harmonic measure upstairs.
Let $x$ be a point on $m$-a.a.\ leaf $L$ of $\LL$,
and let $\tilde x$ be a lift of $x$ to the universal
cover $\DD$ of $L$. Then a probability measure
$\mu_{\tilde x}$ on $\boundary$ is defined using the characteristic
harmonic function $h$ normalized at $\tilde x$ as in (\ref{e8}).
On the other hand, the unit tangent space $T^1_xL$ is identified
with its lift $T^1_{\tilde x}\DD$, and the latter with
$\boundary$ by the visible map. By these identifications,
the measure $\mu_{\tilde x}$ on $\boundary$ corresponds
to a measure $\mu_x$ on $T^1_xL$, the notation being judtified
by the following lemma.
\begin{lemma} \label{31}
The measure $\mu_x$ is independent of the choice
of a lift $\tilde x$ of $x$.
\end{lemma}
\bd
We have only to prove that if $\gamma$ is a deck transformation
of the covering map $\DD\to L$, then
$\mu_{\gamma \tilde x}=\gamma\mu_{\tilde x}$.
In this proof, we need a refined notation: the minimal
positive harmonic function associated to $\xi\in\boundary$
is denoted by $k_{\xi,\tilde x}$ in order to keep in mind the
point $\tilde x$ where it is normalized.
Clearly we have
$$
k_{\gamma\xi,\gamma\tilde x}\circ\gamma=k_{\xi,\tilde x}.
$$
On the other hand by the definition of $\mu_{\tilde x}$,
the characteristic harmonic function $h$ normalized at $\tilde x$
is given by
$$
h=\int_{\boundary} k_{\xi,\tilde x}d\mu_{\tilde x}(\xi).$$
Therefore
\begin{equation} \label{e31}
h\circ\gamma^{-1}=\int_{\boundary}k_{\gamma\xi,\gamma\tilde
x}d\mu_{\tilde x}(\xi)
=\int_{\boundary}k_{\xi,\gamma\tilde x}d(\gamma\mu_{\tilde x})
(\xi).
\end{equation}
Now by Corollary \ref{4+},
$h\circ\gamma^{-1}$ is a constant multiple of $h$,
normalized at the point
$\gamma \tilde x$. Therefore we have:
\begin{equation} \label{e31a}
h\circ\gamma^{-1}=\int_{\boundary}k_{\xi,\gamma\tilde
x}d\mu_{\gamma\tilde x}(\xi).
\end{equation}
Comparing (\ref{e31}) with (\ref{e31a}),
the uniqueness of the probability measure
shows that $\mu_{\gamma\tilde x}=\gamma\mu_{\tilde x}$.
\qed
\bigskip
The inclusion $T_x^1\LL\hookrightarrow N$ induces
a map from $\PPP(T_x^1\LL)$ to $\PPP (N)$
among the spaces of the probability measures.
The image of $\mu_x$ by this map is also denoted by the same
letter, by abuse of notations.
Recall that if $(X,\mu)$ is a measured space and $(Z,\mathcal B)$ is a
Borel space, then a map $\psi:X\to Z$ is called measurable
if for any $B\in\mathcal B$, $\psi^{-1}(B)$ is a measurable set.
Of course this depends only on the equivalence class of the measure $\mu$.
If $Z=\PPP(Y)$, the space of the probability measures of
a compact metric space $Y$, then $\psi:X\to\PPP(Y)$ is
said to be measurable if it is measurable with respect to
the Borel structure of $\PPP(Y)$ associated with the
weak* topology.
This is equivalent to saying that $x\mapsto\psi(x)(f)$
is measurable for any continuous function $f$ on $Y$.
\begin{lemma} \label{32}
The assignment $M\ni x\mapsto\mu_x\in\PPP(N)$ is measurable with respect to
the harmonic measure $m$.
\end{lemma}
\noindent
{\sc Proof.}
Since for any local chart $U\times Z$ of $\LL$, $U$
is assumed to be a domain in $\DD$, the inclusion map of $U\times Z$ into
$M$ can be extended using leafwise geodesics to a lamination preserving
submersion $\varphi:\DD\times Z\to M$ in such a way that it
is a local isometry on each leaf. The set $\DD\times Z$ is
called a {\em prolonged local chart} of $\LL$.
Associated to it we have a prolonged local chart
$\DD\times\boundary\times Z$ for $\HH$.
By Theorem \ref{2}, the harmonic measure $m$ restricted to a local
chart $U\times Z$ is given by
$$m\vert_{U\times Z}=\int_Zh{\rm vol}\,d\nu,$$
where $h$ is a measurable function defined on $U\times Z$,
harmonic on a plaque $U\times z$ for $\nu$-a.a.\ $z$.
For the prolonged
local chart $\DD\times Z$ let $m\vert_{\DD\times Z}$
be the {\em lift} of $m$ to $\DD\times Z$, i.e., the integral
of the counting measure on the fiber of the submersion
$\DD\times Z\to M$ over $m$.
Then we still have
\begin{equation} \label{e32}
m\vert_{\DD\times Z}=\int_Z h\, {\rm vol} d\nu,
\end{equation}
where $h$ is an obvious extension.
Notice that a slight generalization of Theorem \ref{2} shows that
$h$ is measurable with respect to $m\vert_{\DD\times Z}$.
Denote by $\PPP\HH_u$ the space of positive harmonic functions
taking value 1 at $u\in\DD$. Then there is an affine homeomorphism
of $\PPP\HH_u$ with $\PPP(\boundary)$.
Let $(u,z)\in \DD\times Z$ corresponds to $x\in M$ by the submersion.
The measure $\mu_x=\mu_{(u, z)}$ of $\boundary$ is associated
to the function $h(\cdot,z)/h(u,z)\in\PPP\HH_u$ by the above homeomorphism.
\begin{sublemma} \label{32a}
The assignment
$\DD\times Z\ni(u,z)\mapsto\mu_{(u,z)}\in\PPP(\boundary)$
is measurable with respect to $m\vert_{\DD\times Z}$.
\end{sublemma}
\bd
The measure $m\vert_{\DD\times Z}$ is equivalent to ${\rm vol}\otimes\nu$.
Therefore by Fubini,
there is a vol-conull subset $\DD_*$
such that for any poin $u\in\DD_*$, the set
$\{z\in Z;\,h(u,z)<\alpha\}$
is $\nu$-measurable for any $\alpha\in\mathbb Q$.
It is routine to show then for any any $u\in\DD_*$
and $\alpha\in\R$,
the set $\{z\in Z;\,h(u,z)<\alpha\}$
is $\nu$-measurable.
For any $u\in\DD_*$, the assignment to $z\in Z$ of the
harmonic function
$h(\cdot,z)/h(u,z)$ in $ \mathcal P\HH_u$
is $\nu$-measurable with respect to the $\sigma$-algebra
$\mathcal B(\PPP\HH_u)$ of the
pointwise convergence topology on $\DD_*$.
In fact for any $v\in\DD_*$ and $a>0$, the set
$$
\{z\in Z;\, h(v,z)>ah(u,z)\}=
\bigcup_{\alpha\in\mathbb Q}(\{z\in Z;\,h(v,z)\geq\alpha\}\cap\{
z\in Z; ah(u,z)<\alpha\})$$
is $\nu$-measurable.
The $\sigma$-algebra $\mathcal B(\PPP\HH_u)$
coincides with the $\sigma$-algebra of the compact open topology.
In fact for $(a,b)\subset\R$ and a compact ball $D$ of $\DD$,
the set
$$ \PPP\HH_u(D,(a,b))=\{f\in\PPP\HH_u;\, f(D)\subset(a,b)\}
$$
belongs to $\mathcal B(\PPP\HH_u)$, since
for a subset $\{u_j\}_{j\in\N}\subset \DD_*\cap D$ dense in $D$, we have
\begin{eqnarray*}
&\PPP\HH_u(D,(a,b))=\bigcup_{n\in\N}\{f\in\PPP\HH_u;\, f(D)
\subset[a+n^{-1},b-n^{-1}]\}\\
&=\bigcup_{n\in\N}\bigcap_j\{f\in\PPP\HH_u;\, f(u_j)
\in [a+n^{-1},b-n^{-1}]\},
\end{eqnarray*}
and this subset belongs to $\mathcal B(\PPP\HH_u)$.
A general compact subset $K\subset\DD$ can be written
as the decreasing intersection of finite unions of compact balls $D_n$,
and the like defined set $\PPP\HH_u(K,(a,b))$ also belongs
to $\mathcal B(\PPP\HH_u)$, since
$$\PPP\HH_u(K,(a,b))=\bigcup_n\PPP\HH_u(D_n,(a,b)).$$
The space $\mathcal P\HH_u$ with the compact open topology is homeomorphic
to the space $\PPP(\boundary)$ with the ${\rm weak}^*$ topology.
This shows the $\nu$-measurability
of $\mu_{(u,z)}$
in the variable $z$ for any $u\in\DD_*$. On the other hand
the measure $\mu_{(u,z)}$ is continuous in the variable $u$
for any $z\in Z$.
Let $f:\boundary\to\R$ be an arbitrary continuous function
and fix it for a while.
For any $a\in\R$, define
$$
S(a)=\{(u,z)\in U\times Z;\, \mu_{(u,z)}(f)\geq a\}.$$
The proof of the sublemma
is complete if we show that $S(a)$ is a measurable set.
For any $z\in Z$, define the $z$-slice $S(a)_z\subset U$
by
$$
S(a)\cap(U\times z)=S(a)_z\times z.$$
Similarly define the $u$-slice $S(a)_{u}\subset Z$
for any $u\in U$.
Then $S(a)_{u}$ is $\nu$-measurable for any $u\in\DD_*$, and $S(a)_{z}$ is
closed for any $z\in Z$.
Moreover since $\mu_{(u,z)}(f)$ is a continuous function
of $u$,
$\{S(a_k)_z\}$ forms a (closed) neighbourhood system of $\{S(a)_z\}$
for a sequence $a_k\uparrow a$.
Choose a compact ball $D\subset \DD$ and define
$$S(a)_D=\{z\in Z;\, S(a)_z\cap D\neq\emptyset\}.
$$
Then $S(a)_D$ is $\nu$-measurable.
In fact, for a dense subset $\{u_j\}_{j\in\N}$ of $D\cap\DD_*$,
we have
$$
S(a)_D=\{z\in Z;\, S(a_k)_z\cap\{u_j\}\neq\emptyset,\ \forall k\}
=\bigcap_{k}\bigcup_j S(a_k)_{u_j}.
$$
Now let $\{\mathcal D_i\}$ be a sequence of coverings of $\DD$ by
countably many compact
balls such that ${\rm mesh}(\mathcal D_i)\to0$ as $i\to\infty$.
Define
$$
S(a)^i=\bigcup_{D\in\mathcal D_i}D\times S(a)_D.$$
Then $S(a)^i$ is measurable.
On the other hand, since $S(a)_z$ is closed, we have
$S(a)_z=\bigcap_iS(a)_z^i$. That is, $S(a)=\bigcap_iS(a)^i$
and $S(a)$ is measurable,
completing the proof.
\qed
\bigskip
\noindent
Sublemma \ref{32a} implies in particular for any local chart $U\times Z$,
the assignment
$$
U\times Z\ni (u,z)\mapsto \mu_{(u,z)}\in\PPP(\boundary)
$$
is measurable.
Define a map $\iota_{(u,z)}:\boundary\to U\times\boundary\times Z$
by $\iota_{(u,z)}(\xi)=(u,\xi,z)$.
Consider a map
$$
\psi:U\times Z\times \PPP(\boundary)\to\PPP(U\times \boundary\times Z)
$$
defined by $\psi(u,z,\mu)=\iota_{(u,z)}\mu$.
Consider also a map $\phi:\PPP(U\times \boundary\times Z)\to\PPP(N)$ induced
by the inclusion.
If $(u,z)\in U\times Z$ corresponds to $x\in M$, then
\begin{equation} \label{mu}
\mu_x=(\phi\circ\psi)(u,z,\mu_{(u,z)}).
\end{equation}
The proof of Lemma \ref{32} is complete if we show that the RHS
of (\ref{mu}) is a measurable function of $(u,z)$.
Here we have:
\begin{sublemma} \label{32c}
The map
$\psi:U\times Z\times\PPP(\boundary)\to\PPP(U\times \boundary\times Z)$
is continuous.
\end{sublemma}
\bd
Denote by $C_c(U\times\boundary\times Z)$ the
space of the continuous functions with compact supports.
Consider a product $f\circ p_1\,g\circ p_2$
($f\in C_c(U\times Z)$, $g\in C(\boundary)$,
$p_1:U\times\boundary\times Z\to U\times Z$,
$p_2:U\times\boundary\times Z\to \boundary$, the canonical projections).
Then clearly
$$
\psi(u,v,\mu)(f\circ p_1 g\circ p_2)=f(u,v)\mu(g)$$
is a continuous function of $(u,v,\mu)$.
On the other hand, finite sums of the products $f\circ p_1\,g\circ p_2$
form a dense subset of $C_c(U\times \boundary\times Z)$
in the topology of the uniform convergence on compact sets.
Standard argument shows that $\psi(u,v,\mu)(F)$ is continuous for
any $F\in C_c(U\times\boundary\times Z)$, finishing the proof.
\qed
\bigskip
On the other hand, $\phi$ is obviously continuous.
The RHS of (\ref{mu}) is now shown to be a measurable function
of $(u,z)$,
completing the proof of Lemma \ref{32}.
\qed
\bigskip
Integrating the measurable system of probability measures
$\{\mu_x\}_{x\in M}$
over $m$, we obtain a probability measure $\lambda(m)$ of $N$, called
the {\em canonical lift} of $m$.
\begin{theorem} \label{33}
For any harmonic measure $m$ of $M$,
the canonical lift $\lambda(m)$ is pointed harmonic.
\end{theorem}
\bd
Recall that on a prolonged local chart $\DD\times Z$, the lift of the
harmonic measure $m$ is written as in (\ref{e32}), and
the canonical lift $\lambda(m)$ on the associated prolonged
local chart $\DD\times \boundary\times Z$ disintegrates
on $\DD\times\boundary\times z$ to (a constant multiple of)
\begin{equation} \label{e34}
\int_{\DD}h(u)\mu_u\,d{\rm vol}(u),
\end{equation}
where $\mu_u$ is the probability measure in $\PPP(\boundary)$
determined by the equality
$$
\frac{h(v)}{h(u)}=\int_{\boundary}\frac{k_\xi(v)}{k_\xi(u)}d\mu_u(\xi),
\ \ \forall v\in\DD,$$
where
$k_\xi$ is the minimal harmonic function normalized
at the base point $\tilde x$.
In order to disintegrate further the measure in (\ref{e34})
on $\DD\times\xi\times z$,
we have to transform the measure $\mu_u$ which
depends on $u\in\DD$ to a fixed measure $\mu_{\tilde x}$.
First of all we have
$$
h(v)=\int_{\boundary}h(u)\frac{k_\xi(v)}{k_\xi(u)}d\mu_u(\xi)
=\int_{\boundary}k_\xi(v)\frac{h(u)}{k_\xi(u)}\frac{d\mu_u}
{d\mu_{\tilde x}}(\xi)d\mu_{\tilde x}(\xi).
$$
Hence by the uniqueness of the probability measure, we have
$$
\frac{h(u)}{k_\xi(u)}\frac{d\mu_u}{d\mu_{\tilde x}}(\xi)=1,$$
showing that
$$
\int_{\DD}h(u)\mu_u\,d{\rm vol}(u)
=\int_{\DD}k_\xi(u)\mu_{\tilde x}\,d{\rm vol}(u).$$
This implies that the lift of the measure $\lambda(m)$
disintegrates on $\DD\times \xi\times z$ to a constant
multiple of $k_\xi\,{\rm vol}$, completing the proof.
\qed
\bigskip
Conversely given any pointed harmonic measure on the leafwise
unit tangent space $N$,
its push down is a harmonic measure on
$M$ by Theorem \ref{2}. It is easy to show the following
theorem by analogous computation.
\begin{theorem} \label{15}
A harmonic measure on a compact hyperbolic
lamination $(M,\LL,g)$ corresponds one to one to a
pointed harmonic measure on its leafwise unit
tangent bundle $(N,\HH,\check g)$, by the operations of
taking the canonical lift
and pushing down.
\end{theorem}
\begin{example} \label{16}
If $M$ is a closed oriented hyperbolic surface,
considered as a single leaf lamination, then the unique
harmonic measure $m$ is the (normalized) area form.
The canonical lift $\lambda(m)$ on the
unit tangent bundle $T^1M$ is the (normalized) Haar measure.
\end{example}
\begin{remark}
In case $d=1$ the minimal parabolic subgroup $P$ of $G$
acts on the leafwise tangent bundle $N$
of a compact 2 dimensional hyperbolic lamination from the right
in such a way that the orbit lamination
is the stable foliation $\HH$,
and a probability measure of $N$ is pointed harmonic if and only
if it is invariant by the action of $P$.
Theorem \ref{15} in this case is already obtained in [M]
and [BM] by a somewhat different dynamical method.
For higher dimension we do not have such description of
pointed harmonic measures.
\end{remark}
\section{The dichotomy}
Let $m$ be a harmonic measure on a compact hyperbolic lamination
$(M,\LL,g)$. For $m$-a.a.\ leaf $L$, we have defined a measure
class $[\mu_L]$ on the boundary $\boundary$ of the universal cover
$\DD$ of $L$. In this section we shall prove Theorem \ref{1},
i.e., that for an ergodic harmonic measure $m$, either the support
$K_L={\rm Supp}([\mu_L])$, called the {\em characteristic set}
of $L$, is a singleton for any $m$-a.a.\ leaf,
or is the total space $\boundary$.
The argument closely follows the proof of Proposition 1 [MV],
in which the authors attribute the original idea to Etienne Ghys.
To begin with, let us notice the following fact.
Let $\Gamma$ be the group of deck transformations of the
covering map $\DD\to L$. In the proof of Lemma \ref{31},
we have shown that $\mu_{\gamma \tilde x}=\gamma\mu_{\tilde x}$
for any $\gamma\in\Gamma$.
On the other hand the equivalence class
of the measure $\mu_{\tilde x}$ does not depend on the choice of
the particular point $\tilde x$ from $\DD$, as is explained in the
beginning of Sect.\ 4.
This shows that $\gamma K_L=K_L$.
Given a closed subset $K$ of $\boundary$ which is not a singleton,
the {\em convex hull} of $K$, denoted by $C(K)$,
is the convex hull in $\DD$ of the union of
all the geodesics joining two points
of $K$.
It is a closed convex subset of $\DD$, and the assignment
$K\mapsto C(K)$ is $G$-equivariant, where $G$ is the group
of all the orientation preserving isometries of $\DD$.
Therefore we have:
\begin{lemma} \label{17}
Assume $K_L$ is not a singleton. Then the convex hull $C(K_L)$
of $K_L$, as well as its closed $r$-neighbourhood $N_L(r)$ {\rm ($r>0$)},
is a $\Gamma$-invariant subset of $\DD$.
\end{lemma}
Choose a prolonged local chart $\DD\times Z$, and denote
the characteristic set of the leaf of $\LL$ corresponding
to $\DD\times z$ by $K_z$. Denote by $\CC(\boundary)$
the space of nonempty closed subsets of $\boundary$,
equipped with the $\sigma$-algebra ${\mathcal B}_{\mathcal C}$ of
the Hausdorff topology.
\begin{lemma} \label{18}
The assignment $Z\ni z\mapsto K_z\in\CC(\boundary)$ is $\nu$-measurable with respect to
${\mathcal B}_{\mathcal C}$.
\end{lemma}
\bd
For any open
subset $U$ of $\boundary$, define ${\mathcal C}(\boundary)_U$
to be the open subset of $\CC(\boundary)$ consisting of those
closed sets which intersects $U$. It is well known, easy to show,
that the open sets $\CC(\boundary)_U$
generate the $\sigma$-algebra ${\mathcal B}_{\mathcal C}$.
Therefore it suffices to show that the set
$$
Z_U=\{z\in Z;\, K_z\in \CC(\boundary)_U\}$$
is $\nu$-measurable. Choose a countable family $\{f_i\}$
of nonnegtive continuous functions supported in $U$ such that
the union of their support is $U$, and take a base point
$\tilde x\in\DD_*$, where $\DD_*$ is the subset defined in
the proof of Sublemma \ref{32a}. Then
the set $Z_U$ consists of exactly those points $z$
such that $\mu_{(\tilde x,z)}(f_i)>0$ for some $i$.
The $\nu$-measurable dependence of $\mu_{(\tilde x,z)}$
in the variable $z$
established in the proof of Sublemma \ref{32a} completes the proof.
\qed
\begin{definition}
(1) Let $M_{\rm I}$ be the union of $m$-a.a.\
leaves $L$ such that the characteristic set $K_L$ is a singleton.
(2) Let $M_{\rm II}$ be the union of $m$-a.a.\ leaves $L$ such that
$K_L=\boundary$.
(3) Let $M_{\rm III}=M\setminus(M_{\rm I}\cup M_{\rm II})$.
\end{definition}
Lemma \ref{18} implies that the three subsets are $m$-measurable.
Since they are saturated and the harmonic measure $m$ is
ergodic, only one of them is conull.
Henceforth we assume that $M_{\rm III}$ is conull and deduce a
contradiction,
which is sufficient for the proof of Theorem \ref{1}.
For any $m$-a.a.\ leaf $L$ and for $r>0$,
consider the image of $N_L(r)$ by the covering map $\DD\to L$.
Taking their union for any $m$-a.a.\ leaf $L$,
we get a subset of $M$, denoted by $N(r)$.
\begin{lemma} \label{19}
The subset $N(r)$ is measurable.
\end{lemma}
\bd
Denote by $\CC(\DD\cup\boundary)$ the set of nonempty
closed subsets of the compactification $\DD\cup\boundary$,
equipped with the Hausdorff topology.
Then the map from $\CC(\boundary)$ to $\CC(\DD\cup\boundary)$
which assigns to $K$ the closure
of the $r$-neighbourhood of the convex hull
of $K$ is clearly continuous.
Consider a prolonged local chart $\DD\times Z$ and again let $K_z$
denote the characteristic set of the leaf in $\LL$ which
corresponds to $\DD\times z$. Also denote by $N_z(r)\subset\DD$
the closed $r$-neighbourhood of the convex hull of $K_z$.
Then by the above observation and by Lemma \ref{18},
the map
$$Z\ni z\mapsto N_z(r)\cup K_z\in\CC(\DD\cup\boundary)$$
is measurable. In particular for any
open subset $U$ of $\DD$, the set
$$
\{z\in Z;\, N_z(r)\cap U\neq\emptyset\}
$$
is a measurable
subset of $Z$.
Let us show that the union $N_Z(r)=\bigcup_z N_z(r)\times z$ is a measurable
subset of $\DD\times Z$. Choose a sequence of open coverings
of $\DD$,
${\mathcal U}_1\prec{\mathcal U}_2\prec\cdots$, such that
mesh$({\mathcal U}_i)\to 0$. Define
$$
N_Z(r)^i=\bigcup_{U\in{\mathcal U}_i}U\times\{z;\, N_z(r)
\cap U \neq\emptyset\}.
$$
Then the set $N_Z(r)^i$ is measurable, and hence
$N_Z(r)=\bigcap_iN_Z(r)^i$ is also measurable.
Now the image of $N_Z(r)$ by the submersion of $\DD\times Z$ to $M$
is measurable.
In fact, $N_Z(r)$ is a union of a null set and a
Borel set. The image of a null set is null by the
definition of the lift $m\vert_{\DD\times Z}$ of $m$.
On the other hand the image
of a Borel set by a countable to one Borel map is Borel.
This is a well known fact about standard Borel spaces,
and follows e.g. from [Ke] Corollary 15.2 and [S]
Theorem 1.3.
Now the set $N(r)\subset M$ is a finite union of measurable sets
and is measurable.
\qed
\bigskip
Let us finish the proof of Theorem \ref{1}.
Recall we are assuming that $M_{\rm III}$ is conull in way of
contradiction.
Since
$M=\bigcup_rN(r)$ mod 0, we have $m(N(r))>0$ for some $r$.
By Theorem \ref{5}, the measure $\PP_m$ on the space $\Omega$
of leafwise paths is ergodic with respect to
the shift semiflow $\theta_t$. This means that for $\PP_m$-almost any
path the average time of stay in the set $X_0^{-1}(N(r))$
is equal to $\PP_m(X_0^{-1}(N(r)))=m(N(r))$.
In other words for $m$-a.a.\ $x$, $W^x$-almost
surely we have
\begin{equation} \label{e10}
\lim_{t\to\infty}\frac{1}{t}dt\{s\in[0,t] ;\, X_s\in N(r)\}=m(N(r))>0,
\end{equation}
where $dt$ denotes the Lebesgue measure on $[0,\infty)$.
But by Lemma \ref{17}, the inverse image
$p^{-1}(N(r))$ of the universal covering map $p:\DD\to L$
of $m$-a.a.\ leaf $L$ coincides with
the set $N_L(r)$, the closed $r$-neighbourhood of the
convex hull of the characteristic set $K_L$.
Since $K_L\neq\boundary$, there is a closed
nondegenerate interval $I$ contained in $\boundary\setminus K_L$.
For any point $x$ on $L$, the set of
paths whose lifts hit $I$ has positive $W^x$-measure.
On the other hand for those paths the limit of (\ref{e10}) must be 0,
since there is a neighbourhood of $I$
in $\DD\cup\boundary$ which does not intersect $N(r)$. A contraction.
Theorem \ref{1} is now proved.
\bigskip
\begin{example}
For any harmonic measure $m$ of a compact hyperbolic lamination,
the canonical lift $\lambda(m)$ of $m$, a pointed
harmonic measure of the leafwise
unit tangent bundle, is of type I. Especially
the unique ([G], [DK]) harmonic measure of
the Anosov foliation on
the unit tangent bundle of a closed oriented hyperbolic surface
is of type I.
\end{example}
Ergodic completely invariant measures are typical examples of
harmonic measures of Type II. But there are some more.
An example is in order. Let
$\Sigma=\Gamma\setminus{\mathbb D}^2$, where
$\Gamma<PSL(2,\R)$ is a purely hyperbolic cocompact
Fuchsian group.
Choose any homomorphism $\rho:\Gamma\to {\rm Homeo}(Z)$
to the group of the homeomorphisms of a compact metric space $Z$
which satisfies the following conditions.
(1) The homomorphism $\rho$ is not injective.
(2) There is no $\rho(\Gamma)$-invariant measure on $Z$.
Let $M=\Gamma\setminus({\mathbb D}^2\times Z$), where the action of $\Gamma$
is by deck transformation on the first factor and by $\rho$ on
the second. Then the horizontal lamination
$\{{\mathbb D}^2\times z\}$ on ${\mathbb D}^2\times Z$
induces a lamination $\LL$ on $M$, called the {\em suspension}
of $\rho$.
Let $m$ be any ergodic harmonic measure of $\LL$, and notice that
$m$ is
not completely invariant by (2).
\begin{proposition}
The above ergodic harmonic measure $m$ is of type II.
\end{proposition}
\bd
By Theorem \ref{2}, the harmonic measure $m$
determines the class of a probability measure $\nu$
on $Z$. The measure $\nu$ is quasi-invariant by the action
of $\rho(\Gamma)$.
Assume for contradiction that $m$ is of type I. Then
for the prolonged local chart ${\mathbb D}^2\times Z$,
the charcteristic set $K_z$ ($z\in Z$) is a singleton
for $\nu$-a.a.\ $z$. This yields a measurable map
$k: Z\to\oboundary$, by Lemma \ref{18}. The map $k$
is $\Gamma$-equivariant with respect to $\rho$ and the Fuchsian group
action on $\oboundary$, i.e., we have
$$
k(\rho(\gamma)z)=\gamma k(z)\ \ \mbox{ for all }\ \gamma\in\Gamma,
\ \ \nu{\rm -a.a.}\ \ z\in Z.
$$
The push forward measure
$k\nu$ is kept quasi-invariant by the Fuchsian group, and
in particular its support is an infinite set.
Choose a nontrivial $\gamma\in\Gamma$ from the kernel of $\rho$,
and let $F$ be a Borel fundamental domain of $\gamma$ for
its action on
$\oboundary\setminus{\rm Fix}(\gamma)$. Then we have $\nu(k^{-1}(F))>0$.
On the other hand we have
$$
k^{-1}\gamma F=\rho(\gamma^{-1})k^{-1}\gamma F=k^{-1}F\ \ {\rm mod}\ 0.
$$
Thus we have $\nu(\emptyset)=\nu(k^{-1}F\cap k^{-1}\gamma F)>0$.
A contradiction.
\qed
Finally let us pose some problems.
\begin{question} It is known [K2] that
a compact hyperbolic lamination with a type I ergodic
harmonic measure is an amenable
measured foliation in the sense of [AR].
Is the converse true?
\end{question}
\begin{question}
For an ergodic harmonic measure of type I of a compact
hyperbolic lamination of dimension $d$+$1$, the characteristic
exponent satisfies $\lambda=d^2$. Is it true for type
II measure that $\lambda<d^2$?
\end{question}
\begin{question}
For an injective homomorphism from $\Gamma$ (as above)
to $PSL(2,\R)$ with dense image,
is the harmonic measure
of the suspension foliation type II?
\end{question}
|
\section{Introduction}
\subsection{Gromov--Witten invariants and integrable hierarchies}
\label{sec:introgen}
Gromov--Witten theory deals with the study and the computation of intersection numbers on moduli spaces of holomorphic maps from a source Riemann surface to a compact K\"ahler manifold $X$. Denote by $\overline{\mathcal M}_{g,n}(X,\beta)$ the Kontsevich compactification of the moduli space of degree $\beta \in H_2(X,\mathbb{Z})$ stable maps from $n$-pointed genus $g$ curves to $X$. The Gromov--Witten invariants of $X$ are defined as
\beq
\left\langle \tau_{p_1}(\phi_{\alpha_1}) \dots \tau_{p_n}(\phi_{\alpha_n}) \right\rangle_{g,n,\beta}^X :=
\int_{[\overline{\mathcal M}_{g,n}(X,\beta)]^{\rm vir}} \prod_{i=1}^n \operatorname{ev}^*_i(\phi_{\alpha_i})\psi_i^{p_i},
\eeq
where $[\overline{\mathcal M}_{g,n}(X,\beta)]^{\rm vir}$ is the virtual fundamental
class of $\overline{\mathcal M}_{g,n}(X,\beta)$, $\phi_{\alpha_i}\in H^\bullet(X, \mathbb{C})$ are
arbitrary co-homology classes of $X$, $\operatorname{ev}_i:\overline{\mathcal M}_{g,n}(X,\beta)\to X$
is the evaluation map at the $i^{\rm th}$ marked point, and $\psi_i=c_1(\mathcal{L}_i)$ are the first Chern classes of the universal cotangent line bundles $\mathcal{L}_i$ on $\overline{\mathcal M}_{g,n}(X,\beta)$. When $p_i=0$ for all $i$, these invariants have an interpretation as a ``count'' (in a suitable sense) of holomorphic curves of genus $g$ and degree $\beta$ inside $X$, subject to the constraint of intersecting $n$ generic cycles given by the Poincar\'e duals of the classes $\phi_{\alpha_i}$. \\
We know from examples
\cites{Witten:1990hr, Kontsevich:1992ti, MR2199226},
and have limited general evidence both in the Fano and the Calabi--Yau case
\cites{Eguchi:1997jd, MR1901075, MR2208418, Bershadsky:1993cx},
that Gromov--Witten invariants of a target space $X$ could be subject to a mysterious web of constraints relating them to one another, and a long-standing problem in the subject has been to lift at least part of the mystery. An influential conjecture
stemming from Witten's influential work on two-dimensional topological
gravity \cite{Witten:1990hr} asserts that a full explanation should be
provided by the theory of integrable hierarchies of non-linear PDEs. More
precisely, introduce formal symbols $\epsilon$ and $t^{\alpha,p}$, where $\alpha \in
\{1,\dots, h_X\}$, $h_X :=\dim_\mathbb{C} H^\bullet(X, \mathbb{C})$ and $p \in \mathbb{N}$; the
set $\l\{t^{\alpha,p}\r\}_{\substack{\alpha \in \{1, \dots, h_X\} \\ p\in \mathbb{N}}}$ will be shorthandedly written as $\mathbf{t}$. Moreover let $\phi_1$ correspond to the unity of $H^\bullet(X)$ and denote the formal variable $t^{1,0}$ with $x$. We define the all-genus, full-descendent Gromov--Witten potential of $X$ as the formal power series
\bea
\mathcal{F}^X(\epsilon, \mathbf{t}) = \sum_{g \geq 0} \epsilon^{2g-2}\sum_{ \beta \in
H_2(X,\mathbb{Z})}\sum_{n \geq 0} \sum_{\substack{\alpha_1, \ldots, \alpha_n \\p_1,\ldots,p_n}}
{\prod_{i=1}^n t^{\alpha_i,p_i} \over n!}
\left\langle \tau_{p_1}(\phi_{\alpha_1}) \dots \tau_{p_n}(\phi_{\alpha_n}) \right\rangle_{g,n,\beta}^X.
\label{eq:pot}
\end{eqnarray}
\noindent The Gromov--Witten/Integrable Systems correspondence can then be stated as follows:
\begin{conj}
\label{conj:integrableGW}
Let $\mathcal{F}^X(\epsilon, \mathbf{t})$ denote the all-genus full descendent Gromov--Witten potential of $X$. Then there exists a Hamiltonian integrable hierarchy of PDEs such that $\epsilon^2 \mathcal{F}^X(\epsilon, \mathbf{t})$ is the logarithm of a $\tau$--function associated to one of its solutions. The variables $t^{\alpha,p}$ are identified with times of the hierarchy, and the genus counting variable $\epsilon$ with a perturbative parameter in a small dispersion expansion of the equations.
\end{conj}
\noindent By ``small dispersion expansion'' we mean that, in terms of the basic fields $u_\alpha(\mathbf{t})$
\beq
u_\alpha(\mathbf{t}):=\epsilon^2 \frac{{\partial}^2 \mathcal{F}^X(\epsilon, \mathbf{t})}{{\partial} x {\partial} t^{\alpha,0}},
\label{eq:tau}
\eeq
the equations of the hierarchy should take the form of a formal gradient expansion
\beq
\frac{{\partial} u_\alpha}{{\partial} t^{b,p}}= \sum_{g=0} \epsilon^{2g}
\sum_{\beta=1}^{h_X}A^{[g]}(\mathbf{u}, \mathbf{u}_x, \mathbf{u}_{xx}, \ldots,\mathbf{u}^{(2g+1)} ).
\label{eq:dispexp}
\eeq
In \eqref{eq:dispexp} $A^{[g]}$ are degree $2g+1$ homogeneous polynomials in $\mathbf{u}^{(n)}$, where we have defined
\beq
\deg \frac{{\partial}^n u_\alpha}{{\partial} x^n}=n \quad \forall \alpha.
\label{eq:grading}
\eeq
A constructive proof of this conjecture - i.e., an explicit characterization of the hierarchy associated to the Gromov--Witten theory of a given target space $X$ - would be a
far-reaching result, both in principle and computationally.
However, to find out whether such an integrable structure can be found and effectively described is in general a tough task, and
the catalogue of rigorous and complete answers to this question is restricted to a discouragingly low number of examples:
\ben
\item $X=\mathrm{pt}$, that is, intersection theory on the Deligne--Mumford compactification of the moduli space of curves. The Witten--Kontsevich theorem states \cite{Witten:1990hr,Kontsevich:1992ti} that the KdV hierarchy is the relevant integrable system in this case;
\item $X=\mathbb{P}^1$, in which case the associated system is the extended Toda hierarchy \cite{Eguchi:1994in, MR2092034, MR2199226, Milanov:2006cu, MR2309158};
\item $X=(\mathbb{P}^1)^{\circlearrowleft T\simeq \mathbb{C}^*}$, where $T$ is the canonical torus action on $\mathbb{P}^1$. The relevant hierarchy is a reduction of 2D-Toda \cite{MR2049645, MR2439568, MR2199226}.
\end{enumerate}
For each of the three cases above, a few proposals have been made to extend the correspondence to orbifolds of the form $[X/G]$, where $G$ is a finite group \cite{MR1950944, rossi-2008, MR2433616, pauljohnsonthesis}; the corresponding candidate hierarchies should be reductions of KP (resp. 2D-Toda) for $X=\mathrm{pt}$ (resp. $X=\mathbb{P}^1$). Unfortunately, apart from this very limited bestiary, the goal to have a general constructive proof of Conjecture \ref{conj:integrableGW} appears to be out of reach at the moment. In fact, even adding new examples to the above list seems to be a very challenging problem: the next-to-simplest case of the complex projective plane $\mathbb{P}^2$ is already hard to tackle, and it is as of today unsolved. \\
On the other hand, recent developments \cite{Aganagic:2003db, Aganagic:2003qj}
strongly indicate a natural new arena to push forward the study of the
Gromov--Witten/Integrable Systems correspondence: the local theory of toric
Calabi--Yau threefolds. In this context, physics-inspired dualities
have provided an
impressive quantity of new insights, including conjectural proposals for the solutions of the
non-equivariant theory \cite{Aganagic:2003db,
Bouchard:2007ys} and remarkable connections to other areas of Mathematics: examples include
other moduli space problems in Algebraic Geometry \cite{MR2264664, moop} and quite different subjects like quantum topology
\cite{Gopakumar:1998ki, Ooguri:1999bv, Marino:2001re} and modular forms
\cite{Aganagic:2006wq}. On one hand, it is natural to speculate that the high degree of
solvability of the theory could be {\it explained} by underlying integrable
structures; on the other, such a rich web of mathematical interconnections renders the possibility to elucidate
the role of integrability in this context an even more appealing goal. \\
\subsection{Main results}
\label{sec:introresults}
In this paper we begin to address this problem by studying the integrable structures that govern the equivariant Gromov--Witten theory of Calabi--Yau rank two bundles over the complex projective line - that is, differential neighbourhoods of a (not necessarily isolated) rational curve inside a Calabi--Yau threefold. By Grothendieck's theorem, such bundles split
into a sum of line bundles: $\mathcal{O}_{\mathbb{P}^1}(n_1)\oplus \mathcal{O}_{\mathbb{P}^1}(n_2)$, $n_i \in \mathbb{Z}$; by the Calabi--Yau condition, we must have that $k:=-n_1=n_2+2$. We will denote by $X_k$ the total spaces of these bundles. Moreover, we will consider their equivariant Gromov--Witten theory with respect to a $T\simeq \mathbb{C}^*$ torus action, which covers the trivial action on the base $\mathbb{P}^1$ and rotates the fibers:
\beq
X_k := \mathcal{O}_{\mathbb{P}^1}(-k)\oplus \mathcal{O}_{\mathbb{P}^1}(k-2)^{\circlearrowleft T}, \quad k\in \mathbb{Z}.
\label{eq:Xk}
\eeq
In many cases, we will take $T$ to act with identical (resp. opposite) characters on the two fibers; we will refer to these choices as the {\it diagonal} (resp. {\it anti-diagonal}) case. \\
In spirit, our study we will be very close to the perturbative philosophy of Dubrovin--Zhang \cite{dubrovin-2001} for the non-equivariant Gromov--Witten theory of Fano manifolds with $(p,p)$ co-homology\footnote{This represents the prototypical family of target spaces whose big quantum co-homologies satisfy the technical assumptions necessary for Dubrovin and Zhang's machinery to work, like commutativity and semi-simplicity of the quantum product and a well-defined grading.}.
Let us briefly recall the main lines of their strategy. In their case, the whole hierarchy is constructed according to the following two-step process:
\ben
\item find a closed form description of its genus zero approximation (the {\it Principal Hierarchy});
\item find a reconstruction procedure to incorporate the higher genus corrections.
\end{enumerate}
Step (1) is based on the datum of a Frobenius manifold, that is, a solution of the Witten--Dijkgraaf--Verlinde--Verlinde equations possessing a distinguished dependence on one of its variables (the unity direction) and obeying a quasi-homogeneity condition (existence of the Euler field). Out of these data, it was shown in \cite{Dubrovin:1992dz} how to associate a quasi-linear, non-dispersive Hamiltonian hierarchy and a $\tau$-function coinciding with the genus zero Gromov--Witten partition function. Step (2) is much more involved, and strongly relies on the the existence of a local bi-Hamiltonian structure, as well as on the assumption of semi-simplicity of the quantum product and of Virasoro constraints on the dispersionful $\tau$-function \cite{Dubrovin:1997bv, dubrovin-2001}. \\
We will try to transfer some of the guiding principles of \cite{dubrovin-2001}
to the case at hand. A major obstacle is the fact that equivariant quantum
co-homology rings do not satisfy all axioms of a Frobenius manifold, and in particular the quasi-conformality of the prepotential. Still, the arguments of \cite{Dubrovin:1992dz} show that Step (1) above is {\it almost independent} of the presence of an Euler vector field, the only requirement being that the prepotential be known in closed form. In other words, bi-Hamiltonianity is not required to reconstruct the Principal Hierarchy; the existence of a grading operator is only needed to fix completely a canonical basis of flows. \\
For the case of the local theory of $\mathbb{P}^1$ in the diagonal and anti-diagonal case, and for the resolved conifold with a generic $(\mathbb{C}^*)^2$-fiberwise action, we have complete control on the prepotential both from the $A$-model \cite{Bryan:2004iq} and the $B$-model side \cite{MR2276766}. This will be sufficient for us to construct in a completely explicit way the relevant tree-level hierarchies. \\
From a geometer's point of view, however, the real utility of a clear link with integrable hierarchies resides in the possibility to effectively perform Step (2), namely, to give a complete recipe to solve the all-genus, full-descendent theory in terms of a dispersive deformation of the Principal Hierarchy. This would be particularly valuable for the case at hand, where little is known about possible higher genus relations between descendent invariants. For this second step, however, it looks hopeless to generalize the methods of Dubrovin--Zhang for the construction of the dispersive tail, as the validity of some of their key assumptions, like existence of Virasoro symmetries annihilating the $\tau$-function, is unclear, if not in jeopardy in our case. \\
Still, in one example we can find a way out. It turns out that for the resolved conifold $\mathcal{O}_{\mathbb{P}^1}(-1)\oplus \mathcal{O}_{\mathbb{P}^1}(-1)$ with anti-diagonal action the Principal Hierarchy coincides with the long-wave limit of the so-called Ablowitz--Ladik lattice \cite{MR0377223}. The latter can be regarded as a complexified version of the discretized non-linear Schr\"odinger hierarchy, and appears as a particular reduction of the 2D-Toda hierarchy. Explicit knowledge of a candidate dispersive integrable model allows us to give a full reconstruction of the dispersive flows. It is tempting to speculate that this particular deformation could be the one that verifies Conjecture \ref{conj:integrableGW} in this case.
\begin{conj}
The all-genus, full descendent Gromov--Witten potential of the resolved conifold $X_{1}$ in the equivariantly Calabi--Yau case is the logarithm of a $\tau$-function of the Ablowitz--Ladik hierarchy.
\label{conj:AL}
\end{conj}
If proven, this statement would add a fourth item to the list we presented in
Sec. \ref{sec:introgen}. We have various reasons to believe that this
conjecture is true. First of all, it was shown in \cite{MR2462355} that, for
$2$-component integrable systems like the ones we consider in this paper,
integrability very often breaks down when we turn on dispersive
perturbations. This is for example the case of the generalized Fermi--Pasta--Ulam systems, for which the procedure of discretizing space derivatives never preserves involutivity of the flows except for exponential non-linearities (i.e. for the Toda lattice). Having {\it one} dispersive integrable candidate is an already fortunate circumstance and, if we trust the statement of Conjecture \ref{conj:integrableGW}, it should be taken very seriously. \\ The second, and much more cogent piece of evidence that we provide is given by the following
\begin{thm}
Conjecture \ref{conj:AL} is true for $g\leq 1$.
\label{thm:main}
\end{thm}
The key idea in our proof will be to establish a so-called
``quasi-triviality'' property for the Ablowitz--Ladik hierarchy at the leading
order of the dispersive expansion. Although differing in the way we obtained
it, due to the apparent absence of a second compatible local Poisson bracket for the Ablowitz--Ladik system, our final result comes very close to analogous statements in the bi-Hamiltonian case \cite{Dubrovin:1997bv, dubrovin-2001}.\\ Finally, we will exploit the possibility to reconstruct the dispersive flows, order by order in the parameter $\epsilon$, to give higher genus tests of our proposal. In particular we verify the following non-trivial implication of Conjecture \ref{conj:AL}:
\begin{thm}
Under Assumption \ref{assm:tausymm}, let $\mathcal{F}(\epsilon,\mathbf{t})$ be the Ablowitz--Ladik $\tau$-function which reduces for $\epsilon \to 0$ to the topological $\tau$-function of the Principal Hierarchy of $X_{1}$ with anti-diagonal action. Then its reduction to small phase space at $\mathcal{O}(\epsilon^4)$ coincides with the genus $2$ primary Gromov--Witten potential of the resolved conifold in the equivariantly Calabi--Yau case, possibly up to the degree zero term.
\label{cor:g=2}
\end{thm}
For the reasons that we have outlined at the end of section
\ref{sec:introgen}, we believe that this new example of the
Gromov--Witten/Integrable Systems correspondence could be a good starting point
for new insights in toric Gromov--Witten theory. A partial list of the
questions to be answered include the relationship of our hierarchies with the
physicists' open invariants of toric Calabi--Yau threefolds and the
Eynard--Orantin recursion \cite{Eynard:2007kz, Bouchard:2007ys}, a local mirror
symmetry description of the hierarchies in the framework of spectral curves
and the universal Whitham hierarchy \cite{Krichever:1992qe, Aganagic:2003qj},
the study of the fate of the Virasoro conjecture \cite{Eguchi:1997jd} in the
equivariant case,
multi-parameter generalizations ({\it e.g.} the
``closed topological vertex''), a Kontsevich-like description via random
matrix ensembles, and physical applications for the geometric engineering of
extended $\mathcal{N}=2$ $U(1)$ gauge theories \cite{Marshakov:2006ii} in five dimensions. We plan to return on some of this points in the near future. \\
{\noindent \bf Acknowledgements.} It is a pleasure to thank {\it in primis}
Boris Dubrovin for his many insightful comments and suggestions, that
crucially helped us to solve many puzzles in this project. The present work
would not have been undertaken without his influence. I am moreover grateful
to Vincent Bouchard, Motohico Mulase and Brad Safnuk for inviting me to
present part of this material at the AIM workshop ``Recursion structures in
topological string theory and enumerative geometry'' in Palo Alto; I would
also like to thank the participants for their valuable comments. I am likewise
happy to thank Tom Coates for inviting me to visit Imperial College in
November 2009, as well as for his interest, support, and helpful discussions
on related subjects. I also benefitted from discussions with Guido Carlet,
Renzo Cavalieri and Paolo Rossi. \\
This work was supported by a post-doc fellowship of the Fonds National Suisse (FNS); partial support from a ``Progetto Giovani 2009'' grant of the Gruppo Nazionale per la Fisica Matematica (GNFM) is also acknowledged.
\section{The genus zero theory of local $\mathbb{P}^1$ and integrable hierarchies}
\label{sec:displess}
\subsection{Hamiltonian integrable hierarchies from associativity equations}
\label{sec:disphier}
In this section we sketchily review the general construction of dispersionless Hamiltonian hierarchies from associativity equations. The details can be found in the original literature on the subject \cite{Dubrovin:1992dz, Dubrovin:1994hc}; see also \cite{rossi-2009} for a recent and very readable account of this material. \\
Let $\mathcal{V}$ be a $n$-dimensional vector space over a field $\mathbb{K}$. We will denote by $\mathcal{N} :=\mathcal{L}(S^1, \mathcal{V})=\{\mathbf{u}:S^1 \to\mathcal{V}\}$ the formal loop space of $\mathcal{V}$; the components of the formal maps $\mathbf{u} \in \mathcal{N}$ will often be written as $u^\alpha(x)$, where $x\in S^1$ and $\alpha=1,\dots, n$. $\mathcal{N}$ carries naturally the structure of a linear space over $\mathbb{K}$; a distinguished subspace of its dual $\mathcal{N}^*$ is given by the so called {\it local functionals}
\beq
F[\mathbf{u}] := \int_{S^1} f(x, \mathbf{u},\mathbf{u}_x,\mathbf{u}_{xx},\ldots,\mathbf{u}^{(k)}, \ldots) \mathrm{d} x,
\eeq
where $\mathbf{u}^{(k)}$ denotes the $k^{\rm th}$ $x$-derivative of $\mathbf{u}$. The adjective ``local'' refers to the fact that we require the density $f$ to be a {\it differential polynomial}, i.e., to depend polynomially\footnote{In this formal setting and in absence possibly of a well defined analytic theory of functions when $\mathbb{K}\neq\mathbb{C}$ or $\mathbb{R}$, a non-polynomial functional dependence should be thought of as a non-truncating formal power series expansion in $u^\alpha(x)$.} on $\mathbf{u}^{(k)}$ for $k>0$. The set of local functionals on $\mathcal{N}$ will be called $LF(\mathcal{N})$. We want to define a Hamiltonian infinite dimensional dynamical system on $\mathcal{N}$ via the following data:
\bit
\item a local Poisson bracket
\beq
\l\{ u^\alpha(x), u^\beta(y) \r\} = \sum_{j=0}^m a^{\alpha\beta}_j(\mathbf{u},\mathbf{u}_x,\mathbf{u}_{xx},\ldots,\mathbf{u}^{(n)}, \ldots) \delta^{(j)}(x-y),
\label{eq:parpoisgen}
\eeq
for some integer $m\in \mathbb{N}$ and differential polynomials $a^{\alpha\beta}_{j}$; we have denoted by $\delta^{(j)}(x-y)$ the $j^{\rm th}$ distributional derivative of Dirac's $\delta$-function. By bilinearity and the functional Leibnitz rule, the Poisson bracket of elements $F, G \in \mathcal{N}$ is
\beq
\l\{ F, G\r\} = \int_{S^1 \times S^1} \frac{\delta F}{\delta u^\alpha(x)} \frac{\delta G}{\delta u^\beta(y)} \l\{ u^\alpha(x), u^\beta(y) \r\} \mathrm{d} x \mathrm{d} y.
\eeq
The Poisson structure on $\mathcal{N}$ is said to be of {\it hydrodynamic type} if $a^{\alpha\beta}_j= \delta_{j1}\eta^{\alpha\beta}$ for a constant, symmetric, non-degenerate matrix $\eta^{\alpha\beta}$;
\item Hamiltonian flows on $\mathcal{N}$ generated by Hamiltonians $H[\mathbf{u}]\in LF(\mathcal{N})$ via
\beq
\mathbf{u}_{t} = \{\mathbf{u}, H[\mathbf{u}]\}.
\eeq
\end{itemize}
\begin{defn}
Let $ \l\{ u^\alpha(x), u^\beta(y) \r\}$ be a hydrodynamic Poisson bracket on $\mathcal{N}$
and $H_{\alpha,p}[\mathbf{u}] \in LF(\mathcal{N})$, $\alpha=1,\dots,n$, $p \in \mathbb{N}$, be a countably infinite sequence of independent local functionals
\beq
H_{\alpha,p}[\mathbf{u}]=\int_{S^1}h_{\alpha,p}(\mathbf{u}(x), \mathbf{u}_x(x), \ldots,x)\mathrm{d} x.
\eeq
Then:\\
\ben
\item the equations
\beq
\frac{{\partial} \mathbf{u}}{{\partial} t^{\alpha,p}} = \{\mathbf{u}, H_{\alpha,p}[\mathbf{u}]\}
\label{eq:hierarchygen}
\eeq
are said to make up a {\it Hamiltonian integrable hierarchy of PDEs} if they satisfy the involutivity condition
\beq
\{H_{\alpha,p}[\mathbf{u}], H_{\beta,q}[\mathbf{u}]\}=0 \quad \forall \alpha,\beta,p,q;
\eeq
\item the hierarchy is said to possess a $\tau$-structure it there exists a potential ${\partial}_x\ln \tau$ for the integrability condition
\beq
{\partial}_{t^{b,q}} h_{\alpha,p-1}={\partial}_{t^{\alpha,p}}
h_{\beta,q-1}={\partial}_x{\partial}_{t^{\alpha,p}}{\partial}_{t^{b,q}} \ln \tau.
\eeq
$\tau(\mathbf{u},\mathbf{u}_x,\mathbf{u}_{xx},\ldots,\mathbf{u}^{(n)}, \ldots)$ is called a $\tau${\it-function} of the hierarchy;
\item the hierarchy is said to be {\it dispersionless} if the system \eqref{eq:hierarchygen} is quasi-linear, i.e., if the densities
$h_{\alpha,p}$ do not depend on derivatives $\mathbf{u}^{(k)}$ of the fields for $k\geq 1$.
\end{enumerate}
\end{defn}
It was suggested by Witten \cite{Witten:1990hr} that Conjecture
\ref{conj:integrableGW} should have a description in this framework, with the
Hamiltonian densities $h_{\alpha,p}$ being related to 2-point ``big phase space''
correlators in a topological field theory coupled to gravity. For the genus
zero theory this was formalized in fairly large generality, and in a
completely explicit way, in the work of Dubrovin \cite{Dubrovin:1992dz,
Dubrovin:1994hc}. We will now review it in the case we will be interested in
of the $T\simeq (\mathbb{C}^*)^k$-equivariant Gromov--Witten theory of a K\"ahler target manifold
$X$. We will assume that $T$ acts with compact fixed loci $F$. \\
Take $\mathcal{V}_X:=QH_T^\bullet(X)$ to be the big equivariant quantum co-homology ring of $X$; in this case $\mathbb{K} = \mathbb{C}(\lambda_1, \dots, \lambda_k)$ is the field of fractions of $H_T^\bullet(\mathrm{pt})$. Suppose moreover that $\mathcal{V}_X^{\mathrm{odd}}=0$, and pick a basis $\phi_\alpha$, $\alpha=1,\dots, h_X$ of $\mathcal{V}_X$, where $h_X=\dim_{\mathbb{K}} \mathcal{V}_X$ and $\phi_1 = \mathbf{1}_{\mathcal{V}_X}$; a generic element of $\mathcal{V}_X$ will be written $\mathbf{u}=\sum_\alpha u^\alpha \phi_\alpha$ with $u^\alpha\in \mathbb{K}$. The genus zero primary Gromov--Witten potential of $X$ is a formal analytic function $F_0:\mathcal{V}_X\to \mathbb{K}$,
\beq
F_0(\mathbf{u}) = \sum_{j,d=0}^\infty \frac{1}{j!}\Big\langle \overbrace{\mathbf{u}, \dots, \mathbf{u}}^{j \hbox{ \footnotesize times}} \Big\rangle_{0,j,d}^X = \sum_{j=0}^\infty \sum_{\substack{\alpha_1 \dots \alpha_n}} f_{\alpha_1 \dots \alpha_j} u^{\alpha_1}\dots u^{\alpha_j},
\label{F0series}
\eeq
satisfying:
\ben
\item $\eta_{\alpha\beta}:=\partial_1\partial^2_{\alpha\beta}F_0(\mathbf{u})$ is a nondegenerate, constant symmetric matrix;
\item $F_0$ obeys the following set of third order, non-linear PDEs
\beq
\partial^3_{\alpha\beta\gamma} F_0 \eta^{\gamma\delta} \partial^3_{\delta\epsilon\zeta} F_0=\partial^3_{\alpha\epsilon\gamma} F_0 \eta^{\gamma\delta} \partial^3_{\delta\beta\zeta} F_0.
\label{wdvv}
\eeq
It is intended that indices are raised with the non-degenerate contravariant 2-tensor $\eta^{\alpha\beta}=(\eta^{-1})_{\alpha\beta}$, and we use Einstein's convention to sum over repeated indices.
\end{enumerate}
\begin{rem}
\label{rem:euler}
It should be stressed at this point that, as opposed to the usual definition of a Frobenius manifold \cite{Dubrovin:1994hc}, we do not have a quasi-homogeneity condition obeyed by $F_0$. This is due to the fact that the ground field $\mathbb{K}$ has a non-trivial grading in this case, and therefore the natural Euler operator, keeping track of the equivariant de Rham degree, is not $\mathbb{K}$-linear.
\end{rem}
\noindent Eq. \eqref{wdvv} implies the following fact. Define the 1-parameter family of connections on $T^*\mathcal{V}_X$
\beq
D_z := d + \Gamma,
\eeq
where the Christoffel symbol $\Gamma_\alpha$ in components reads $(\Gamma_\alpha)_\beta^\gamma := z c_{\alpha\beta}^\gamma$ and $z\in\mathbb{K}$. Notice that because of integrability of $c_{\alpha\beta\gamma}$ and \eqref{wdvv} we have
\beq
D_z^2=0 \quad \forall z,
\label{eq:flatcond}
\eeq
that is, the connection is flat. Its horizontal sections $\omega^{(\beta)}_\alpha d u^\alpha = d h^{(\beta)}$, where $(\beta)=1,\dots, h_X$ labels a fundamental set of solutions of \eqref{eq:flatcond}, should come from a basis $f^{(\beta)} \in Fun(\mathcal{V}_X)$ of solutions of the holonomic system of PDEs
\beq
\partial^2_{\alpha \beta} h^{(\delta)} = z c_{\alpha \beta}^\gamma \partial_{\gamma} h^{(\delta)}, \quad \delta=1, \ldots, h_X.
\label{flatfunct}
\eeq
We will call the solutions of \eqref{flatfunct} the {\it flat functions} of $\mathcal{V}_X$. Their duals $h_\alpha(u,z):=\eta_{\alpha\beta} h^{\beta} (u,z)$ can always be normalized such that
\bea
\label{eq:norm1}
h_\alpha(\mathbf{u},0)&=&w_\alpha=\eta_{\alpha\beta}u^\beta, \\
\label{eq:norm2}
{\partial}_\gamma h_\alpha(\mathbf{u},z)\eta^{\gamma\delta} {\partial}_\delta h_\beta(\mathbf{u},-z) &=& \eta_{\alpha\beta}, \\
\label{eq:norm3}
{\partial}_1 h_\alpha(\mathbf{u},z) &=& z h_\alpha(\mathbf{u},z) + \eta_{1\alpha}.
\end{eqnarray}
\begin{rem}
Eqs. \eqref{eq:norm1}-\eqref{eq:norm3} do not fix completely the ambiguity in the choices of the $z$-dependent constants of integration of \eqref{flatfunct}. In the ordinary Frobenius manifold case such ambiguity could be dealt with by imposing additional conditions coming from the existence of the Euler vector field. In the cases we are interested in such a procedure will have to be performed otherwise (see Sec. \ref{sec:conifold}).
\end{rem}
Solutions of WDVV relate to the theory of Hamiltonian dispersionless systems in the following way. Endow the loop space $\mathcal{N}_X := L(S^1, \mathcal{V}_X)$ with the hydrodynamic Poisson bracket
\beq
\l\{ u^\alpha(x), u^\beta(y) \r\} = \eta^{\alpha\beta} \delta'(x-y).
\label{eq:poissbrack}
\eeq
Then the Taylor coefficients of the $z$-expansion of $h_\alpha(\tau; z)$ with respect to $z$,
\beq
h_\alpha(\mathbf{u},z) =: \sum_{z=0}^\infty h_{\alpha,p-1}(\mathbf{u}) z^p,
\label{eq:hap}
\eeq
define dispersionless Hamiltonian densities on $\mathcal{N}_X$. The system of $1^{\rm st}$ order quasi-linear PDEs
\beq
\frac{{\partial} \mathbf{u}}{ {\partial} t^{\alpha,p}} = \l\{\mathbf{u}, \int_{S^1}h_{\alpha,p}(\mathbf{u}(x))\r\} \mathrm{d} x
\label{eq:princhier}
\eeq
will be called the {\it Principal Hierarchy} of $X$. We have the following
\begin{thm}[Dubrovin]
\label{thm:dubr}
The set of Hamiltonians $H_{\alpha,p}=\int_{S^1}h_{\alpha,p} \mathrm{d} x $ mutually Poisson--commute with respect to the Poisson bracket \eqref{eq:poissbrack}. Let $u^\alpha(\mathbf{t})$ solve the system (\ref{eq:princhier}) with boundary condition
\beq
\label{eq:initial}
u_\alpha(\mathbf{t})\bigg|_{\substack{t^{a,p}=0 \\ \hbox{\rm \footnotesize for } p>0}}={\partial}^2_{1,\alpha} F_0(u^1+x,u^2,u^3,\dots..., u^n)
\eeq
and define for all times
\bea
\left\langle\bra\tau_{p}\phi_\alpha \tau_q \phi_{\beta} \right\rangle\ket_0 &=& \frac{1}{2\pi i} \oint \oint \frac{dzdw}{z+w}\l( {\partial}_\gamma h_\alpha(\mathbf{u}(\mathbf{t}),z)\eta^{\gamma\delta} {\partial}_\delta h_\beta(\mathbf{u}(\mathbf{t}),w)- \eta_{\alpha\beta}\r), \nn \\ \\
{\partial}^2_{x,t^{\alpha,0}} \mathcal{F}_0(x+t^{1,0},t^{2,0}, \dots ) &:=& \frac{1}{2} \sum\left\langle\bra\tau_{p}\phi_\alpha \tau_q \phi_{\beta} \right\rangle\ket_0(\mathbf{t}) t^{\alpha,p} t^{\beta,q} + \sum \left\langle\bra\tau_{p}\phi_\alpha \tau_1 \phi_1 \right\rangle\ket_0(\mathbf{t}) t^{\alpha,p} \nn \\ &+& \frac{1}{2} \left\langle\bra\tau_{1}\phi_1 \tau_1 \phi_{1} \right\rangle\ket_0(\mathbf{t}), \\
\langle\langle \phi_{\alpha,p} \phi_{\beta,q}\dots \rangle\rangle_0 &:=& \partial_{t^{\alpha,p}} \partial_{t^{\beta,q}}\dots\partial_{\dots} \mathcal{F}_0(\mathbf{t}).
\label{eq:dercorr}
\end{eqnarray}
Then $\mathcal{F}_0$ is the logarithm of a $\tau$ function for the hierarchy \eqref{eq:princhier}. It moreover satisfies
\beq
\begin{array}{cccr}
\mathcal{F}_0|_{t^{a,p}=0 \hbox{ \rm \scriptsize for } p>0} &=& F_0(t^{a,0}) & \hbox{\rm (reduction to primaries)} \\
\partial_x \mathcal{F}_0 &=& \sum t^{\alpha,p}\partial_{t^{\alpha,p-1}} \mathcal{F}_0 +\frac{1}{2} \eta_{\alpha \beta} t^{\alpha,0} t^{\beta,0} & \hbox{\rm (string equation)}\\
\langle\langle\phi_{\alpha,p} \phi_{\beta,q} \phi_{\gamma,r} \rangle\rangle_0 &=&\langle\langle \phi_{\alpha,p-1} \phi_{\delta,0}\rangle\rangle_0\eta^{\delta \epsilon} \langle\langle \phi_{\epsilon,0} \phi_{\beta,q} \phi_{\gamma,r} \rangle\rangle_0 & \hbox{\rm (genus zero TRRs)}
\end{array}
\label{eq:dubrTRR}
\eeq
\end{thm}
For the purpose of the Gromov--Witten/Integrable Systems correspondence this construction has a number of very attractive features, together with a few weak points. The main virtue of this construction is that it does not depend on the details of $X$, apart from the requirement that $\mathcal{V}_X^{\mathrm{odd}}=0$;
moreover, it provides an explicit construction of the integrable hierarchy starting from primary data, thus yielding a constructive proof of Conjecture \ref{conj:integrableGW} at the leading order in $\epsilon$ (i.e. in the genus zero subsector). However, to make it work we need to have control on $F_0$ in {\it closed} form - no implicit, recursive or up-to-inversion-of-the-mirror-map form will do the job. Any polynomial truncation of (\ref{F0series}) affects dramatically the form of the three-point couplings $c_{\alpha\beta\gamma}$ and therefore the flat functions. In other words, we must know explicitly {\it all} the coefficients $f_{\alpha_1\dots \alpha_j}$ in (\ref{F0series}). This limitation turns out to be very constraining in the context of ordinary Gromov--Witten theory, where it basically reduces the list of viable examples to the cases of $X=\mathrm{pt}$ and $X=\mathbb{P}^1$ we mentioned in Sec. \ref{sec:introgen}. However, since the construction does not depend on the existence of an Euler vector field, we might expect to find new examples in the context of equivariant Gromov--Witten theory. Indeed, as we are going to argue, the local theory of rational curves inside Calabi--Yau threefolds evades this limitation in a large number of cases.
\subsection{The resolved conifold}
\label{sec:dispconif}
Let us then consider the target spaces $X_k$ of \eqref{eq:Xk}. We begin with
the rigid case $k=1$, and consider its equivariant theory with respect to a
$T\simeq (\mathbb{C}^*)^2$ fiberwise action rescaling the fibers. Let $H(X_1) :=
H_T^\bullet(X_1) \simeq H^\bullet(F_1) \otimes \mathbb{C}(\lambda_1,
\lambda_2)$ denote the localized $T$-equivariant cohomology of $X_1$ and $F_1
\simeq \mathbb{P}^1$ be the fixed locus of the $T$-action, that is, the null
section of $X_1\to \mathbb{P}^1$. Let moreover $(1,p)$ denote the canonical
basis of $H(X_1)$ (regarded as a free $\mathbb{C}(\lambda_1,
\lambda_2)$-module), where $1$ and $p$ denote respectively the lifts to
$T$-equivariant co-homology of the identity and the K\"ahler class of the base $\mathbb{P}^1$, and write $\mathbf{u} =: v+w p$, i.e. $v:= u^1$, $w:=u^2$ with $v,w \in \mathbb{C}(\lambda_1,\lambda_2)$. We separate the degree zero (``classical'') and positive degree (``quantum'') parts of the genus zero Gromov--Witten potential of $X_1$ as
\beq
F^{X_1}_0(\mathbf{u}) = F^{X_1}_{0, \mathrm{cl}}(\mathbf{u}) + F^{X_1}_{0,\mathrm{qu}}(\mathbf{u}),
\label{treesplit}
\eeq
where
\bea
F^{X_1}_{0, \mathrm{cl}}(\mathbf{u}) &=& \frac{1}{3!}\int_{[\mathbb{P}^1]} \frac{\mathbf{u} \cup \mathbf{u} \cup \mathbf{u}}{e(\mathcal{N}_{X_1/F_1})}, \nn \\
F^{X_1}_{0,\mathrm{qu}}(\mathbf{u}) &=& \sum_{d >0} e^{d w} N^{(1)}_{0,d}, \nn \\
N^{(1)}_{g,d} &=& \int_{[(X_1)_{g,0,d}]^{vir}} 1.
\end{eqnarray}
A special feature of the $(\mathbb{C}^*)^2$-equivariant theory of the resolved conifold is that the invariants $N^{(1)}_{0,d}$ have a closed expression for all $d$ \cite{MR2276766}
given by the Aspinwall-Morrison multi-covering formula \cite{Aspinwall:1991ce}
\beq
N^{(1)}_{0,d}=\frac{1}{d^3}.
\label{eq:aspmorr}
\eeq
$X_1$ then belongs to the list of fortunate cases where a closed form expression for the genus zero Gromov--Witten invariants of all degrees, and therefore for the prepotential, is known in terms of special functions. Explicitly we have
\bea
F^{X_1}_{0, \mathrm{cl}}(v,w) &=& \frac{1}{3!}\int_{[\mathbb{P}^1]} \frac{(v+w p)^3}{(\lambda_1-p)(\lambda_2-p)} \nn \\ &=& \frac{1}{3! \lambda_1 \lambda_2}\int_{[\mathbb{P}^1]} (v+w p)^3\l(1+\l(\frac{1}{\lambda_1}+\frac{1}{\lambda_2}\r)p\r)
\label{clterm1}
\end{eqnarray}
and hence
\beq
F_0^{X_1}= \frac{v^3}{3!}\frac{\lambda_1+\lambda_2}{\lambda_1^2 \lambda_2^2} +\frac{1}{2 \lambda_1 \lambda_2} v^2 w +\operatorname{Li}_3(e^w),
\label{eq:prepconifold}
\eeq
where we have introduced the polylogarithm function
\beq
\operatorname{Li}_j(x) = \sum_{n=1}^\infty \frac{x^n}{n^j}.
\eeq
This is all is necessary to apply the machinery of Sec. \ref{sec:disphier}. For future use, we state the following
\begin{lem} Consider the following solution of WDVV
\beq
F_0= \frac{Pv^3}{3!} +\frac{Qv^2 w}{2} +\operatorname{Li}_3(e^w).
\eeq
Then the general integral of the flatness conditions \eqref{flatfunct} reads
\beq
f(v,w,z,P,Q) = A(w,z,P,Q)\frac{e^{v z}}{z}+ B(z),
\label{eq:genintflat}
\eeq
where
\bea
A(w,z,P,Q) &=&
c_1(z) \, _2F_1\left(-\Delta_+,-\Delta_-;1;e^w\right) \nn \\ &+& c_2(z) \, _2F_1\left(1+\Delta_-,1+\Delta_+;\frac{z P }{Q ^2}+2;1-e^w\right)
\left(1-e^w\right)^{\frac{z P }{Q ^2}+1},
\label{eq:genintA} \\
\Delta_\pm &=&\frac{z \left(P \pm\sqrt{P^2-4 Q
^3}\right) }{2 Q ^2}. \nn
\end{eqnarray}
\end{lem}
\noindent {\it Proof.} The form \eqref{eq:genintflat} follows from \eqref{flatfunct} with $\alpha=v$. Putting $\alpha=\beta=w$ in \eqref{flatfunct} yields a Fuchsian ODE for $A$ as a function of $w$,
\beq
{\partial}^2_{ww}A =\frac{e^w }{1-e^w}\left(\frac{z^2 A(w)}{Q}-\frac{P z {\partial}_w A(w)}{Q^2}\right),
\eeq
whose general integral has the form \eqref{eq:genintA}.\begin{flushright}$\square$\end{flushright}
For the prepotential \eqref{eq:prepconifold} we have
\beq
\Delta_+=z \lambda_1, \quad \Delta_-=z\lambda_2, \quad
\frac{P}{Q^2}=\lambda_1+\lambda_2.
\eeq
Let us fix a normalization of the corresponding flat functions $h^\alpha(v,w;z)$,
\beq
h^\alpha(v,w;z) = A^\alpha(w,z,\lambda_1, \lambda_2)\frac{e^{v z}}{z}+ B^\alpha(z),
\label{eq:unnormff}
\eeq
in order for the flows to satisfy the string axiom and the genus zero TRRs. Eq. \eqref{eq:norm3} fixes $B^\alpha(z)$ to be
\beq
B^\alpha(z)=-\frac{\delta^{v,\alpha}}{z}.
\eeq
To fix completely $A^\alpha(w,z,\alpha,\beta, \lambda)$ we use the fact that it is related \cite{MR1677117} to the fundamental solution $S_{\alpha\beta}$ of the Gauss-Manin system as
\beq
A^\alpha={\partial}_v h^\alpha\Big|_{v=0},
\eeq
\beq
{\partial}_v h^\alpha=:S_0^\alpha=:J^\alpha,
\eeq
that is to say, it corresponds to the $\alpha$-component of the $J$-function at $v=0$. The Coates--Givental theorem \cite{MR2276766} prescribes it to take the form
\beq
\label{eq:Jfunconifold}
J(v=0,w, z, \lambda_1, \lambda_2)=e^{z p \log{q(w)}} \sum_{d\geq 0}\frac{\prod_{m=-d+1}^0\left(-p+m/z+\lambda_1\right)\left(-p+m/z+\lambda_2\right)}{\prod_{m=1}^d\left(p+m/z\right)^2}q(w)^d,
\eeq
where $q(w)$ is the inverse mirror map. We have the following
\begin{prop}
\label{prop:norm}
For the normalized flat functions \eqref{eq:unnormff} we have
\bea
A^\alpha(w,z,\lambda_1, \lambda_2) &=& c^\alpha_1(z,\lambda_1, \lambda_2) \, _2F_1\left(-z \text{$\lambda_1$},-z \text{$\lambda_2$};1;e^w\right) \nn \\ &+& c^\alpha_2(z, \lambda_1,\lambda_2) \, _2F_1\left(z \text{$\lambda_1$}+1,z \text{$\lambda_2$}+1;z (\text{$\lambda_1$}+\text{$\lambda_2$})+2;1-e^w\right) \nn \\ & \times & \left(1-e^w\right)^{z (\text{$\lambda_1$}+\text{$\lambda_2$})+1},
\label{eq:Anorm}
\end{eqnarray}
where:
\bea
\label{eq:cv1}
c^v_1(z,\lambda_1, \lambda_2) &=& 1, \\
c^v_2(z,\lambda_1, \lambda_2) &=& 0, \\
\label{eq:cw1}
c^w_1(z,\lambda_1, \lambda_2) &=& -z \l[\psi ^{(0)} (z \lambda_1+1)+\psi ^{(0)}(z \lambda_2+1)+2 \gamma\r], \\
c^w_2(z,\lambda_1, \lambda_2) &=& -\frac{z \Gamma (z \lambda_1+1) \Gamma (z \lambda_2+1)}{\Gamma (z (\lambda_1+\lambda_2)+2)}.
\label{eq:cw2}
\end{eqnarray}
In \eqref{eq:cw1}, $\gamma$ is the Euler-Mascheroni constant, while $\psi ^{(0)}(x)$ is the polygamma function
\beq
\psi ^{(0)}(z)=\frac{d\log{\Gamma(z)}}{dz}
\eeq
\label{thm:normflat}
\end{prop}
\noindent {\it Proof.} The $\mathcal{O}(z)$ term of the expansion of the $J$-function \eqref{eq:Jfunconifold} is the statement that the mirror map is {\it trivial} in this case
\beq
\log{q}=w \hbox{ (mod } 2\pi i).
\eeq
Let us examine the summand in \eqref{eq:Jfunconifold} above more closely, starting from the numerator. The finite product gives, remembering that $p^2=0$,
\beq
\bary{l}
\prod_{m=-d+1}^0\left(-p+m/z+\lambda_1\right)\left(-p+m/z+\lambda_2\right) = \\
\prod_{m=0}^{d-1}\l[p \left(\frac{2 m}{z}-\lambda_{1}-\lambda_{2}\right)+\frac{m^2-z \lambda_{1} m-z \lambda_{2} m+z^2 \lambda_{1} \lambda_{2}}{z^2}\r]=
\frac{\left(\frac{1}{z^2}\right)^d \Gamma (d-z \lambda_{1}) \Gamma (d-z \lambda_{2})}{\Gamma (-z \lambda_{1}) \Gamma (-z \lambda_{2})} \\
-\l[\frac{\left(\frac{1}{z^2}\right)^d z \Gamma (d-z \lambda_{1}) \Gamma (d-z \lambda_{2}) \left(\psi ^{(0)}(-z \lambda_{1})-\psi ^{(0)}(d-z
\lambda_{1})+\psi ^{(0)}(-z \lambda_{2})-\psi ^{(0)}(d-z \lambda_{2})\right)}{\Gamma (-z \lambda_{1}) \Gamma (-z \lambda_{2})}\r]p,
\eary
\eeq
while for the inverse of the denominator we obtain simply
\beq
\frac{1}{\prod_{m=1}^d\left(p+m/z\right)^2}= \frac{\left(z^2\right)^d}{\Gamma
(d+1)^2}-\frac{2 z \left(z^2\right)^d H_d}{\Gamma (d+1)^2}p,
\eeq
where $H_d$ is the $d^{\rm th}$ harmonic number. For the $v$-component, this means that we should have
\beq
A^v(w,z,\lambda_1,\lambda_2)=
\sum_{d\geq 0} \frac{e^{wd} \Gamma (d-z \lambda_{1}) \Gamma (d-z \lambda_{2})}{\Gamma (d+1)^2 \Gamma (-z \lambda_{1}) \Gamma (-z \lambda_{2})}.
\eeq
Comparison with \eqref{eq:Anorm} sets
\beq
c^{v}_1 =1, \qquad c^{v}_2=0.
\eeq
On the other hand, the component $J^{w}$ of the $J$-function in the direction of the volume form is a series that looks as follows
\bea
J^{w}(0, w,z,\lambda_1,\lambda_2) &=& z w A^v(w,z,\lambda_1,\lambda_2) +
\sum_{d\geq 0} \Bigg[ e^{w d}\frac{z \Gamma (d-z \lambda_{1}) \Gamma (d-z \lambda_{2})}{\Gamma (d+1)^2 \Gamma (-z \lambda_{1}) \Gamma (-z \lambda_{2})} \nn \\ &\times& \left(-2 H_d + \sum_{i=1,2}\l(\psi ^{(0)}(d-z \lambda_{i}) - \psi ^{(0)}(-z \lambda_{i})\r)\right) \Bigg].
\end{eqnarray}
The term proportional to $z w$ term comes from the $e^{zp\log{q}}$ prefactor of the $I$ function. Let us then fix the coefficients $c^{w}_i(z)$ by Taylor-expanding \eqref{eq:Anorm} at $q=\exp{w}=0$. We get an expansion of the form $a \log{q}+b+o(1)$
\bea
A^w(w,z,\lambda_1,\lambda_2) &=&
- \frac{c^w_2(z,\lambda_1,\lambda_2) \Gamma (z
(\lambda_{1}+\lambda_{2})+2)}{\Gamma (z \lambda_{1}+1) \Gamma (z
\lambda_{2}+1)} \log (q) -c^w_1(z,\lambda_1,\lambda_2) \nn \\ &-&
\frac{c^w_2(z,\lambda_1,\lambda_2) \Gamma(2+ z \lambda_1 + z \lambda_2)(
\psi^{(0)}(z \lambda_{1}+1)+\psi^{(0)}(z \lambda_{2}+1)+2 \gamma)}{\Gamma (z
\lambda_{1}+1) \Gamma (z \lambda_{2}+1)} \nn \\ &+& \mathcal{O}(q),
\end{eqnarray}
while from the explicit form of the $J$-function we get
\beq
A^w(w,z, \lambda_1, \lambda_2)=z \log{q} + \mathcal{O}(q).
\eeq
Matching the logarithmic coefficient gives \eqref{eq:cw2}, while the $\mathcal{O}(1)$ term yields \eqref{eq:cw1}. This completely fixes the form of the deformed flat coordinates; it is straightforward to check that the normalization conditions \eqref{eq:norm1}-\eqref{eq:norm3} are satisfied. \begin{flushright}$\square$\end{flushright}
Theorem \ref{thm:dubr} and Proposition \ref{thm:normflat} together complete the construction of the dispersionless hierarchy that governs the genus zero Gromov--Witten theory of the resolved conifold. To see what its first flows look like, take the $z$-expansion of the densities (see \eqref{E-expansion:1}-\eqref{E-expansion:2})
\bea
h_v(v,w,z,\lambda_1,\lambda_2) &=&
\frac{w \lambda_{1} \lambda_{2}+v (\lambda_{1}+\lambda_{2})}{\lambda_{1}^2 \lambda_{2}^2}+\frac{1}{\lambda_{1}^2 \lambda_{2}^2}\Bigg[\frac{1}{2}
(\lambda_{1}+\lambda_{2}) \left(v^2+2 \lambda_{1} \lambda_{2} \text{Li}_2\left(e^w\right)\right) \nn \\ &-& \frac{1}{6} \lambda_{1}
\lambda_{2} \left(-6 v w-6
(\lambda_{1}+\lambda_{2}) \l(\text{Li}_2\left(1-e^w\right)+w \operatorname{Li}_1(w)+\frac{\pi^2}{6}\r)\right)\Bigg] z \nn \\ &+& O\left(z^2\right), \\
h_w(v,w,z,\lambda_1,\lambda_2) &=&
\frac{v}{\lambda_{1} \lambda_{2}}+\left(\frac{v^2}{2 \lambda_{1} \lambda_{2}}+\text{Li}_2\left(e^w\right)\right) z+O\left(z^2\right).
\end{eqnarray}
The first two flows are then
\bea
\frac{{\partial} v}{{\partial} t_{1,0}} &=& v_x, \\
\frac{{\partial} w}{{\partial} t_{1,0}} &=& w_x, \\
\frac{{\partial} v}{{\partial} t_{2,0}} &=& \lambda_1 \lambda_2 \frac{e^w}{1-e^w} w_x, \\
\frac{{\partial} w}{{\partial} t_{2,0}} &=& v_x -(\lambda_1+\lambda_2) \frac{e^w}{1-e^w} w_x.
\label{eq:firstflowsconifold}
\end{eqnarray}
Eliminating $v$ and putting $t:=t_{2,0}$ we obtain the non-linear wave equation
\beq
w_{tt}=\lambda_1 \lambda_2 \l(\frac{e^w}{1-e^w} w_x\r)_x-(\lambda_1+\lambda_2) \l(\frac{e^w}{1-e^w} w_x\r)_t.
\label{eq:nlinconifold}
\eeq
In Sec. \ref{sec:ablo} we will see how this relates, in one notable case, to known examples in the theory of integrable hierarchies.
\label{sec:conifold}
\subsection{The diagonal action}
\label{sec:diagonal}
Let us move on to the general case \eqref{eq:Xk} of $X_k$. In the first place we restrict $T$ to be isomorphic to a one dimensional torus acting diagonally on the two fibers. We adapt, with obvious meaning of symbols, the conventions of Sec. \ref{sec:conifold} for the potentials, the genus zero invariants, and the equivariant co-homology classes of $X_k$, appending an index $k$ and a superscript $di$ (for ``diagonal'') whenever necessary. \\
The choice of a diagonal action is special for two reasons. First, this is the case that corresponds to the invariants defined by Bryan and Pandharipande in \cite{Bryan:2004iq}. Secondarily, it surprinsigly turns out to be a {\it subcase} of the $(\mathbb{C}^*)^2$-equivariant theory of the resolved conifold we treated in the previous section. The quantum tail of the prepotential indeed \cite{MR2276766, Forbes:2006sj} has for all $k$ the Aspinwall-Morrison like form
\beq
N^{(k, {\rm di})}_{0,d}=\frac{1}{d^3}.
\eeq
On the other hand, the classical piece is given by
\bea
F^{X_k, {\rm di}}_{0, \mathrm{cl}}(v,w) &=& \frac{1}{3!}\int_{[\mathbb{P}^1]} \frac{(v+w p)^3}{(\lambda-kp)(\lambda+kp-2p)} \nn \\ &=& \frac{1}{3! \lambda^2}\int_{[\mathbb{P}^1]} (v+w p)^3\l(1+\frac{2p}{\lambda}\r)
\label{cltermk}
\end{eqnarray}
and hence
\beq
F_0^{X_k, {\rm di}}= \frac{1}{3}\l(\frac{v}{\lambda}\r)^3 +\frac{1}{2 \lambda^2} v^2 w +\operatorname{Li}_3(e^w) \qquad \forall k \in \mathbb{Z}.
\label{eq:prepxkdi}
\eeq
Therefore, our results in the previous section apply, {\it a fortiori}, to the theory with generic $k$ and $\lambda_1=\lambda_2$.
\subsection{The anti-diagonal action}
Another case of special interest is given by the reduction to the case of a $T\simeq\mathbb{C}^*$ fiberwise action with {\it opposite} characters on the two fibers. In this case, the equivariant Euler class of $X_k$ is trivial
\beq
e_T(X_k)=0,
\eeq
that is, $X_k$ is equivariantly Calabi--Yau. The notation will follow the same conventions as in the previous two sections, with a superscript $ad$ for ``anti-diagonal'' added whenever needed. \\
It was conjectured in general for toric Calabi--Yau threefolds, and verified explicitly for the case at hand \cite{Bryan:2004iq}, that the invariants in the equivariantly Calabi--Yau case are the ones that most closely make contact with the physics prediction based on topological open/closed duality. In particular the authors of \cite{Bryan:2004iq} could prove the following formula, which could be regarded as a specialization to $X_k$ of the topological vertex formalism of \cite{Aganagic:2003db}:
\begin{thm}[Bryan-Pandharipande]
\label{t:bryanp}
The {\rm fixed-degree $d>0$, all-genus} Gromov - Witten potentials of $X_k^{\rm ad}$ are given by the following sum over partitions
\beq
\label{bryanp}
\sum_{g\geq 0} \epsilon^{2g-2} N^{(k), {\rm ad}}_{g,d} = (-1)^{d(k-1)}\sum_\rho \left(\frac{\mathrm{dim}_Q \rho}{d!}\right)^2 Q^{c_\rho(1-k)}.
\eeq
In (\ref{bryanp}), $\rho$ is a Young diagram (a partition of length
$l(\rho)$), $c_\rho$ is its total content, $Q:=e^{i\epsilon}$, $h(\Box)$ is
the hooklength of a box in $\rho$ and
\beq
\frac{\mathrm{dim}_Q \rho}{d!} = \prod_{\Box \in \rho} \left(2
\sin{\frac{h(\Box)\epsilon}{2}} \right)^{-1}.
\eeq
\end{thm}
\noindent As we stressed in Sections \ref{sec:introresults} and \ref{sec:disphier}, a key point in our analysis is the construction of the hierarchy governing the genus zero theory starting from a {\it closed-form} solution of WDVV. To see this, we should be able to obtain a closed expression for the {\it all-degree, genus zero} invariants starting from \eqref{bryanp}. This is the content of the next
\begin{prop}[\cite{Caporaso:2006gk}]
The quantum part $F^{X_k,{\rm ad}}_{0,qu}(v,w)$ of the $A$--model prepotential of $X_k$ with anti-diagonal action is
\bea
\label{potX1}
F^{X_1,{\rm ad}}_{0, qu}(\tau) &=& \operatorname{Li}_3(e^{w}), \\
\label{potX2}
F^{X_2,{\rm ad}}_{0, qu}(\tau) &=& -\operatorname{Li}_3(e^{w}), \\
\label{potXk}
F^{X_k,{\rm ad}}_{0, qu}(\tau) &=& (-)^{k-1}e^{-w} {}_{n_k+3}F_{n_k+2}\Bigg[1,1,1,1,\frac{1}{n_k},\frac{2}{n_k}, \dots, 1-\frac{1}{n_k}; 2,2,2,2, \frac{1}{n_k-1},\nonumber \\ & & \ldots, 1-\frac{1}{n_k-1}; (-1)^k \left(\frac{n_k}{n_k-1}\right)^{n_k-1} n \exp({w})\Bigg] \qquad (k>2). \nonumber \\
\end{eqnarray}
where $n_k=(k-1)^2$.
\end{prop}
\noindent Eq. \eqref{potX1} is a corollary of \eqref{eq:aspmorr}; the other two expressions were obtained in \cite{Caporaso:2006gk} by an asymptotic analysis in $\epsilon$ of the sum over partitions \cite{Bryan:2004iq} based on saddle-point techniques for a suitably constructed random matrix model. A mirror symmetry confirmation, based on Birkhoff factorization applied to the Coates--Givental twisted $I$-function, was given in \cite{Forbes:2006ab}. \\
To complete the computation of the prepotential we just have to add the degree zero contribution. We get
\beq
F_0^{X_k,{\rm ad}}(v, w) = \frac{2-2k}{3!}\left(\frac{v}{\lambda}\right)^3-\frac{1}{2} w \left(\frac{v}{\lambda}\right)^2+F^{X_k,{\rm ad}}_{0, qu}(w).
\label{eq:pottotXk}
\eeq
The case $k=1$ is obviously a reduction of the case of Sec. \ref{sec:conifold} for $\lambda_1=-\lambda_2=\lambda$; inspection shows moreover that the case $k=2$ coincides with the one of Sec. \ref{sec:diagonal} upon sending $F_0 \to -F_0$. \\
The situation for $k>1$ case is instead radically different. A closed form solution for the flat functions seems too hard to obtain; still the Hamiltonian densities $h_{\alpha,p}$ can be computed and normalized, as we have done before, order by order in $p$. The kind of equations that we find seem totally new: defining the ``Yukawa coupling'' $Y_k(w):= {\partial}^3_{www} F^{X_k,{\rm ad}}_0(w)$ we have
\beq
Y_k(w) =\frac{1}{n_k} -\frac{1}{n_k} {}_{n_k-1}F_{n_k-2}\Bigg[\frac{1}{n_k}, \dots, \frac{n_k-1}{n_k}; \frac{1}{n_k-1}, \ldots, \frac{n_k-2}{n_k-1}; \frac{ (-1)^k n_k^{n_k} e^{-w}}{(n_k-1)^{n_k-1}}\Bigg]
\label{eq:yukk}
\eeq
and we find for instance for the $t:=t_{2,0}$-flow
\bea
\partial_tv(x,t) &=& \left\{v(x,t), H_{w,0} \right\} = Y_k (w) w_x, \\
\partial_tw(x,t) &=& \left\{w(x,t), H_{w,0} \right\}= (v)_x + (2k-2)Y_k(w) w_x,
\end{eqnarray}
which reduces to a wave equation with hypergeometric\footnote{In fact it was shown in \cite{phdthesis-brini} how to give for $k=3$ a purely algebraic expression for the Yukawa \eqref{eq:yukk}; the final result though sheds little more light on the nature of the equation \eqref{firstflowk}.} non-linearity
\beq
(w)_{tt}=\left(Y_k(w) w_x\right)_x+(2k-2)\left(Y_k(w) w_x\right)_t.
\label{firstflowk}
\eeq
\section{The resolved conifold at higher genus and the Ablowitz--Ladik hierarchy}
In this section we address the problem of deforming the hierarchies we constructed in Sec.~\ref{sec:displess} in order to incorporate higher genus corrections, and we will succesfully find a way to do it in the case $k=1$ with anti-diagonal action. After reviewing in Sec. \ref{sec:disppert} the general problem of constructing Hamiltonian integrable perturbations of dispersionless systems, we will exploit the connection of the Principal Hierarchy of $X_1^{\rm ad}$ with a known integrable lattice to construct a candidate dispersive deformation whose $\tau$-function corresponds to higher genus Gromov--Witten potentials. A quasi-triviality property at $\mathcal{O}(\epsilon^2)$ will be established in Sec. \ref{sec:quasitr}, and a $\tau$-structure will be defined at this order and used to prove Theorem \ref{thm:main}. Finally in Sec. \ref{sec:g2} we point out the difficulties and subtleties of the higher genus case, and provide a non-trivial $g=2$ test of Conjecture \ref{conj:AL}.
\subsection{Dispersive perturbations of Hamiltonian systems}
\label{sec:disppert}
In the terminology of Sec.~\ref{sec:introresults}, we have performed Step (1) of the construction of the hierarchies relevant to establish Conjecture \ref{conj:integrableGW} for the local theory of $\mathbb{P}^1$, at least for the case $k=1$ and for the diagonal and anti-diagonal action. A full answer needs a prescription to perform Step~(2), that is to say, to find a way to unambiguously determine the coefficients $A_{\alpha,p}^{[g]}$ in \eqref{eq:dispexp}, or within the framework of Hamiltonian hierarchies, the dispersive corrections $H^{[g]}$ of the $g=0$ Hamiltonians
\beq
\label{eq:dispdef1}
H^{\rm disp}_{\alpha,p}[\mathbf{u}, \epsilon]=H^{[0]}_{\alpha,p}[\mathbf{u}] + \epsilon H_{\alpha,p}^{[1]}[\mathbf{u}] + \epsilon^2 H_{\alpha,p}^{[2]}[\mathbf{u}]+\dots,
\eeq
for local functionals $H_{\alpha,p}^{[n]}[\mathbf{u}]$
\beq
H_{\alpha,p}^{[j]}[\mathbf{u}] = \int_{S^1} h_{\alpha,p}^{[j]}(\mathbf{u}, \mathbf{u}_x, \mathbf{u}_{xx}, \dots, \mathbf{u}^{(j)}) \mathrm{d} x,
\eeq
where $h_{\alpha,p}^{[j]}$ is a differential polynomial, homogeneous of degree $j$ with respect to the grading \eqref{eq:grading}, and we have appended a superscript $[0]$ to the dispersionless Hamiltonians of the Principal Hierarchy \eqref{eq:princhier}. The statement of integrability is then that
\beq
\l\{H^{\rm disp}_{\alpha,p}[\mathbf{u}, \epsilon], H^{\rm disp}_{\beta,q}[\mathbf{u}, \epsilon]\r\}=0
\label{eq:dispinv}
\eeq
as a formal power series in $\epsilon$. \\
As we emphasized in Sec. \ref{sec:introresults}, in the context of the equivariant Gromov--Witten theory there are no general methods available to date to determine recursively $H_{\alpha,p}^{[n]}[\mathbf{u}]$ starting from the Hamiltonians of the Principal Hierarchy. However, suppose that a dispersive completion
\beq
\label{eq:dispdef2}
H^{\rm disp}_{\overline{\alpha},\overline{p}}[\mathbf{u}, \epsilon]=\sum_{k=0}^\infty
\epsilon^k H^{[k]}_{\overline{\alpha},\overline{p}}[\mathbf{u}]
\eeq
of {\it one} Hamiltonian $\alpha=\overline{\alpha}$, $p=\overline{p}$ be known. We have the following
\begin{thm}[\cite{MR2462355}]
Let $H^{[0]}_{\alpha,p}[\mathbf{u}]$, $\alpha=1,\dots, n$, $p \in \mathbb{N}$ be Hamiltonian local functionals of a dispersionless hierarchy of integrable PDEs
\beq
\{H^{[0]}_{\alpha,p}[\mathbf{u}],H^{[0]}_{\beta,q}[\mathbf{u}]\}=0
\eeq
and let $H^{\rm disp}_{\overline{\alpha},\overline{p}}[\mathbf{u}, \epsilon]$ be a dispersive deformation \eqref{eq:dispdef2} of the Hamiltonian flow $\alpha=\overline{\alpha}$, $p=\overline{p}$ for one pair $(\overline{a}, \overline{p})$ and given local functionals $H^{[k]}_{\overline{\alpha},\overline{p}}[\mathbf{u}]$. Then if a dispersive completion of $H^{\rm disp}_{\alpha,p}[\mathbf{u}, \epsilon]$ preserving involutivity of the flows $\forall \epsilon$ exists
\beq \{H^{\rm disp}_{\alpha,p}[\mathbf{u}, \epsilon],H^{\rm disp}_{\beta,q}[\mathbf{u}, \epsilon]\}=0 \quad \forall \alpha,\beta, p, q, \eeq
it is unique. In such a case, there exists a formal sum of linear differential operators
\bea
\label{doper}
&&
D=\sum_{k=0}^\infty \epsilon^n D^{[k]},
\nn\\
&&
\nn\\
&&
D^{[0]}={\rm id}, \quad D^{[k]} = \sum b^{[k]}_{i_1, \dots, i_n} (u_{1}, \dots, u^{(k)}_{1}, \dots, u_{n}, \dots, u^{(k)}_{n})
\frac{{\partial}^{\sum_j i_j}}{{\partial} u_1^{i_1} \dots {\partial} u_n^{i_{n}}},
\end{eqnarray}
such that
\beq
\int_{S^1}D^{[k]} h^{[0]}_{\alpha,p}(\mathbf{u})\mathrm{d} x= \int_{S^1}h^{[k]}_{\alpha,p}(\mathbf{u}, \mathbf{u}_x, \mathbf{u}_{xx}, \dots, \mathbf{u}^{(k)}) \mathrm{d} x = H^{[k]}_{\alpha,p}[\mathbf{u}]
\eeq
satisfies the involutivity condition \eqref{eq:dispinv}.
In \eqref{doper}, the coefficients $b^{[k]}_{i_1, \dots, i_n}$ are differential polynomials and $\sum_{j=1}^n i_j \leq \left[ \frac{3k}2\right]$.
\label{thm:dop}
\end{thm}
The theorem implies in our case that if a perturbation \eqref{eq:dispdef2} of {\it one} Hamiltonian of the Principal Hierarchy is integrable, then the involutivity condition \eqref{eq:dispinv} singles out an operator \eqref{doper}, which order by order in $\epsilon$ reconstructs the dispersive tail of {\it all} flows. This operator is uniquely defined, modulo total derivatives and the relations \eqref{flatfunct} defining the dispersionless Hamiltonian densities.
\subsection{The resolved conifold and the Ablowitz--Ladik hierarchy}
\label{sec:ablo}
Consider now the Principal Hierarchy for the resolved conifold in the equivariantly Calabi--Yau case $\lambda_1=-\lambda_2=\lambda$. In this case the prepotential is
\beq
F_0^{X_1, \rm ad}=-\frac{1}{2\lambda^2}v^2 w + \operatorname{Li}_3(e^w)
\label{eq:f0conad}
\eeq
and, from the fact that $\eta_{vv}=0$, the non-linear wave equation \eqref{eq:nlinconifold} has a vanishing rectangular term
\beq
w_{tt}=-\lambda^2 \l(\frac{e^w}{1-e^w} w_x\r)_x.
\eeq
This equation was recognized in \cite{MR2462355} to be related to the dispersionless limit of the {\it Ablowitz--Ladik lattice} \cite{MR0377223}. We will here review, almost verbatim, the arguments of \cite{MR2462355} relating the solution of WDVV \eqref{eq:f0conad} to such an integrable lattice. The basic flow of the system is
\bea\label{abla1}
&&
i\,\dot a_n = -\frac12\,(1-a_n b_n) (a_{n-1}+a_{n+1})+ a_n,
\nn\\
&&
\\
&&
i\, \dot b_n = \quad\frac12\,(1-a_n b_n)(b_{n-1} + b_{n+1}) - b_n,
\nn
\end{eqnarray}
where $\{a_n, b_n :\mathbb{Z} \to \mathbb{C}\}$. Introducing new variables
\bea\label{abla6}
&&
u_n =-\log(1-a_n b_n),
\nn\\
&&
\\
&&
y_n = \frac1{2i} \left( \log{\frac{a_n}{a_{n-1}}} -\log{\frac{b_n}{b_{n-1}}} \right),
\nn
\end{eqnarray}
the evolution \eqref{abla1} can be written as a Hamiltonian flow generated by
\beq\label{abla7}
H_{\mathrm{AL}}=\sum_n \sqrt{\left(1-e^{-u_n}\right)\left(1-e^{-u_{n-1}}\right)}\, \cos y_n
\eeq
with the Poisson bracket
\beq\label{abla8}
\{ u_n, y_m\} = \delta_{n,m-1} -\delta_{n,m},\quad \{ u_n,u_m\}=\{ y_n, y_m\}=0.
\eeq
By taking the long-wave expansion we continuously interpolate the discrete dependent variables $u_n$, $t_n$ through functions $u(X,t)$, $y(X,t)$
\beq
u_n =u(\epsilon n, \epsilon t), \quad y_n = y(\epsilon n, \epsilon t).
\eeq
This leads, at the leading order in $\epsilon$, to the dispersionless system
\bea\label{abla9}
&&
u_t ={\partial}_X\left[\left( e^{u}-1\right)\, \sin y\right],
\nn\\
&&
\\
&&
y_t\,= {\partial}_X \left[ e^{-u} \cos y\right].
\nn
\end{eqnarray}
In order to make contact with the Principal Hierarchy of the resolved conifold, we will follow the argument of \cite{MR2462355} replacing $v(X)$, $y(X)$ by
\bea\label{uw1}
x &:=& i \lambda X, \\
v(x)&:=& i y(x) \lambda, \\
w(x)&:=& \frac{i\epsilon \lambda {\partial}_x}{e^{i\epsilon \lambda {\partial}_x} -1}\, u(x).
\end{eqnarray}
In this way, the Poisson brackets of $w$ and $v$ take the standard form \eqref{eq:poissbrack}, and the Hamiltonian \eqref{abla7} becomes upon interpolation
\bea\label{eq:hal}
H_{\rm AL}&=& \int h_{\rm AL}\, \mathrm{d} x \nn \\ &=& \int \sqrt{ \left( 1-\exp\left\{ \frac{1-e^{i\epsilon \lambda {\partial}_x}}{i\epsilon \lambda {\partial}_x}\, w\right\}\right)
\left(1- \exp\left\{ \frac{e^{i\epsilon \lambda {\partial}_x}-1}{i\epsilon \lambda {\partial}_x}\, w\right\}\right)}\, \cosh\left( \frac{v}{\lambda}\right)\, \mathrm{d} x, \nn \\
\end{eqnarray}
\beq
h_{\rm AL}=
\left(-1+e^{w}\right) \cosh \left(\frac{v}{\lambda }\right)-\frac{\left(e^{w} \lambda ^2 \cosh
\left(\frac{v}{\lambda }\right) \left(4 \left(-1+e^{w}\right) w_{xx}-3 (w_{xx})^2\right)\right) \epsilon ^2}{24
\left(-1+e^{w}\right)}+O\left(\epsilon ^4\right).
\label{eq:halpert}
\eeq
It turns out that the Ablowitz--Ladik lattice admits an infinite set of
conserved currents \cite{MR0377223}; as opposed to the Toda case, these
currents do not come straightforwardly from a bi-Hamiltonian recursion
associated to a local Poisson pencil, due to the non-existence of an Euler vector field for the prepotential \eqref{eq:f0conad}. It is easy to show that at the $\mathcal{O}(\epsilon^0)$ an infinite number of them\footnote{In fact all of them, with the sole exception \cite{MR2462355} of the one generating phase shifts of $a_n$, $b_n$.} coincide with the densities of the Principal Hierarchy associated to the prepotential \eqref{eq:f0conad}: the condition for a dispersionless density $f(v(x),w(x))$ to be in involution with the dispersionless Ablowitz--Ladik hamiltonian gives
\beq
\l\{H_{\rm AL}^{[0]} , \int_{S_1}f \r\} = 0 \Leftrightarrow {\partial}^2_{ww}f+\l(\frac{\lambda^2 e^w}{1-e^w}\r){\partial}^2_{vv} f =0,
\label{eq:linweq}
\eeq
which is implied by \eqref{flatfunct} for $\alpha=\beta$. \\
This connection provides us with a viable candidate hierarchy to relate to the Gromov--Witten theory of $X_1$ with $e_T(X_1)=0$, and led us to our Conjecture \ref{conj:AL} connecting the dispersionful Ablowitz--Ladik system with the all-genus theory of $X_1$. To this aim, and as a first step towards the reconstruction of the dispersionful hierarchy, let us remark here that by Theorem \ref{thm:dop}, Eq. \eqref{eq:hal} offers us a way to effectively construct the dispersive flows:
\begin{prop}
The $D$-operator for the Ablowitz--Ladik hierarchy reads at $\mathcal{O}(\epsilon^2)$
\bea
D_{\rm AL} f &=& f+ \epsilon^2\Bigg[
\frac{e^{w(x)} \left(-1+2 e^{w(x)}\right) w'(x)^2
f_{vv} \lambda ^4}{24
\left(-1+e^{w(x)}\right)^2}+\frac{e^{w(x)} w'(x)^2
f_{vvw} \lambda ^4}{12
\left(-1+e^{w(x)}\right)} \nn \\ &+& \frac{e^{w(x)} w'(x) v'(x)
f_{vvv} \lambda ^4}{6
\left(-1+e^{w(x)}\right)}+\frac{v'(x)^2 f_{vv}
\lambda ^2}{-12+12 e^{-w(x)}}+\frac{1}{12} v'(x)^2
f_{vvw} \lambda ^2\Bigg] + \mathcal{O}(\epsilon^4). \nn \\
\label{eq:dop1loop}
\end{eqnarray}
\end{prop}
\noindent Eq. \eqref{eq:dop1loop} was obtained, with minor discrepancies due to a different choice of variables, in \cite{MR2462355}. A sketch of the proof can be found in Appendix \ref{sec:appdop}. Applying \eqref{eq:dop1loop} to the densities $h_{\alpha,p}$ of the Principal Hierarchy we find
\bea
h_v^{[2]}(v,w,v_x,w_x) &=& \frac{1}{24
\left(-1+e^{w(x)}\right)^2} \Bigg[e^{w(x)} \lambda ^2 \left(-w(x)+2 e^{w(x)} (w(x)+1)-2\right) w'(x)^2 \nn \\
&-& 2 \left(-1+e^{w(x)}\right) \left(e^{w(x)} (w(x)-1)+1\right) v'(x)^2\Bigg] z^2+O\left(z^3\right), \\
h_w^{[2]}(v,w,v_x,w_x) &=& \frac{e^{w(x)} \left(\left(-1+2 e^{w(x)}\right) \lambda ^2 w'(x)^2-2 \left(-1+e^{w(x)}\right) v'(x)^2\right) z}{24 \left(-1+e^{w(x)}\right)^2} \nn \\ &+& \frac{e^{w(x)}}{24
\left(-1+e^{w(x)}\right)^2} \Bigg[4
\left(-1+e^{w(x)}\right) w'(x) v'(x) \lambda ^2 \nn \\ &+& v(x) \left(\left(-1+2 e^{w(x)}\right) \lambda ^2 w'(x)^2-2 \left(-1+e^{w(x)}\right) v'(x)^2\right)\Bigg] z^2+O\left(z^3\right). \nn \\
\label{eq:h1loop}
\end{eqnarray}
As an example, the leading order correction to \eqref{eq:nlinconifold} reads
\bea
w_{tt} &=& \l(\frac{-\lambda^2 e^{w}}{1-e^w}w_x\r)_x- \frac{e^{w(x)}}{24 \left(-1+e^{w(x)}\right)^4} \Bigg[\left(1+4 e^{w(x)}+e^{2 w(x)}\right) \lambda ^2 w'(x)^3 \nn \\ &+& \l(-2+2e^{2 w(x)}\r) \left(v'(x)^2-2 \lambda ^2 w''(x)\right) w'(x)-2 \left(-1+e^{w(x)}\right)^2 \nn \\ & \times & \left(2 v'(x) v''(x)-\lambda ^2 w^{(3)}(x)\right)\Bigg]_x\epsilon^2+\mathcal{O}(\epsilon^4).
\end{eqnarray}
\subsection{Quasi-triviality and genus one Gromov--Witten invariants}
\label{sec:quasitr}
A key ingredient in the Dubrovin--Zhang analysis of bi-Hamiltonian evolutionary
hierarchies, which proved instrumental in their proof \cite{MR2092034} of the
$\mathbb{P}^1$/Toda correspondence, is the fact the hierarchy verifying Conjecture
\ref{conj:integrableGW} satisfies a {\it quasi-triviality} property:
\begin{defn}
A transformation of the form
\beq
u^\alpha \to z^\alpha = u^\alpha+\sum_{k=1}^\infty \epsilon^{2k} F_{k}^\alpha(\mathbf{u}, \mathbf{u}_x, \dots, \mathbf{u}^{(m(k))}),
\label{eq:quasimiura}
\eeq
with $m(k)$ a monotonically increasing, positive integer-valued sequence and $F_{k}^\alpha$ a degree $k$ rational function of $\mathbf{u}^{(j)}$ for $j>0$, will be called a quasi-Miura transformation. The hierarchy \eqref{eq:dispexp} is called quasi-trivial if there exists a quasi-Miura transformation reducing it, in the new variables, to its dispersionless $\epsilon=0$ truncation.
\end{defn}
The quasi-Miura transformation \eqref{eq:quasimiura} is said to be $\tau$-symmetric if there exists a formal power series
\beq
\mathcal{F}(\epsilon, \mathbf{u}, \mathbf{u}_x, \dots) = \sum_{k=0}^\infty \epsilon^{2k} \mathcal{F}_{k}(\mathbf{u}, \mathbf{u}_x, \dots, \mathbf{u}^{(m(k)-2)})
\label{eq:Fvsde2F}
\eeq
where again $\mathcal{F}^{[k]}(\mathbf{u}, \mathbf{u}_x, \dots, \mathbf{u}^{(m(k)-2)})$ is a degree $k$ rational function of $\mathbf{u}^{(j)}$ for $j>0$, such that
\beq
F_{k,\alpha}(\mathbf{u}, \mathbf{u}_x, \dots, \mathbf{u}^{(m(k))})= \frac{{\partial}^2 \mathcal{F}_k}{{\partial} x {\partial} t^{\alpha,0}}(\mathbf{u}, \mathbf{u}_x, \dots, \mathbf{u}^{(m(k)-2)})
\label{eq:tausymmmiura}
\eeq
Comparing \eqref{eq:tausymmmiura} with \eqref{eq:tau} identifies $\mathcal{F}_k$ as
the $\mathcal{O}(\epsilon^k)$ dispersive correction to the logarithm of the dispersionless $\tau$ function, i.e., in the case of the Principal Hierarchy \eqref{eq:princhier}, of the genus zero topological $\tau$-function of Theorem \ref{thm:dubr}. Conjecture \ref{conj:integrableGW} then states that precisely one such object should correspond to the full descendent, all-genus Gromov--Witten potential. \\
The rest of this section is devoted to the use of a quasi-triviality transformation to give a proof of Conjecture \ref{conj:AL} at $\mathcal{O}(\epsilon^2)$, that is, for $g = 1$ Gromov--Witten invariants. Our proof consists of the following three steps:
\bit
\item[(a)] proving that the Ablowitz--Ladik hierarchy is quasi-trivial at $\mathcal{O}(\epsilon^2)$;
\item[(b)] fixing a suitable choice of dependent variables leading to a $\tau$-symmetric transformation \eqref{eq:tausymmmiura} at $\mathcal{O}(\epsilon^2)$;
\item[(c)] proving that the logarithm of the $\tau$-function thus obtained coincides with the genus one, full descendent potential of $X_1$ with anti-diagonal action.
\end{itemize}
In the context of the usual theory of (conformal) semi-simple Frobenius
structures, step (a) is a consequence of the theory of $(0,n)$ Poisson pencils
on the loop space, while step (b) and (c) follow from the axiom of
linearization of Virasoro constraints. In absence of bi-Hamiltonianity and
possibly Virasoro-type constraints, we will need to perform steps (a)-(c) ``by
hand'', guided to some extent by the analogy with the Extended Toda hierarchy. \\
\noindent Let us start from step (a):
\begin{thm}
There exists an infinitesimal time-$\epsilon$ canonical quasi-Miura transformation
\beq
u^\alpha \to z^\alpha = u^\alpha+\epsilon \l\{u^\alpha, K\r\}+\mathcal{O}(\epsilon^2),
\label{eq:quasiK}
\eeq
where
\beq
K = -\frac{\epsilon}{24}\int_{S^1} \Bigg[ F_+(x) \log
F_+(x) + F_-(x) \log
F_-(x) - 2 \log \left(-1+e^{w(x)}\right) v'(x)\Bigg] \mathrm{d} x
\label{eq:K}
\eeq
and
\beq
F_\pm(x):= v'(x) \pm \sqrt{\frac{e^{w(x)}}{-1+e^{w(x)}}} \lambda
w'(x)
\eeq
which sends solutions of the Principal Hierarchy to those of its $\mathcal{O}(\epsilon^2)$ correction \eqref{eq:h1loop}.
\label{thm:K}
\end{thm}
\noindent To prove it we will make use of the following technical lemma from \cite{MR2462355}:
\begin{lem}
A density $h(\mathbf{u}, \mathbf{u}_x, \ldots)$, depending at most rationally on the jet variables $\mathbf{u}^{(n)}$ for $n>1$ is a total derivative ${\partial}_x g(\mathbf{u}, \mathbf{u}_x, \ldots)$ if and only if
\beq
\frac{\delta}{\delta \mathbf{u}(y)} \int_{S^1} h(\mathbf{u}(x), \mathbf{u}_x(x), \ldots) \mathrm{d} x=0.
\eeq
\label{lem:totalder}
\end{lem}
\noindent{\it Proof of Theorem \ref{thm:K}.} The proof follows from a very lengthy, but straightforward application of Lemma \ref{lem:totalder}. The Hamiltonian densities $h^{\rm qt}_{\alpha,p}(\mathbf{z}, \mathbf{z}_x, \mathbf{z}_{xx})$ obtained by composition of the dispersionless densities $h_{\alpha,p}(\mathbf{u})$ with the $\mathcal{O}(\epsilon^2)$ quasi-Miura transformation \eqref{eq:quasiK} are given by
\beq
h^{\rm qt}_{\alpha,p}(\epsilon, \mathbf{z}, \mathbf{z}_x, \mathbf{z}_{xx}, \dots) :=h_{\alpha,p}(\mathbf{u}(\epsilon, \mathbf{z}, \mathbf{z}_x, \dots))=h_{\alpha,p}(\mathbf{z})+\epsilon\l\{h_{\alpha,p}, K\r\}+o(\epsilon^2)
\eeq
since the trasformation generated by $K$ is canonical. In particular the $\mathcal{O}(\epsilon^2)$ correction is
\beq
h^{[2], \rm qt}_{\alpha,p}(\mathbf{z}, \mathbf{z}_x, \mathbf{z}_{xx}) = \epsilon\l\{h_{\alpha,p}, K\r\}.
\eeq
We claim that this reproduces the leading dispersive correction \eqref{eq:h1loop} of the dispersionless Ablowitz--Ladik flows. In the following we denote by $h^{\rm Dop}_{\alpha,p}$ the densities \eqref{eq:h1loop} we got by acting with the $D$-operator \eqref{eq:dop1loop}, to distinguish them from the ones obtained via \eqref{eq:quasiK}; accordingly, the corresponding generating functions will be written $h^{\rm Dop}_{\alpha}(z)$ and $h^{\rm qt}_{\alpha}(z)$. Define
\beq
r_{\alpha}(z) := h^{\rm qt}_{\alpha}(z)-h^{\rm Dop}_{\alpha}(z).
\eeq
We refrain here from reproducing the exact form of $r_{\alpha}(z)$, which would
occupy alone a few pages\footnote{We will be happy to provide the interested
reader with the details of this calculation.}. A direct computation using \eqref{eq:dop1loop} and \eqref{eq:K} shows that
\bea
\frac{\delta}{\delta \mathbf{u}(x)} \int_{S^1} r^{[2]}_{\alpha}(z)(\mathbf{u}, \mathbf{u}_y, \mathbf{u}_{yy}) \mathrm{d} y =
0.
\end{eqnarray}
Therefore we conclude that
\beq
H^{\rm qt, [2]}_{\alpha,p}= H^{\rm Dop, [2]}_{\alpha,p}.
\label{eq:HqtHdop}
\eeq
\begin{flushright}$\square$\end{flushright}
\begin{rem}
Even though the Hamiltonian flows they generate are equal by \eqref{eq:HqtHdop}, the Hamiltonian densities $h^{\rm qt}_{\alpha,p}$ and $h^{\rm Dop}_{\alpha,p}$ differ greatly in form; in particular the quasi-Miura transformation \eqref{eq:quasiK} introduces a rational, rather than polynomial, dependence of $h^{\rm qt}_{\alpha,p}$ on the jet variables $\mathbf{u}^{(n)}$ for $n \geq 1$, which should disappear only at the level of the flows by \eqref{eq:HqtHdop}. For the basic dispersionless Hamiltonian of the Ablowitz--Ladik lattice we have for example
\beq
\l\{(1-e^w)\cosh\l(\frac{v}{\lambda}\r), K \r\} = -\frac{P(\mathbf{u}, \mathbf{u}_x, \mathbf{u}_{xx}, \mathbf{u}^{(3)})}{48
\left(-1+e^{w(x)}\right)^2 \left(e^{w(x)} \lambda ^2 w'(x)^2-\left(-1+e^{w(x)}\right) v'(x)^2\right)^2}
\eeq
where $P(\mathbf{u}, \mathbf{u}_x, \mathbf{u}_{xx}, \mathbf{u}^{(3)})$ looks like
\bea
P &=& e^{w(x)-\frac{v(x)}{\lambda }} \lambda ^2 \Bigg[\left(1+e^{\frac{2 v(x)}{\lambda }}\right) \Big(-3 e^{3 w(x)} \lambda ^4
w'(x)^6+e^{w(x)} \left(-1+e^{w(x)}\right) \lambda ^2 \nn \\ &\times& \left(e^{w(x)} \left(1+2 e^{w(x)}\right) w''(x) \lambda ^2+\left(1+5
e^{w(x)}\right) v'(x)^2\Big) w'(x)^4\r) -2 e^{w(x)} \left(-1+e^{w(x)}\right)^2 \nn \\
&\times& \lambda ^2 \left(e^{w(x)} \lambda ^2 w^{(3)}(x)-2 v'(x)
v''(x)\right) w'(x)^3-2 e^{w(x)} \left(-1+e^{w(x)}\right)^2 v'(x)^4+\dots
\Bigg]. \nn \\
\end{eqnarray}
\end{rem}
It is particularly instructive to consider the result of the quasi-Miura transformation on the variable $w$. We find
\beq
\epsilon \l\{w, K\r\}=-\lambda^2 \epsilon^2 \frac{{\partial}^2\tilde\mathcal{F}_1(w, v_x, w_x)}{\partial x^2},
\label{eq:quasitrw}
\eeq
where
\beq
\tilde\mathcal{F}_1(w, v_x, w_x):=\frac{1}{24}\log\l( v'(x)^2+\frac{\lambda^2 e^{w(x)}}{1-e^{w(x)}}w'(x)^2\r)+\frac{1}{12}\operatorname{Li}_1(e^{w(x)}).
\label{eq:F1tilde}
\eeq
On the other hand, for the $v$ variable we find
\beq
\epsilon \l\{v, K\r\}=-\lambda^2 \epsilon^2 \l(\frac{{\partial}^2\tilde\mathcal{F}_1(w, v_x, w_x)}{\partial x \partial t^{2,0}}+ \frac{1}{6} \frac{{\partial}^2 \operatorname{Li}_1(e^{w})}{\partial x \partial t^{2,0}} \r) \neq -\lambda^2 \epsilon^2 \frac{{\partial}^2\tilde\mathcal{F}_1(w, v_x, w_x)}{\partial x \partial t^{2,0}},
\label{eq:quasitrv}
\eeq
that is, the quasi-triviality transformation \eqref{eq:quasiK} is {\it not}
$\tau$-symmetric. Indeed, except for the KdV case, the incompatibility between
canonicity and $\tau$-symmetry of the quasi-Miura transformation seems to be a
fairly generic fact\footnote{This point was strongly emphasized to us in an
enlightening discussion with B.~Dubrovin.}, also for bi-Hamiltonian systems
like the Extended Toda hierarchy. Quite interestingly, the similiarity with
the case of the Gromov--Witten theory of $\mathbb{P}^1$ and the Extended Toda
hierarchy is even more striking, as the discrepancy between the transformation
generated by $K$ and the one via the logarithm of a dispersive $\tau$-function
is present in only {\it half} of the change of variables, and it is equal to
twice the term independent of space derivatives (in the language of
\cite{Dubrovin:1997bv}, the $G$-function) inside $\tilde\mathcal{F}_1$. \\
It should be
emphasized that the form of the $G$-function in the Toda case relies on the
existence of a grading condition for the Frobenius manifold associated to
$QH^\bullet(\mathbb{P}^1)$, which in particular allows to fix the degree zero terms at
$g=1$ \cite{Dubrovin:1997bv}. As we will see, in our case this is achieved by shifting the
form of the $g=1$ full-descendent free energy by the constant map term
\beq
\langle \tau_0(1) \rangle^X_{1,1,0}=- \int_{\mathcal M_{1,1}} \lambda_1=-\frac{1}{24},
\eeq
where $\lambda_1=c_1(\mathbb{E})$ is the Chern class of the Hodge
line bundle on $\overline{\mathcal M}_{1,1}$, through
\beq
\tilde \mathcal{F}_1 \to \mathcal{F}_1:=\tilde \mathcal{F}_1 - \frac{w(x)}{24}.
\label{eq:dzeroshift}
\eeq
Inspired by the analogy with the $\mathbb{P}^1$/Toda case, we are then led to consider the following $\tau$-symmetric ansatz for the choice of dependent
variables at $\mathcal{O}(\epsilon^2)$:
\beq
u_\alpha \to u_\alpha + \epsilon^2 \frac{{\partial}^2\mathcal{F}_1(w, v_x, w_x)}{\partial x \partial t^{\alpha,0}}+\mathcal{O}(\epsilon^4),
\label{eq:tausymf1}
\eeq
where
\beq
\mathcal{F}_1=\frac{1}{24}\log\l( v'(x)^2+\frac{\lambda^2
e^{w(x)}}{1-e^{w(x)}}w'(x)^2\r)+\frac{1}{12}\operatorname{Li}_1(e^{w(x)})- \frac{w(x)}{24}.
\label{eq:F1}
\eeq
We are now in position to prove Theorem \ref{thm:main}: by the definition of quasi-triviality, the $\mathcal{O}(\epsilon^2)$ correction to the logarithm of the $\tau$-function $\mathcal{F}_0$ in \eqref{eq:dercorr} should be obtained by plugging into \eqref{eq:F1} the solution of the Principal Hierarchy with initial data \eqref{eq:initial}; the result of the composition will be again denoted by the same symbol $\mathcal{F}_1(\mathbf{t})$. Accordingly, we introduce genus 1 ``big correlators''
\beq
\left\langle\bra \tau_{p_1}\phi_{\alpha_1}, \dots, \tau_{p_k}\phi_{\alpha_k} \right\rangle\ket_1 := \frac{{\partial}^{k} \mathcal{F}_1(\mathbf{t})}{{\partial}{t^{\alpha_1,p_1}} \dots \partial {t^{\alpha_k,p_k}}}.
\label{eq:dercorr1loop}
\eeq
\noindent{\it Proof of Theorem \ref{thm:main}.} The statement for $g=0$ was
proven in Sec. \ref{sec:dispconif} (Theorem \ref{thm:dubr} and Proposition
\ref{prop:norm}). For $g=1$, notice that the descendent Gromov--Witten
invariants of $X_1$ are completely determined in a recursive fashion by the
following two formulas: the first is the $g=1$ case of the higher genus
multi-covering formula found in \cite{Marino:1998pg, Gopakumar:1998jq,
Katz:1999xq, MR1728879, Bryan:2004iq} that yields the primary potential of $X_1$ as
\beq
F_1^{X_1}(t^{2,0})=- \frac{t^{2,0}}{24} + \sum_{d=1}^\infty N^{(1)}_{1,d} e^{d t^{2,0}}=- \frac{t^{2,0}}{24}+\frac{1}{12}\operatorname{Li}_1(e^{t_{2,0}}).
\label{eq:primpotg1}
\eeq
The second is the set of $g=1$ topological recursion relations \cite{Dijkgraaf:1990nc, Kontsevich:1997ex, MR1672112}
\bea
\left\langle\tau_{p}(\phi_\alpha) \right\rangle_{1,1,d}^{X_1} &=&
\sum_{d_1+d_2=d}\left\langle\tau_{p-1}(\phi_\alpha) \tau_{0}(\phi_\nu)
\right\rangle_{0,2,d_1}^{X_1} \eta^{\mu \nu}\left\langle\tau_{0}(\phi_\mu) \right\rangle_{1,1,d_2}^{X_1}
\nn \\ &+& \frac{1}{24} \eta^{\mu \nu}\left\langle\tau_{p-1}(\phi_\alpha) \tau_{0}(\phi_\nu) \tau_{0}(\phi_\mu) \right\rangle_{0,3,d}^{X_1},
\label{eq:g1trr}
\end{eqnarray}
that fully determine genus one descendent invariants in terms of the genus one primaries and genus zero descendents.
We will prove here that the Ablowitz--Ladik $\tau$-function \eqref{eq:F1} implies both \eqref{eq:primpotg1} and \eqref{eq:g1trr}. \\
\noindent Consider first the small phase reduction of $\mathcal{F}(\mathbf{t})$, i.e.
\bea
u^\alpha(\mathbf{t})\l|_{\substack{t^{\alpha,p}=0 \\ \hbox{\footnotesize for } p>0}}\r. &=& t^{\alpha,0}+\delta^{\alpha,0}x.
\label{eq:initcond}
\end{eqnarray}
Replacing into \eqref{eq:F1} we find
\bea
\mathcal{F}_1(\mathbf{t})\l|_{\substack{t^{\alpha,p}=0 \\ \hbox{\footnotesize for }
p>0}}\r. &=& - \frac{t^{2,0}}{24}+\frac{1}{12}\operatorname{Li}_1(e^{t^{2,0}}),
\end{eqnarray}
which proves \eqref{eq:primpotg1}. \\
\noindent To see that $\mathcal{F}_1(\mathbf{t})$ correctly embeds the full-descendent information too, we follow \cite{Dubrovin:1997bv} and compute
\bea
\mathcal{C}_{\alpha,p} &:=& \frac{{\partial} \mathcal{F}_1}{{\partial} t^{\alpha,p}}-\frac{{\partial}^2 \mathcal{F}_0}{{\partial} t^{\alpha,p-1} {\partial} t^{\nu,0}}\eta^{\mu\nu}\frac{{\partial} \mathcal{F}_1}{{\partial} t^{\mu,0}} \nn \\ &=&
\sum_{\gamma=1,2}\l[\frac{{\partial} \mathcal{F}_1}{{\partial} u^\gamma} \frac{{\partial} u^\gamma}{{\partial} t^{\alpha,p}}+\frac{{\partial} \mathcal{F}_1}{{\partial} u_x^\gamma} {\partial}_x \l(\frac{{\partial} u^\gamma}{{\partial} t^{\alpha,p}}\r) - {\partial}_\nu h^{[0]}_{\alpha,p-1}\eta^{\mu\nu}\l(\frac{{\partial} \mathcal{F}_1}{{\partial} u^{\gamma}}\frac{{\partial} u^\gamma}{{\partial} t^{\mu,0}}+ \frac{{\partial} \mathcal{F}_1}{{\partial} u_x^{\gamma}}{\partial}_x\frac{{\partial} u^\gamma}{{\partial} t^{\mu,0}} \r)\r] \nn \\
&=& \sum_{\gamma=1,2}\Bigg[\frac{{\partial} \mathcal{F}_1}{{\partial} u^\gamma} {\partial}_\nu h^{[0]}_{\alpha,p-1} \frac{{\partial} u^\gamma}{{\partial} t^{\mu,0}}\eta^{\mu \nu}+
\frac{{\partial} \mathcal{F}_1}{{\partial} u_x^\gamma} {\partial}_x \l( {\partial}_\nu h^{[0]}_{\alpha,p-1} \frac{{\partial} u^\gamma}{{\partial} t^{\mu,0}}\eta^{\mu \nu}\r) \nn \\
&-& {\partial}_\nu h^{[0]}_{\alpha,p-1}\eta^{\mu\nu}\l(\frac{{\partial} \mathcal{F}_1}{{\partial} u^{\gamma}}\frac{{\partial} u^\gamma}{{\partial} t^{\mu,0}}+ \frac{{\partial} \mathcal{F}_1}{{\partial} u_x^{\gamma}}{\partial}_x\frac{{\partial} u^\gamma}{{\partial} t^{\mu,0}} \r)\Bigg] \nn \\
&=& \sum_{\gamma=1,2}\frac{{\partial} \mathcal{F}_1}{{\partial} u^\gamma_x} {\partial}_x {\partial}_\nu h^{[0]}_{\alpha,p-1} \frac{{\partial} u^\gamma}{{\partial} t^{\mu,0}}\eta^{\mu \nu},
\label{eq:Cap}
\end{eqnarray}
where in the second line we used the chain rule and the fact that
\beq
{\partial}_\nu h^{[0]}_{\alpha,p-1} = \left\langle\bra \tau_{p-1} (\phi_\alpha) \tau_{0} (\phi_\nu) \right\rangle\ket_0 = \frac{{\partial}^2 \mathcal{F}_0}{{\partial} t^{\alpha,p-1} {\partial} t^{\nu,0}},
\eeq
while in the third line we used the genus zero topological recursion relations \eqref{eq:dubrTRR}
\bea
\frac{{\partial} u^\alpha}{{\partial} t^{\beta,q}} &=& \eta^{\alpha\gamma}\frac{{\partial}^3 \mathcal{F}_0}{{\partial} t^{\gamma,0} {\partial} t^{1,0} {\partial} t^{\beta,q}} = \eta^{\alpha\gamma}\frac{{\partial}^2 \mathcal{F}_0}{ {\partial} t^{\delta,0} {\partial} t^{\beta,q-1}} \eta^{\delta \epsilon}\frac{{\partial}^3 \mathcal{F}_0}{{\partial} t^{\gamma,0} {\partial} t^{1,0} {\partial} t^{\delta,0}} \nn \\
&=& \frac{{\partial} u^\alpha}{{\partial} t^{\delta,0}}\eta^{\delta \epsilon} {\partial}_\epsilon h^{[0]}_{\alpha,p-1}.
\end{eqnarray}
Moreover Eq. \eqref{eq:firstflowsconifold} with $\lambda_1=-\lambda_2=\lambda$ and the explicit form \eqref{eq:F1} yield
\bea
\sum_{\gamma=1,2} \frac{{\partial} \mathcal{F}_1}{{\partial} u^\gamma_x} \frac{{\partial} u^\gamma}{{\partial} t^{\mu,0}}\eta^{\mu \nu} &=& \frac{{\partial} \mathcal{F}_1}{{\partial} v_x}\frac{{\partial} v}{{\partial} t^{\nu,0}}+\frac{{\partial} \mathcal{F}_1}{{\partial} w_x}\frac{{\partial} w}{{\partial} t^{\nu,0}} =\frac{\eta^{\mu \nu}}{24} \frac{2 v'(x) {\partial}_{t^{\mu,0}} v + 2 \frac{\lambda^2 e^{w(x)}}{1-e^{w(x)}} w'(x) {\partial}_{t^{\mu,0}} w}{v'(x)^2 + \frac{\lambda^2 e^{w(x)}}{1-e^{w(x)}}w'(x)^2} \nn \\
&=&-\frac{\lambda^2}{12} \delta^{\nu,0}.
\end{eqnarray}
This implies
\beq
\mathcal{C}_{\alpha,p}=-\frac{\lambda^2}{12} \partial_x \partial_w h^{[0]}_{\alpha,p} =-\frac{\lambda^2}{12}\frac{{\partial}^3 \mathcal{F}_0}{{\partial} t^{1,0} {\partial} t^{2,0} {\partial} t^{\alpha,p-1}}=\frac{1}{24}\eta^{\mu\nu} \frac{{\partial}^3 \mathcal{F}_0}{{\partial} t^{\alpha,p-1} {\partial} t^{\nu,0} {\partial} t^{\mu,0}}.
\eeq
Combining the last equality with the definition of $\mathcal{C}_{\alpha,p}$ in the first line of \eqref{eq:Cap} we obtain
\beq
\left\langle\bra\tau_{p}(\phi_\alpha) \right\rangle\ket_1 = \left\langle\bra\tau_{p-1}(\phi_\alpha) \tau_{0}(\phi_\nu) \right\rangle\ket_0 \eta^{\mu \nu}\left\langle\bra\tau_{0}(\phi_\mu) \right\rangle\ket_1+\frac{1}{24} \eta^{\mu \nu}\left\langle\bra\tau_{p-1}(\phi_\alpha) \tau_{0}(\phi_\nu) \tau_{0}(\phi_\mu) \right\rangle\ket_0,
\eeq
which, setting $t^{\alpha,p}=0$ for $p>0$ in \eqref{eq:dercorr1loop} and expanding in $e^{t_{2,0}}$, implies \eqref{eq:g1trr}.
\begin{flushright}$\square$\end{flushright}
\subsection{A look at the higher genus theory}
\label{sec:g2}
A natural further step would be to generalize the results of the previous
section to higher genus Gromov--Witten invariants. As usual, however, the
degree of difficulty undergoes a phase transition as soon as $g>1$, and the search of straightforward generalizations of the methods we used becomes unwieldy. In particular, the construction of the quasi-triviality transformation appears to be very hard, let alone finding a suitable $\tau$-symmetric form to compare with the higher genus Gromov--Witten potentials. \\
However, there is still room for a number of non-trivial tests of Conjecture \ref{conj:AL}. To see this, recall that in the previous section we found three ways to construct the $\mathcal{O}(\epsilon^2)$ dispersive tail of the Ablowitz--Ladik hierarchy:
\ben[(i)]
\item via the $D$-operator \eqref{eq:dop1loop};
\item via the canonical quasi-Miura transformation \eqref{eq:quasiK};
\item via the $\tau$-symmetric quasi-Miura transformation \eqref{eq:tausymf1}.
\end{enumerate}
(i) and (ii) are equivalent by Theorem \ref{thm:K}, and (ii) and (iii) because
rational Miura transformations form a group. At the level of the flows the
statement of Theorem \ref{thm:K} is stronger than that, meaning that this
equivalence is realized as an equality of the form of the equations of the
hierarchy \eqref{eq:HqtHdop}. By \eqref{eq:quasitrw} and \eqref{eq:quasitrv}
this is not the case for (ii) and (iii), where
$\tau$-symmetry is broken in the canonical setup; moreover, the canonical free
energy $\tilde\mathcal{F}$ such that $w^{[\rm d-op] }(\mathbf{t})={\partial}^2_x
\tilde{\mathcal{F}}(\mathbf{t})$ coincides with the topological free energy $\mathcal{F}$ up
to a Miura transformation, consisting in a shift by terms
whose restriction to primary invariants involves only degree zero terms.
The situation is schematized in Table \ref{tab:disp}. \\
\begin{table}[h]
\begin{tabular}{ccccc}
\bf D-operator & & \bf Canonical q.t. & & \bf $\tau$-symmetric q.t. \\
\hline
\\
$w^{[\rm d-op]}(\mathbf{t}, \epsilon)$ & $=$ & $w^{[\rm c.q.t.]}(\mathbf{t}, \epsilon)$ & $=$ & $w^{[\rm \tau-s.q.t.]}(\mathbf{t}, \epsilon)|_{d>0}$ \\
$v^{[\rm d-op]}(\mathbf{t}, \epsilon)$ & $=$ & $v^{[\rm c.q.t.]}(\mathbf{t}, \epsilon)$ & $\neq$ & $v^{[\rm \tau-s.q.t.]}(\mathbf{t}, \epsilon)_{d>0} $
\end{tabular}
\vspace{.5cm}
\caption{\small Relations between solutions of the dispersionful Ablowitz--Ladik
hierarchy at $\mathcal{O}(\epsilon^2)$. The equality between the second and the third
column holds up to a Miura transformation whose restriction to the
small phase space involves only degree zero terms.}
\label{tab:disp}
\end{table}
The objects to construct for the purpose of computing higher genus Gromov--Witten invariants are $v^{[\rm \tau-s.q.t.]}(\mathbf{t}, \epsilon)$ and $w^{[\rm \tau-s.q.t.]}(\mathbf{t}, \epsilon)$ at higher order in $\mathcal{O}(\epsilon)$; as we emphasized, the relevant $\tau$-symmetric quasi-Miura transformation seems however very difficult to obtain. On the other hand, focusing on the first line we see that at the leading order in the perturbative expansion in $\epsilon$ we have
\beq
w^{[\rm d-op]}(\mathbf{t}, \epsilon) = w^{[\rm \tau-s.q.t.]}(\mathbf{t}, \epsilon) = \frac{{\partial}^2 \mathcal{F}}{{\partial} x^2}(\mathbf{t})+ \mathcal{O}(\epsilon^4)
\label{eq:wdopwtsqt0}
\eeq
up to terms related to constant maps contribution upon restriction to primary
fields, as in \eqref{eq:dzeroshift}. In the following we will make the following important
\begin{assm}
The equality \eqref{eq:wdopwtsqt0} holds true at genus $\mathcal{O}(\epsilon^4)$:
\beq
w^{[\rm d-op]}(\mathbf{t}, \epsilon) = w^{[\rm \tau-s.q.t.]}(\mathbf{t}, \epsilon) = \frac{{\partial}^2 \mathcal{F}}{{\partial} x^2}(\mathbf{t})+ \mathcal{O}(\epsilon^6)
\label{eq:wdopwtsqt}
\eeq
up to a quasi-Miura transformation whose restriction to primaries is
determined by {\rm degree zero} Gromov--Witten invariants.
\label{assm:tausymm}
\end{assm}
That is, even if we do not know the form of $w^{[\rm d-op]}(\mathbf{t}, \epsilon)$
as a rational function in the derivatives of the fields beyond $\mathcal{O}(\epsilon^2)$, we assume that it is a double
derivative of some (rational) local functional $\mathcal{F}$. In our situation, we
have little guidance for the construction of the right quasi-Miura
transformation which determines the form of $\mathcal{F}$ at higher genus, but on the other hand
computing higher order corrections to the $D$-operator \eqref{eq:dop1loop},
and therefore to $w^{[\rm d-op]}(\mathbf{t})$, is just a matter of
computational time and stamina. Indeed the involutivity condition
\eqref{eq:dispinv} gives a self-contained way to find $w^{[\rm
d-op]}(\mathbf{t}, \epsilon)$ at any order in $\epsilon$, thus allowing a complete
recursive reconstruction of the $\epsilon$ expansion of the flows. As for the genus
one case, we might be missing a possible contribution from constant maps here; the computation below
will indeed be insensitive to degree zero invariants.\\
\noindent {\it Proof of Theorem \ref{cor:g=2}:} we will work out the
consequences of Theorem \ref{thm:dop} at $\mathcal{O}(\epsilon^4)$ by putting
$H_{\alpha,p}=H_{\rm AL}$, where the $\mathcal{O}(\epsilon^4)$ expansion of $H_{\rm AL}$ was given in \eqref{eq:halpert}. This determines the two-loops $D$-operator of the Ablowitz--Ladik lattice; for future use we give here the expression of the coefficients $b^{[4]}_{vv}$ and $b^{[4]}_{vvv}$
\beq
D^{[4]}f = b^{[4]}_{vv} f_{vv} + b^{[4]}_{vvv} f_{vvv} + \dots
\eeq
where, up to a total derivative, we have
\bea
\label{eq:MISTIMO0}
b^{[4]}_{vv} &=& \frac{\lambda ^2}{{5760 \left(-1+e^{w(x)}\right)^4}} \Bigg[
-8 e^{w(x)} \left(-1+e^{2 w(x)}\right) v'(x)^4+4 \left(-1+e^{w(x)}\right) \Big(e^{w(x)} \nn \\ & \times &\left(11+7 e^{w(x)}\right)-2\Big)
\lambda ^2 w''(x) v'(x)^2-8 \left(-1+e^{w(x)}\right) \left(e^{w(x)} \left(-1+19 e^{w(x)}\right)+2\right) \lambda ^2 \nn \\ & \times & w'(x)
v''(x) v'(x)+\lambda ^2 \Big(\left(4-e^{w(x)} \left(15 e^{w(x)} \left(-4+e^{w(x)}\right)+19\right)\right) \lambda ^2 w''(x)
w'(x)^2 \nn \\ &+& 8 e^{w(x)} \left(-1+e^{w(x)}\right) \Big(\left(e^{w(x)} \left(-7+3 e^{w(x)}\right)+1\right) \lambda ^2
w''(x)^2+\left(-1+e^{w(x)}\right) \nn \\ & \times & \left(5+7 e^{w(x)}\right) v''(x)^2\Big)\Big),
\Bigg] \\
b^{[4]}_{vvv} &=& \frac{1}{17280} \Bigg[\frac{\lambda ^4}{\left(1-e^{w(x)}\right)^3}\Bigg(
2 \Big(4 w(x) \left(-1+e^{w(x)}\right)^3+3 \Big(4 \log \left(-1+e^{w(x)}\right)
\nn \\ & \times &
\left(-1+e^{w(x)}\right)^3+e^{w(x)} \left(e^{w(x)} \left(-107+32 e^{w(x)}\right)+131\right)-46\Big)\Big) w'(x) v'(x)
w''(x) \lambda ^2\Bigg) \nn \\ &+& \left(\frac{3 \left(e^{w(x)} \left(-119+183
e^{w(x)}\right)+46\right)}{\left(-1+e^{w(x)}\right)^3}-12 \log \left(-1+e^{w(x)}\right)-4 w(x)\right) w'(x)^2 v''(x) \lambda
^2 \nn \\ &-& \frac{144 e^{w(x)} \left(1+e^{w(x)}\right) w''(x) v''(x) \lambda ^2}{\left(-1+e^{w(x)}\right)^2}-12 \Bigg(-\frac{3
\left(-5+9 e^{w(x)}\right)}{\left(-1+e^{w(x)}\right)^2}-2 \log \left(-1+e^{w(x)}\right) \nn \\ &+& 6 w(x)\Bigg) v'(x)^2 v''(x)\Bigg].
\label{eq:MISTIMO}
\end{eqnarray}
As an application of \eqref{eq:dop1loop} and \eqref{eq:MISTIMO}, consider the $t^{2,1}$ flow generated by $H_{2,1}$. For reasons that will be clear in a moment, we would like to compute the solutions $v^{[\rm d-op]}(t^{1,0}, t^{2,0}, t^{2,1}, \epsilon)$ and $w^{[\rm d-op]}(t^{1,0}, t^{2,0}, t^{2,1}, \epsilon)$, with all other times set to zero, of the $t^{2,1}$ flow. This will lead us to a proof of Theorem \ref{cor:g=2}. \\
\noindent From \eqref{eq:genintA}, \eqref{eq:unnormff} and \eqref{eq:cv1}-\eqref{eq:cw2} we have that
\beq
h^{[0]}_{2,1}(v,w)=-\frac{v^3}{6 \lambda^2}+v \operatorname{Li}_2(e^w)
\eeq
and therefore
\beq
h^{\rm d-op}_{2,1} := D_{AL} h^{[0]}_{2,1}=h^{[0]}_{2,1}+ \epsilon^2 h^{[2]}_{2,1} + \epsilon^4 h^{[4]}_{2,1}+\mathcal{O}(\epsilon^6),
\eeq
where
\bea
h^{[2]}_{2,1} &=& \frac{e^{w(x)}}{24 \left(-1+e^{w(x)}\right)^2} \Bigg[4 \left(-1+e^{w(x)}\right) w'(x) v'(x) \lambda ^2+v(x) \Big(\left(-1+2 e^{w(x)}\right) \lambda ^2 w'(x)^2 \nn \\ &-& 2 \left(-1+e^{w(x)}\Big) v'(x)^2\right)\Bigg], \\
h^{[4]}_{2,1} &=& -\frac{1}{\lambda^2} \l(v b^{[4]}_{vv} + b^{[4]}_{vvv} \r),
\end{eqnarray}
while on the other hand
\beq
h^{\rm d-op}_{1,0}=h^{[0]}_{1,0} = -\frac{v w}{\lambda^2}.
\eeq
Let us solve the dispersive equations
\beq
\frac{{\partial} u^\alpha}{{\partial} t^{2,1}} = \l\{u^{\alpha},\int_{S^1} h^{\rm
d-op}_{2,1}(v,w)\r\}
\eeq
perturbatively in $t^{2,1}$ with the topological Cauchy datum \eqref{eq:initcond}. We find
\bea
w^{[\rm d-op]}(x, t^{2,0},t^{2,1})&=& t^{2,0}+t^{2,1} x+(t^{2,1})^2 \lambda ^2 \log
\left(1-e^{t^{2,0}}\right) + \frac{e^{t^{2,0}} (t^{2,1})^3 x \lambda ^2}{-1+e^{t^{2,0}}}+\dots \nn \\ &+& \left(-\frac{e^{t^{2,0}} (t^{2,1})^2 \lambda ^2}{12 \left(-1+e^{t^{2,0}}\right)^2}+\frac{e^{t^{2,0}} \left(1+e^{t^{2,0}} \right) (t^{2,1})^3 x \lambda ^2}{12
\left(-1+e^{t^{2,0}}\right)^3} + \dots\right) \epsilon ^2 \nn \\ &+& \left(-\frac{e^{t^{2,0}} \left(1+4 e^{t^{2,0}}+e^{2 t^{2,0}}\right) (t^{2,1})^2 \lambda ^2}{240 \left(-1+e^{t^{2,0}}\right)^4} + \dots\right) \epsilon^4+ \mathcal{O}(\epsilon^6).
\label{eq:wsol2l}
\end{eqnarray}
From now on we put $t^{2,0}=:t$. The last line of \eqref{eq:wsol2l} and the assumption \eqref{eq:wdopwtsqt} combined together lead to
\beq
\frac{{\partial}^4 \mathcal{F}_2}{{\partial} x^2 {\partial} (t^{2,1})^2}\Bigg|_{\substack{t^{\alpha,p}=0 \\ \hbox{\footnotesize for } p>0}}=\frac{e^{t} \left(1+4 e^{t}+e^{2 t}\right) }{120 \left(-1+e^{t}\right)^4}=\frac{1}{120}\operatorname{Li}_{-3}\l(e^{t}\r).
\eeq
In Gromov--Witten theory the left hand side would represent the small phase correlator $\left\langle\bra \phi_1, \phi_1, \tau_1 \phi_2, \tau_1 \phi_2 \right\rangle\ket^{X_1}_{2, \mathrm{sps}}$, where we define
\beq
\left\langle \left\langle \tau_{p_1} \phi_{\alpha_1} \dots \tau_{p_k} \phi_{\alpha_k} \right\rangle \right\rangle^{X_1}_{g, \mathrm{sps}}(x,t) :=
\sum_{d,n \geq 0} \frac{1}{n!} \left\langle \tau_{p_1} \phi_{\alpha_1} \dots \tau_{p_k} \phi_{\alpha_k} \overbrace{ x\phi_1 + t \phi_2 \dots x\phi_1 + t\phi_2}^{\text{$n$ times}} \right\rangle_{g,n+k,d}^{X_1}.
\eeq
Applying twice the puncture equation to $\left\langle\bra \phi_1, \phi_1, \tau_1 \phi_2, \tau_1 \phi_2 \right\rangle\ket^{X_1}_{2, \mathrm{sps}}$ we can kill the two descendent insertions and reduce to the double derivative of the primary potential
\beq
2 \sum_{d \geq 0} \sum_{n=0}^\infty \frac{(t)^n}{n!} \left\langle \phi_2, \phi_2, \overbrace{ \phi_2, \dots, \phi_2}^{\text{$n$ times}} \right\rangle_{2,n+2,d} = 2 \frac{{\partial}^2 F_2(t)}{{\partial} t^2},
\eeq
that is
\beq
\frac{{\partial}^2 F_2(t)}{{\partial} t^2} = \frac{1}{240}\operatorname{Li}_{-3}\l(e^{t}\r).
\label{eq:der2g2}
\eeq
This reproduces exactly the higher genus formula for primary Gromov--Witten
invariants of $X_1$ \cite{Marino:1998pg, Gopakumar:1998jq, Katz:1999xq,
MR1728879, Bryan:2004iq}
\beq
F_g^{X_1}(t)=\sum_{d=0}^\infty N^{(1)}_{g,d} e^{d t}=\frac{\l|B_{2g}\r|}{2g(2g-2)!}\operatorname{Li}_{3-2g}(e^{t})+\frac{\l|B_{2g} B_{2g-2}\r|}{2g (2g-2)(2g-2)!}
\label{eq:primpotggen}
\eeq
at genus 2, up to the constant map contribution, and proves Theorem \ref{cor:g=2}. \begin{flushright}$\square$\end{flushright}
\subsection{Higher descendent flows and the Ablowitz--Ladik equations}
By the same token, the complete solution $w(\mathbf{t})={\partial}^2_x \tilde{\mathcal{F}}(\mathbf{t})$ of all flows should
contain information on descendent invariants; however, the discrepancy between
$\tilde{\mathcal{F}}(\mathbf{t})$ and $\mathcal{F}(\mathbf{t})$, which amounts to constant map terms when restricted
to primaries, might also affect positive degree invariants when it comes to
computing descendents. In particular the terms of $\mathcal{O}(t^{2,1})^{n+2}$ of
\eqref{eq:wsol2l} compute the right genus 2 Gromov--Witten invariants with single descendent
insertions at $n$ points only if $n\leq 2$. As for the genus one case, the precise
choice of dependent variables for the Ablowitz--Ladik equations is then crucial for
the computation of Gromov--Witten invariants, and in particular it should
correct the hydrodynamic Poisson structure \eqref{eq:poissbrack}, which is
left invariant by construction in the $D$-operator formalism, by higher order terms in $\epsilon$.\\
It is nonetheless remarkable that the dispersive Ablowitz--Ladik flows in the
$D$-operator form satisfy a number of constraints induced from the topology of
moduli spaces of stable maps.
As an example, a little experimentation at the next few orders in $t^{2,1}$
shows that
\beq
\label{eq:exdesc1}
\frac{1}{n!} \frac{{\partial} w^{[\rm d-op]}(x, t^{2,0},t^{2,1})}{{\partial} (t^{2,1})^n}\bigg|_{t^{2,1}=0}=
\sum_{k=0}^{n} a''_{k,n}(t) x^k
\eeq
with
\beq
a_{k,n}(t)=\l(\bary{c}n \\ k\eary\r) \frac{{\partial}^{k} a_{0,n-k}(t)}{{\partial} t^k}
\label{eq:stringrel}
\eeq
It is noteworthy that the relation \eqref{eq:stringrel}, which in
Gromov--Witten theory would be a consequence of the string axiom, is realized
by the dispersive Ablowitz--Ladik flows; we checked this up to
$\mathcal{O}((t^{2,1})^7)$ (i.e. $n\leq 5$). Along the same lines, it is straightforward to switch on the $t^{1,1}$-flow of the Ablowitz--Ladik hierarchy and see that the dilaton constraint is satisfied too. As an example, for the $\mathcal{O}((t^{2,1})^2)$, $\mathcal{O}(\epsilon^{2g})$ coefficient $\tilde w^{(2)}_g(t^{1,1},t)$ of $w(x,t,t^{1,1},t^{2,1})$ we can give a closed expression for its $t^{1,1}$ dependence
\bea
\label{eq:exdesc3}
\tilde w^{(2)}_g &:=& \sum_{ n\geq 0} \frac{(t^{1,1})^n}{n!} \left\langle\bra \phi_1,\phi_1, \tau_{1}\phi_2, \tau_{1}\phi_2 , \overbrace{\tau_1\phi_1, \dots, \tau_1\phi_1}^{n \hbox{ \footnotesize \rm times}} \right\rangle\ket^{X_1}_{g, \mathrm{sps}} \nn \\
&=&
\left(\frac{1}{1-t^{1,1}}\right)^{2g+2}
\frac{{\partial}^2}{{\partial} y^2} F_g(y)\Bigg|_{y=\left(\frac{t}{1-t^{1,1}}\right)}, \qquad g=0,1,2
\end{eqnarray}
and it is immediate to see that the small-phase space dilaton equation holds
\beq
\left[(1-t^{1,1})\frac{{\partial}}{{\partial} t^{1,1}}-t \frac{{\partial}}{{\partial} t}-2-2g\right] \tilde w^{(2)}_g = 0.
\eeq
|
\section{Introduction}
The interaction between electrons and ion vibrations is responsible for
many fundamental phenomena in solids. For example, as shown in BCS theory in
electron-phonon(EP) interaction induces a weak attraction between electrons near Fermi
surface with opposite momenta and opposite spins and eventually leads to superconductivity\cite{MAHAN1}.
In quasi one-dimensional materials it is well known that EP
coupling usually causes metal-insulator transition.
In the insulating phase a band gap opens at the Fermi energy and simultaneously lattice is distorted resulting
in a larger unit cell. This transition is called Peierls transition\cite{PEIERLS1}.
For simplicity most of theoretical works consider only harmonic ion vibration
and in only a few works higher orders are taken into account.
For example, the effect of anharmonicity on superconductivity was studied by
Freericks et al.\cite{FREERICKS1} more than ten years ago in infinite-dimensional limit. However, it seems that the
anharmonicity represented by the quartic term does not enhance the
superconductivity transition temperature as expected.
Another example is that polaronic properties with anharmonicity were investigated by
Chatterjee and Takada\cite{CHATTERJEE1} some time ago.
Recently a renewed interest in anharmonic ion vibration was raised in
$\beta$-pyrochlore oxides, especially in KOs$_2$O$_6$.
Experiments show that it has unusual behaviors in both normal
state and superconductivity state\cite{YONEZAWA1,YONEZAWA2, YONEZAWA3,
HIROI1,YAMAURA1}.
In KOs$_2$O$_6$ an Os-O network forms a oversized cage and
K ion sits in the cage. As the mass of K ion is small compared with Rb or Cs it oscillates
relatively far from the equilibrium position. Density functional theory shows
that this movement is highly anharmonic and the harmonic term of the potential
can even be zero or negative\cite{KUNES1, KUNES2}.
Some of the anomalous behaviors observed in KOs$_2$O$_6$ was explained by Dahm and one of
the present authors by treating the anharmonic ion vibration
in a quasi-harmonic approximation with temperature-dependent frequency. \cite{DAHM1,DAHM2}.
The spectral function and NMR relaxation rate have been discussed for a model with a general
anharmonic potential\cite{TAKECHI1}. Otsuka et al\cite{OTSUKA1}
has consider the effect of static anharmonic potential on quasi-one-dimensional extended Hubbard model and rich
phase diagrams have been found at finite temperature. However, in quasi-one-dimensional materials,
phonon fluctuation is usually important\cite{LU1}.
In this paper we will use the one-dimensional spinfull Holstein model as
a prototypical model to study the effects of anharmonicity of ion vibrations.
Since here we are more interested in the anharmonicity the model we will study is given by
\begin{eqnarray}
&&\mathcal{H}=\sum_{i}\frac{p_i^{2}}{2m}+V(r_{1}r_2\cdots r_{L})-t\sum_{i\sigma}(c^{\dagger}_{i\sigma}c_{i+1\sigma}+h.c.)\nonumber \\
&&~~~~~~~~~~-\lambda\sum_{i\sigma} r_{i}(n_{i\sigma}-\frac{1}{2})-\mu\sum_{i\sigma}n_{i\sigma} \label{HAM1}
\end{eqnarray}
where $p_i$ and $r_i$ are the momentum and position operator of ions at site $i$,
$V(r_1r_2\cdots r_L)=\sum_{i=1}^{L}(\frac{K}{2}r^2_i+\gamma{r^4_i})$ is the potential
energy of ion vibration with $L$ being the lattice size. $\gamma$ is always nonnegative so that
the ion potential is bounded. $t$ is the hopping term of electrons onto nearest-neighbor
sites. $\sigma$ is the spin index and in our work
we only consider spinful model and then $\sigma$ takes two values $\sigma=\uparrow,\downarrow$.
$c^\dagger_{i\sigma}(c_{i\sigma})$ is the creation(annihilation)
operator for an electron at
site $i$ with spin index $\sigma$, $n_{i\sigma}=c^\dagger_{i\sigma}c_{i\sigma}$ is the electron
number operator at site $i$ for spin $\sigma$. The electrons couple locally with ions and the
coupling constant is given by $\lambda$. $\mu$ is the chemical potential for controlling the electron number.
If $\mu=0$ Hamiltonian $(\ref{HAM1})$ is invariant under the particle-hole transformation
$p_i\rightarrow -p_i,r_i\rightarrow -r_i,c_i\rightarrow
(-1)^{i}c^\dagger_{i}$
and thus it is always half-filled at $\mu=0$.
If $\gamma=0$ Hamiltonian (\ref{HAM1}) reduces to the {\it{standard Holstein model}}.
In the past several decades it has been extensively studied by many authors. The mean-field theory and the
expansion from the strong coupling limit predict that at half-filling for
spinless case there is a phase transition from a disorder phase to a dimerized phase
while for spinful case the ground state is dimerized for any
finite EP coupling and any finite phonon frequency\cite{HIRSCH1}.
However, the latter conclusion was challenged by Wu et al\cite{Wu1}. Based on functional integral analysis they argued
that the fluctuation of phonons could also destroy the dimerization for spinful case when
$\lambda$ is smaller than a critial $\lambda_c$. This conclusion was confirmed later by
accurate density-matrix renormalization group calculations\cite{JECKELMANN1}
as well as quantum Monte carlo simulations\cite{HARDIKAR1}.
However, it is not clear yet what is the effect of anharmonicity $\gamma\ne 0$ in one dimension.
In particular, if the harmonic term $K$ is zero or even negative, the ion potential
becomes double-well. Then a natural question is whether there are some fundamental changes compared
with the harmonic case?
In the following we will clarify these issues by investegating the
Hamiltonian (\ref{HAM1}) by using the determinant
quantum Monte Carlo simulations\cite{BLANKENBECLER1, SANTOS1}. In this model the electrons of two
spin species couple equally to ions, so there is no sign problem.
In our simulations the largest lattice size is up to 32 and finite-size scaling is
used to extrapolate our results to the thermodynamic limit. The inverse temperature $\beta$ is up to
24. We check our results with different $\beta$ to
confirm that the temperature in our simulation is low enough to extract ground state properties.
The time slice $\Delta \tau=0.1$ is used in most of our simulations and we also check the results
carefully by calculations with smaller $\Delta \tau$ and confirm that the results
are not sensitive to $\Delta \tau$.
In this paper we consider only the half-filled case by putting $\mu=0$. Moreover for simplicity we also
fix $m=0.5$ and $t=1$. In Section II,
firstly we will take $\omega=\sqrt{K/m}=1$, and show the results as a function of $\lambda$.
We find that for both $\gamma=0$ and finite $\gamma$ there is a phase transition
from the disordered phase to the dimerized phase. The dimerization occurs
accompanied by the charge density wave(CDW) transition simultaneously.
Secondly we show the phase diagram in $(\lambda, \gamma)$ plane.
Finally we show the order parameters $m_p$ and $m_e$ as a function of
$K$ with fixed $\lambda=1.4$ and $\gamma=0.1$. Similarly, a
transition from the dimerized CDW phase to the disordered phase is found.
In Section III, we will present results on static charge structure factor and spin structure factor.
These data confirm our previous conclusions. By calculating correlation exponents, we find that
for any finite $\lambda$ there is a spin gap. However, a charge gap opens only after
$\lambda>\lambda_c$, or $K<K_c$. In Section IV, we summarize our results.
\section{Dimerization and charge density wave order parameters}
Since Hamiltonian (\ref{HAM1}) with $\gamma=0$ has been extensively studied
then we will focus on the case with a finite $\gamma$ in this work. Data for $\gamma=0$
are calculated just for comparison with the DMRG results\cite{JECKELMANN1}.
When $\gamma$ is larger than zero the anharmonic term increases the elastic energy of ion.
Therefore a straightforward conjecture is that it will suppress the dimerization as well as the CDW order.
In the first step we examine how the order parameters change as a function of $\lambda$
when the anharmonic term $\gamma$ is introduced. For our purpose we define the
staggered correlation function of lattice displacements
\begin{eqnarray}
D_p(l)=\frac{1}{L}\sum_{i}(-1)^{l}\langle r_{i}r_{i+l}\rangle \label{DPL}
\end{eqnarray}
and the staggered correlation function of the electron densities
\begin{eqnarray}
D_e(l)=\frac{1}{L}\sum_{i}(-1)^{l}\langle n_{i}n_{i+l}-1\rangle \label{DEL}
\end{eqnarray}
The order parameters $m_p$ and $m_e$ are correspondingly defined as
\begin{eqnarray}
m_{p}^{2}=\frac{1}{L}\sum_{l}D_p(l) \label{MP}
\end{eqnarray}
and
\begin{eqnarray}
m_{e}^{2}=\frac{1}{L}\sum_{l}D_e(l) \label{ME}
\end{eqnarray}
In the thermodynamic limit $m_{p}$ is expected to be finite in the dimerized
phase while it is zero in the disorder phase. Similarly, $m_e$ is finite
in the
CDW phase and zero in the disorder phase.
\begin{figure}
\includegraphics[width=6cm, height=7cm, clip]{fig1.eps}
\null \vskip-3mm
\caption{(Color online)Finite-size scaling of square of the order parameters $m_p^{2}$(open symbols) and $m_e^{2}$(filled symbols).
The parameters $(\lambda, \gamma)$ for black circles, red squares, green diamonds and blue triangles are $(0.7,0.0), (1.0, 0.1), (1.0, 0.0)$
and $(2.0,0.1)$, respectively.}
\label{FIG1}
\vskip-2mm
\end{figure}
To be specific we first fix $K=0.5$. In Fig. \ref{FIG1} we show both
$m_p^2$ and $m_e^2$ as a function of inverse length of chain with various $(\lambda,\gamma)$.
Now let us see the data without anharmonicity, i.e., $\gamma=0$.
As shown by circles in (a) both $m_p^2$ and $m_e^2$ clearly extrapolate
to zero in the thermodynamic limit for $\lambda=0.7$ while for
$\lambda=1.0$(shown by diamonds in (b)) the order parameters
are finite. This fact tells clearly that there is at leat one phase transition at $\lambda_c$ satisfying $0.7<\lambda_c<1.0$.
Next we turn to see the data with $\gamma>0$. As a representative example, here we
choose $\gamma=0.1$. As $\lambda=1.0$(shown by squares in (a)) both $m_p^2$ and $m_e^2$ vanish in the thermodynamic limit,
which is different from $\gamma=0$ case. However, as we increase EP
coupling $\lambda$ we reenter in the
dimerized CDW phase, which is shown by triangles in Fig. \ref{FIG1}(b) with $\lambda=2$.
As we have seen, a larger $\lambda_c$ is required when the anharmonicity
$\gamma$ is present and this is consistent with our conjecture.
To extract the critical point $\lambda_c$ accurately we calculate the order parameters
by scaning $\lambda$ with fixed $\gamma$.
The order parameters as a function of $\lambda$ are shown for both
\begin{figure}
\includegraphics[width= 7cm, clip]{fig2.eps}
\null \vskip-4mm
\caption{$m_p$(open symbols) and $m_e$(filled symbols) as a function of $\lambda$ for $\gamma=0$(left) and $\gamma=0.1$(right).
The crosses in the left panel are $m_e\times \lambda/K$. }
\label{FIG2}
\vskip-3mm
\end{figure}
$\gamma=0$ in the left panel of Fig. \ref{FIG2} and $\gamma=0.1$ in right panel of Fig. \ref{FIG2}, respectively.
For $\gamma=0$ the data obtained by the QMC are in good agreement with
those obtained by the DMRG\cite{JECKELMANN1}.
The common feature for both $\gamma=0$ and $\gamma=0.1$ is that
both $m_p$ and $m_e$ vanish at the same critical points, and thus there is one critical point.
The critical values $\lambda_c$ are estimated to be 0.8 and 1.4, respectively.
For $\gamma=0$, $m_p$ and $m_e$ is found numerically\cite{JECKELMANN1}to satisfy
\begin{eqnarray}
m_p=\frac{\lambda}{K}m_e \label{MPME}
\end{eqnarray}
However, this relation does not hold for $\gamma=0.1$. In particular, $m_p/m_e$ is smaller than $\lambda/K$.
It is plausible to say that the anharmonic
potential $\gamma$ has "stronger" suppression on the ion dimerization than on the CDW formation.
In Fig. \ref{FIG3} we show the
schematic phase diagram in $(\lambda,\gamma)$ plane for $K=0.5$. We find that as $\gamma$ increases the critical value $\lambda$ also
increases. This results also support our conclusion that the
anharmonicity suppresses the dimerization and CDW order.
\begin{figure}
\includegraphics[width= 7cm, clip]{fig3.eps}
\null \vskip-4mm
\caption{(Color online) Phase diagram in $(\lambda,\gamma)$ plane for $K=0.5$. The solid line is a guide of eyes.}
\label{FIG3}
\vskip-3mm
\end{figure}
As the coefficient of the harmonic term of the potential may be negative and then the total ion
potential has a form of a double-well,
we also calculate the order parameters as a function of $K$ with $\lambda=1.4$ and $\gamma=0.1$,
which are shown in Fig. \ref{FIG4}.
In this figure we see that when $K<K_c=0.5$ the ground state is in the dimerized CDW phase
while for $K>K_c$ it is in the disorder phase. No new
phases are found in our simulations. Thus we may
conclude that negative $K$ favors the dimerization with the CDW order
but does not introduce new phases in this model in one dimension.
\begin{figure}
\includegraphics[width= 7cm, clip]{fig4.eps}
\null \vskip-4mm
\caption{(Color online)Order parameters $m_p$ and $m_e$ are shown as a function of $K$. Here $\lambda$ is set to be 1.4 and $\gamma=0.1$.}
\label{FIG4}
\vskip-3mm
\end{figure}
\section{Static structure factor and correlation exponents}
One-dimensional correlated electronic systems can usually be described by Tomonaga-Luttinger liquids(TLL)\cite{VOIT1}.
In the TLL theory charge degree of freedoms and spin degree of freedoms are
seperated. In particular, decay of the charge and spin correlation functions are determined
by two exponents $K_\rho$ and $K_\sigma$, respectively.
To explore properties of the disorder phase and the dimerized one further we calculate
the charge structure factor and spin structure factor, which are defined by
\begin{eqnarray}
\mathcal{C}(q)=\frac{1}{L}\sum_{ij}e^{iq(i-j)}(\langle n_i n_j\rangle-1). \label{CK}
\end{eqnarray}
and
\begin{eqnarray}
\mathcal{S}(q)=\frac{1}{L}\sum_{ij}e^{iq(i-j)}S_i^z S_j^z. \label{SK}
\end{eqnarray}
In particular, $K_\rho$ and $K_\sigma$ can be calculated\cite{CLAY1,HARDIKAR1} by
\begin{eqnarray}
K_{\rho}=\pi\lim_{q\rightarrow 0}d\mathcal{C}(q)/dq. \label{KRHO}
\end{eqnarray}
and
\begin{eqnarray}
K_\sigma=4\pi \lim_{q\rightarrow 0}d\mathcal{S}(q)/dq, \label{KSIGMA}
\end{eqnarray}
\begin{figure}
\includegraphics[width= 7cm, clip]{fig5a.eps}
\includegraphics[width= 7cm, clip]{fig5b.eps}
\null \vskip-4mm
\caption{(Color online) Top: Static charge structure factor $\mathcal{C}(q)$ for $\lambda=1.0$.
Inset: $\mathcal{C}(\pi)/L$ is shown as a function of $1/L$.
Bottom: Static charge structure factor for $\lambda=2.0$.
Inset: $\mathcal{C}(\pi)/L$ is shown as a function of $1/L$.}
\label{FIG5}
\vskip-3mm
\end{figure}
In Fig. \ref{FIG5} we show $\mathcal{C}(q)$ as a function of $q$ for $\lambda=1.0$ and $\lambda=2.0$ with
$K=0.5$ and $\gamma=0.1$.
In the top panel we find that $\mathcal{C}(q)$ for $L=16,24$ and $32$ fall almost into one curve and
a peak is present at $q=\pi$. In the inset we show $\mathcal{C}(\pi)/L$ as a function of $1/L$. It vanishes
in the thermodynamic limit.
This figure demonstrates that our sizes are large enough and finite size effect is negligible.
One can conclude that there is only a dominant $2k_F$ short-range correlation and no long-range order.
In the bottom panel it is interesting to notice that $\mathcal{C}(q)$ exhibits a singularity
at $q=\pi$ and $\mathcal{C}(\pi)/L$ is finite in the thermodynamic limit.
This figure indicates that a long-range charge-charge correlation is present.
These two figures are consistent with our conclusion that $\lambda=1$
is in the disordered phase and $\lambda=2$ is in the dimerized CDW phase.
Now we turn to study of the spin correlation functions.
\begin{figure}
\includegraphics[width= 7cm, clip]{fig6.eps}
\null \vskip-4mm
\caption{Static spin structure factors for $\lambda=1.0$ and $\lambda=2.0$ with the lattice size
$L=16, 24$ and $32$. $K$ is chosen to be 0.5 and $\gamma$ is 0.1 and
thus the parameter set used in the top panel
is in the disorder phase and that of the bottom is in the dimerized phase. }
\label{FIG6}
\vskip-3mm
\end{figure}
In Fig. \ref{FIG6} we show the data of static spin structure factor for various lattice sizes
for $\lambda=1.0$ and $2.0$ at $K=0.5$ and $\gamma=0.1$.
First, from the collapse of numerical data of various lattice sizes we may
conclude that finite size effect is negalible in the data.
Secondly we observe that a peak is present at $q=\pi$ for both figures.
Moreover, the height of the peak is found to be finite in both $\lambda=1.0$ and $\lambda=2.0$.
We then conclude that in this model there is a dominant short-range $2k_F$ spin correlation in both the disorder
phase and the dimerized phase.
Next we will analyse the charge and spin correlation exponents $K_\rho$ and $K_\sigma$.
Since our model is invariant under spin rotation, we expect that if it
is in the TLL phase $K_\sigma$ is equal to 1.
However, if it is less than $1$ a spin gap is expected to open\cite{CLAY1,HARDIKAR1}.
This quantity is quite sensitive to the opening of a spin gap and therefore we use it
as a criterion to determine whether there is a spin gap or not.
In the top panel of Fig. \ref{FIG7} we show both $K_\rho$ and $K_\sigma$ as a function of
$\lambda$ at $\gamma=0.1$. The behaviors are actually qualitatively similar
to the results of the standard Holstein model, which are obtained by the
Monte carlo simulations \cite{HARDIKAR1}.
Namely, in the weak coupling region, $K_\rho$ is larger than 1. Moreover, it shows nonmonotonous dependence
on $\lambda$: it increases from 1 as $\lambda$ increases from zero, and after it reaches a maximum it starts to decrease.
In the strong coupling region $K_\rho$ is always smaller than one.
The crossover point with the line $K_\rho=1$ is estimated to be between $\lambda=1.3$ and $\lambda=1.4$.
This value is in agreement with the critical point $\lambda_c=1.4$. We believe the small deviation is
due to numerical errors. In the weak coupling region $K_\rho>1$ is intepreted as
the attraction of effective electron-electron interaction mediated by
phonons while in the strong coupling
region $K_\rho<1$ is intepreted as effective repulsive electron-electron
interaction in the dimerized phase.
However, $K_\sigma$ does not show such crossover and it always smaller than $1$ and
monotonously decreases as a function of $\lambda$. We then conclude that a spin
gap exists for any finite $\lambda$.
In the bottom panel we show $K_\rho$ and $K_\sigma$ as a function of $K$ for $\gamma=0.1$.
In this figure negative $K$ is also considered. Again $K_\rho$ crosses
from below 1 to above 1 as $K$ increases.
The cross point seemingly is consistent with the critical point $K_c=0.5$ within our error bars.
In the same way as before, $K_\sigma$ is small
than $1$ in the whole range, indicating the presence of a spin gap. Compared with
the results obtained by the DMRG\cite{JECKELMANN1} and the
QMC\cite{HARDIKAR1} we may conclude that the disorder phase is Luther-Emery liquid.
\begin{figure}
\includegraphics[width= 6cm, clip]{fig7a.eps}
\includegraphics[width= 6cm, clip]{fig7b.eps}
\null \vskip-4mm
\caption{(Color online) Top: $K_\rho$ and $K_\sigma$ as a function of $\lambda$ for $K=0.5$ and $\gamma=0.1$.
Bottom: $K_\rho$ and $K_\sigma$ as a function of $K$ for $\lambda=1.4$ and $\gamma=0.1$.}
\label{FIG7}
\vskip-3mm
\end{figure}
\section{SUMMARIES}
In summary, in this paper we have investigated the effect of anharmonicity on the
one-dimensional half-filled Holstein model.
By studying order parameters, static charge and spin struture factors and correlation exponents we find
a transition from Luther-Emergy liquid to dimerized CDW phase. This is essentially similar to standard Holstein model.
The effect of anharmonicity is to suppress the dimerization as well as
the charge density wave. Recently, the effective
interaction of electrons mediated by anharmonic phonons has been derived and the transition
temperature of CDW of Hamiltonian (\ref{HAM1}) has been discussed\cite{DIKANDE1}. The transition temperature decreases
monotonously as $\gamma$ increases. This behavior is also consistent
with the present conclusions. On the other hand, a double-well potential with a negative quartic term
favors the dimerization and the CDW order.
\acknowledgements
This work is supported by a Grant-in-Aid for Scientific Research on
Innovative Areas "Heavy Electrons" (No. 20102008) and also by Scientific
Research (C) (No. 20540347). J.Zhao is supported by part by the Japan Society for the Promotion of Science(P07036).
Part of computation in this work has been done using the facilities of the Supercomputer Center, Institute for Solid State
Physics, University of Tokyo.
\vspace{3mm}
|
\section{Introduction}
We consider the following variation of Hilbert's fourth problem:
{\sl to construct all Finsler metrics on the two-sphere whose geodesics are circles.}
Not great circles as in Hilbert's problem, just circles. Our primary goal is to
understand how to construct Finsler surfaces with prescribed geodesics. This question,
addressed by Busemann \cite{Busemann:1939, Busemann:1955, Busemann:1957},
Ambartzumian \cite{Ambartzumian:1976}, Alexander \cite{Alexander:1978},
Matsumoto \cite{Matsumoto:1989}, and Arcostanzo \cite{Arcostanzo:1994} will be completely
solved in Section~3. Our secondary goal is to construct all path geometries on the
two-sphere all of whose paths are circles.
We stress that we work with the classical definition of (reversible) Finsler metrics:
\begin{definition*}
A Finsler metric on a manifold $M$ is a continuous function $F : TM \rightarrow
[0 {\, . \, . \,} \infty)$ that is homogeneous of degree one, smooth outside the zero section
and satisfies the {\sl quadratic convexity} condition: at every tangent space $T_xM$
the Hessian of $F^2(x,\cdot)$ (computed using any affine coordinates on $T_xM$) at
any nonzero tangent vector is a positive-definite quadratic form.
\end{definition*}
Indeed, in studying inverse problems in Finsler geometry our predecessors have either
weakened the condition of quadratic convexity to the condition that the restriction of $F$
to each tangent space be a norm (\cite{Arcostanzo:1994}), considered wider generalizations of
Finsler metrics such as $G$-spaces (\cite{Busemann:1939, Busemann:1955, Busemann:1957}), or
considered integrands that are homogeneous of degree one and not necessarily defined in all
tangent directions (\cite{Matsumoto:1989}). If we insist on the quadratic convexity condition
it is because without it we cannot speak of the geodesic flow of the Finsler metric nor define important
invariants such as the flag curvature (see \cite{Bao-Chern-Shen:2000}).
The paper consists of two parts each describing a geometric construction.
{\noindent \sl Part I. Finsler metrics associated to a family of hypersurfaces.}
Consider a manifold $M$ together with a family of hypersurfaces such that through
every point $x \in M$ and every tangent hyperplane $\zeta_x$ in $T_x M$ there is a unique
hypersurface passing through $x$ tangent to $\zeta_x$. Let us assume that this family is
parameterized by a smooth manifold $\Gamma$ and that the projection that sends a tangent
hyperplane $\zeta_x$ to the unique hypersurface that is tangent to it is a submersion from
the space $PT^* M$ of contact elements on $M$ to the parameter manifold $\Gamma$.
\begin{theorema*}\label{First-construction}
If $\mu$ is a smooth positive measure on $\Gamma$, there is a (unique) Finsler metric
$F : TM \rightarrow [0 {\, . \, . \,} \infty)$ such that for any piecewise smooth curve $c$ on
$M$ we have the Crofton-type formula
\begin{equation}\label{Crofton-formula}
\int \!\! F(\dot{c}(t)) \ dt = \int_{\gamma \in \Gamma} \!\! \#(\gamma \cap c)
\ d\mu(\gamma) .
\end{equation}
\end{theorema*}
In general, the authors do not know what is the precise relationship between the family of
hypersurfaces and the Finsler metrics associated to it. However, in two dimensions the
relationship is simple enough:
\begin{theoremb*}\label{2D-case}
If $M$ is a two-dimensional manifold, the Finsler metrics defined by Eq.~(\ref{Crofton-formula}) are precisely
those whose geodesics are the curves of the family parameterized by $\Gamma$.
\end{theoremb*}
In particular, we will conclude in Section~3 that {\sl any path geometry in a two-dimensional manifold
is locally the system of geodesics of a Finsler metric.}
An interesting remark is that while the construction in Theorem~A extends to families of cooriented hypersurfaces
such as the family of horospheres in hyperbolic space, Theorem~B does not. Indeed, if we allow different curves
of $\Gamma$ to be tangent on the condition that their coorientations at the point of tangency be different,
then not only do those curves fail to be geodesics of the Finsler metric defined by Eq.~(\ref{Crofton-formula}), but
metrics associated to different measures may have different unparameterized geodesics. We shall demonstrate this in
Section~4 by taking $\Gamma$ to be the family of horocycles in the hyperbolic plane and using Eq.~(\ref{Crofton-formula})
to construct two Riemannian metrics --- the hyperbolic metric and a metric conformal to it --- with different sets of
unparameterized geodesics.
\noindent {\sl Part II. Circular path geometries on the two-sphere.}
These are families of circles on $S^2$ such that for every (tangent) line-element there
is a unique circle tangent to it. The standard example is the family of great circles on
the sphere.
The construction of all circular path geometries turns out to be quite simple and elegant.
Let us identify ${\mathbb R}^4$ with the algebra of quaternions and let $\pi : S^3 \rightarrow S^2$
be the Hopf fibration ${\bf q} \mapsto {\bf q} {\bf i} \bar{{\bf q}}$. Remark that any submanifold
$\Sigma$ of the Grassmannian ${\widetilde{\rm G}}_2({\mathbb R}^4)$ of oriented two-planes in ${\mathbb R}^4$ engenders a family
of circles $\{\pi(L \cap S^3) : L \in \Sigma \}$ on the two-sphere.
Identifying, as usual, the Grassmannian ${\widetilde{\rm G}}_2({\mathbb R}^4)$ with the product of two-spheres
$S_{-}^2 \times S_{+}^2$ of radius $1/\sqrt{2}$ in $\Lambda^2({\mathbb R}^4) \cong {\mathbb R}^6$
(see Section~4 or Gluck and Warner~\cite{Gluck-Warner:1983} for details), the submanifolds
of ${\widetilde{\rm G}}_2({\mathbb R}^4)$ that correspond to circular path geometries are very easy to construct:
\begin{theoremc*}
Let $\kappa : S_{+}^2 \rightarrow {\mathbb R}$ be a smooth function. The graph of the map
$f_\kappa : S_{+}^2 \rightarrow S_{-}^2$,
$$
f_\kappa ({\bf x}) := \left(\frac{\kappa({\bf x})}{\sqrt{2 + 2\kappa({\bf x})^2}}, 0,
\frac{1}{\sqrt{2 + 2\kappa({\bf x})^2}} \right),
$$
seen as a submanifold of the Grassmannian ${\widetilde{\rm G}}_2({\mathbb R}^4)$, engenders a circular path geometry
if and only if $\kappa$ is odd and satisfies $\nmx{\nabla \kappa } < 1 + \kappa^2$. Moreover, every circular
path geometry can be obtained this way.
\end{theoremc*}
Let us denote by $\gamma_\kappa({\bf x})$ (${\bf x} \in S_+^2$) the oriented circle on the two-sphere obtained by projecting down the intersection of
the three-sphere with the oriented two-plane $(f_\kappa({\bf x}),{\bf x}) \in S_{-}^2 \times S_{+}^2$ by means of the Hopf fibration. Theorems~B and~C immediately imply the following simple description of all Finsler metrics on $S^2$ whose geodesics are circles.
\begin{theoremd*}
Let $\kappa$ be smooth odd function on $S_{+}^2$ such that $\nmx{\nabla \kappa} < 1 + \kappa^2$ and let $\mu$ be
a smooth positive measure on $S_{+}^2$. The Crofton-type formula
\begin{equation}
\int \!\! F(\dot{c}(t)) \ dt = \int_{{\bf x} \in S_{+}^2} \!\! \#(\gamma_\kappa({\bf x}) \cap c) \ d\mu({\bf x})
\end{equation}
defines a Finsler metric on the sphere whose geodesics are circles. Moreover, every Finsler metric whose geodesics are
circles can be obtained this way.
\end{theoremd*}
\section{Finsler metrics associated to a family of hypersurfaces}
In this section we show how to associate Finsler metrics to certain
families of hypersurfaces which are described by the class of double fibrations
\begin{equation}\label{Double-fibration}
\xymatrix{
& {S^* M} \ar[dl]_{\pi_1}\ar[dr]^{\pi_2} & \\
M & & \Gamma ,}
\end{equation}
where $S^*M$ is the bundle of cooriented contact elements on $M$,
$\pi_1 : S^*M \rightarrow M$ is the canonical projection, $\Gamma$ is an oriented
manifold of the same dimension as $M$, $\pi_2 : S^* M \rightarrow \Gamma$ is a second
{\sl Legendrian fibration\/} (i.e., the fibers of $\pi_2$ are Legendrian submanifolds of
$S^*M$ provided with its canonical contact structure), and the map
$\pi_1 \times \pi_2 : S^* M \rightarrow M \times \Gamma$ is an embedding.
For each $\gamma \in \Gamma$, $\hat{\gamma} := \pi_1(\pi_2^{-1}\{\gamma\})$ is a
cooriented hypersurface. Notice that given a cooriented contact element $\xi \in S^*M$,
the hypersurface corresponding to $\pi_2(\xi) \in \Gamma$ is the unique hypersurface
from the family that is tangent to it and has a matching coorientation.
Notice that the description of families of non-cooriented hypersurfaces such as those appearing in
the introduction can be obtained as a special case of~(\ref{Double-fibration}) by taking double covers.
\noindent {\bf Example.}
The double fibration corresponding to all cooriented hyperplanes in ${\mathbb R}^n$ is
\begin{equation}\label{Standard-double-fibration}
\xymatrix{
& {S^{n-1} \times {\mathbb R}^n} \ar[dl]_{\pi_1}\ar[dr]^{\pi_2} & \\
{\mathbb R}^n & & S^{n-1} \times {\mathbb R} ,}
\end{equation}
where the fibrations are given by $\pi_1(\xi,x) = x$ and
$\pi_2(\xi,x) = (\xi, \xi \cdot x)$.
In order to establish Theorem~A, and its generalization to the cooriented setting, we shall quickly review the basic
ideas around Gelfand transforms, Crofton formulas, and cosine transforms developed in~\cite{Alvarez-Fernandes:2007}.
In what follows {\sl we identify the bundle of cooriented contact elements on $M$, $S^* M$, with a hypersurface of
$T^*M$ such that $S_x^* M$ is a star-shaped hypersurface in $T_x^* M$ for each $x \in M$.}
Given the double fibration~(\ref{Double-fibration}) and a volume form $\Omega$ on $\Gamma$, the Gelfand transform of
the volume density $|\Omega|$ is the function
$$
F := \pi_{1*}(\pi_2^* |\Omega|) : TM \longrightarrow {\mathbb R}
$$
whose value at the tangent vector $v \in T_xM$ is computed as follows: (1) construct a volume density on the fiber $\pi_1^{-1}\{x\} = S^*_x M$ by contracting $\pi_2^* |\Omega|$ at each point $\xi \in S^*_x M$ with any vector $\tilde{v} \in T_{\xi} S^* M$ such that
$D\pi_1(\tilde{v}) = v$; (2) integrate this volume density over $S^*_x M$.
\begin{lemma}\label{Gelfand-transform}
Given the double fibration~(\ref{Double-fibration}) and a volume form $\Omega$ on $\Gamma$, the Gelfand transform
$F := \pi_{1*}(\pi_2^* |\Omega|)$ of the volume density $|\Omega|$ is homogeneous of degree one, smooth and positive
outside the zero section, and satisfies the Crofton-type formula
$$
\int \!\! F(\dot{c}(t)) \, dt = \int_{\gamma \in \Gamma} \!\! \#(\hat{\gamma} \cap c)
\, |\Omega|
$$
for any piecewise smooth curve $c$ on $M$.
\end{lemma}
\noindent {\it Proof.\ } The homogeneity, smoothness, and positivity all follow immediately from the construction. The Crofton-type formula is easily established
once we remark that $\pi_2^* |\Omega|$ defines a measure on the manifold $\pi_1^{-1}(c)$ whose dimension is the same as that of $\Gamma$ and hence, by the simplest case of the co-area formula,
\begin{eqnarray*}
\int_c \pi_{1*}(\pi_2^* |\Omega|) = \int_{\pi_1^{-1}(c)} \pi_2^* |\Omega| &=&
\int_{\gamma \in \Gamma} \! \#(\pi_2^{-1}\{\gamma\} \cap \pi_1^{-1}(c)) \, |\Omega| \\
&=&
\int_{\gamma \in \Gamma} \!\! \#(\hat{\gamma} \cap c) \, |\Omega| \, . \hfill \qed
\end{eqnarray*}
\noindent {\bf Example.}
Let us consider the case of ${\mathbb R}^n$ and its family of cooriented hyperplanes where the
double fibration is given in~(\ref{Standard-double-fibration}). If we write the volume
form $\Omega$ on $S^{n-1} \times {\mathbb R}$ as $ \Omega = \omega \wedge dt$, where $\omega$ is
a volume form on $S^{n-1}$, then it is easy to see that
\begin{equation}\label{cosine-transform}
F(x,v) = \int_{\xi \in S^{n-1}} |\xi \cdot v|\, \omega .
\end{equation}
It is known since the work of Pogorelov on Hilbert's fourth problem
(see \cite{Pogorelov:1979}) that for any choice of volume form $\Omega$ the function $F$
in~(\ref{cosine-transform}) is a Finsler metric all of whose geodesics are straight lines.
For the more general double fibration~(\ref{Double-fibration}), we can say nothing about the
geodesics of $F$, but we can show that $F$ is a Finsler metric.
\begin{theorem}\label{Crofton-construction}
If $\Omega$ is a volume form on $\Gamma$, then $F := \pi_{1*}(\pi_2^* |\Omega|)$ is a
Finsler metric on $M$ which, moreover, satisfies the Crofton-type formula
$$
\int \!\! F(\dot{c}(t)) \, dt = \int_{\gamma \in \Gamma} \!\! \#(\hat{\gamma} \cap c)
\, |\Omega|
$$
for any piecewise smooth curve $c$ on $M$.
\end{theorem}
In order to prove this theorem we first establish a formula for $F$ that
generalizes~(\ref{cosine-transform}).
\begin{lemma}
For every point $x \in M$ there exists a volume form on $\omega_x$ defined on $S^*_x M$
such that the restriction of the function $F := \pi_{1*}(\pi_2^* |\Omega|)$ to the tangent
space $T_x M$ is given by the formula
\begin{equation}\label{Cosine-transform-analogue}
F(x,v) = \int_{\xi \in S^*_x M} |\xi(v)| \, \omega_x .
\end{equation}
\end{lemma}
\noindent {\it Proof.\ }
The key remark is that for every covector $\xi \in S^*_x M$ there exists a non-vanishing
$(n-1)$-form $\omega_{(x,\xi)}$ on $T_{(x,\xi)} S^*M$ such that
$$
\left(\pi_2^* \Omega \right)_{(x,\xi)} = \pi_1^* \xi \wedge \omega_{(x,\xi)} .
$$
Notice that ${\rm Ker}\, \xi$ is the tangent space at $\pi_1(\xi)$ of the hypersurface
corresponding to $\pi_2(\xi) \in \Gamma$. In other words,
${\rm Ker}\, \xi = D\pi_1 ({\rm Ker}\, D_{(x,\xi)} \pi_2)$, or
${\rm Ker}\, \pi_1^* \xi = {\rm Ker}\, D_{(x,\xi)}\pi_1 \oplus {\rm Ker}\, D_{(x,\xi)} \pi_2$.
It is now easy to verify that $\pi_1^* \xi \wedge \pi_2^* \Omega = 0$ and hence, by
Cartan's lemma, $\pi_1^* \xi$ is a linear factor of
$\left(\pi_2^* \Omega \right)_{(x,\xi)}$. Notice that while the form
$\omega_{(x,\xi)}$ is not completely determined, its restriction to
${\rm Ker}\, D_{(x,\xi)}\pi_1 = T_\xi S^*_x M$ is. By performing this construction at every
point $\xi \in S^*_x M$ we obtain a non-vanishing $(n-1)$-form on $S^*_x M$ which we
shall call $\omega_x$. Clearly $\omega_x$ depends smoothly on the parameter $x$.
If $v \in T_x M$ is a tangent vector, the volume density constructed on $S^*_x M$
in order to push forward $\pi_2^* |\Omega|$ is just
$(x,\xi) \mapsto |\xi(v)||\omega_x|$. Since $\omega_x$ never vanishes, we may choose a
suitable orientation for $S^*_x M$ and conclude that
$$
F(x,v) = \pi_{1*}(\pi_2^* |\Omega|) (x,v) = \int_{S^*_x M} |\xi(v)|\omega_x .
$$
\qed
\noindent{\it Proof of Theorem~\ref{Crofton-construction}.\ }
In view of Lemma~\ref{Gelfand-transform} we just need to
verify the quadratic convexity condition. It is here that the relation between Gelfand transforms and Crofton
formulas proves useful through Eq.~(\ref{Cosine-transform-analogue}). Indeed, the quadratic convexity of norms
defined by the cosine transform of smooth positive measures is a classical result which probably goes back to Blaschke,
however we could not resist the temptation of presenting the following short proof.
Using the homogeneity of $F$ the problem reduces to showing that if $(v_1,\dots, v_n)$ are affine coordinates
on $T_x M$, the Hessian of $F(x,\cdot)$ at any nonzero tangent vector $v$ is a positive semi-definite quadratic
form whose kernel is the line spanned by $v$. Letting $(\xi_1, \dots, \xi_n)$ be the dual coordinates in $T^*_x M$, a
smattering of distribution theory yields that
\begin{eqnarray*}
\frac{\partial}{\partial v_i} \int_{\xi \in S^*_x M} |\xi(v)|\, \omega_x
&=& \int_{\xi \in S^*_x M} \xi_i \, {\rm sign}(\xi(v)) \, \omega_x , \\
\frac{\partial^2}{\partial v_i \partial v_j} \int_{\xi \in S^*_x M} |\xi(v)|\, \omega_x
&=& 2 \int_{\xi \in S^*_x M} \xi_i \xi_j \, \delta(\xi(v)) \, \omega_x .\\
\end{eqnarray*}
From this formula, it follows that the quadratic form
$$
(D^2_v F)(w) = 2\int_{\xi \in S^*_x M} (\xi(w))^2 \, \delta(\xi(v)) \, \omega_x ,
$$
is positive semi-definite and its kernel is spanned by $v$. \qed
\section{Metrizations of path geometries on surfaces }
Intuitively, a path geometry on a surface $M$ is a family of curves such that for each
line-element in $PT^*M$ there is a unique curve tangent to it. As in the previous section,
it is more convenient to work ``upstairs" in the space of contact elements of $M$.
\begin{definition}
A {\sl path geometry\/} on a two-dimensional manifold $M$ is a smooth Legendrian
foliation of its space of contact elements $PT^*M$ whose leaves are transverse to
the canonical projection $\pi_1 : PT^*M \rightarrow M$. If $l$ is
a leaf of the path geometry, its projection $\pi_1(l)$ is called a {\sl path.}
\end{definition}
We consider the problem of determining whether a path geometry is the system of geodesics
of a Finsler metric. Since the Hopf-Rinow theorem (trivially) extends to Finsler metrics,
the following example of Howard~\cite{Howard:2005} shows, among other things, that there
are relatively simple path geometries which cannot be the geodesics of a Finsler metric.
\begin{theorem}[Howard~\cite{Howard:2005}]
There exists a complete, torsion-free affine connection on the two-torus $T^2$ such that
locally its geodesics can be made to correspond to lines on the plane, but, globally,
given any point $p \in T^2$, there is a second point $q$ such that there is no unbroken
geodesic joining $p$ and $q$.
\end{theorem}
Nevertheless we have the following result:
\begin{theorem}\label{Metrizable-path-geometries}
If the space of leaves of a path geometry on a surface $M$ is a smooth manifold $\Gamma$
and the natural projection $\pi_2 : PT^*M \rightarrow \Gamma$ is a smooth fibration, then
for each smooth positive measure $\mu$ on $\Gamma$ there is a Finsler metric
$F : TM \rightarrow [0 {\, . \, . \,} \infty)$ that
satisfies the Crofton-type formula
\begin{equation}\label{Crofton-for-surfaces}
\int \!\! F(\dot{c}(t)) \ dt = \int_{\gamma \in \Gamma} \!\! \#(\gamma \cap c)
\ d\mu(\gamma) ,
\end{equation}
and whose geodesics are the paths of the given path geometry. Moreover, any Finsler metric whose space
of geodesics is prescribed by $\Gamma$ can be constructed this way.
\end{theorem}
Locally every path geometry satisfies the hypotheses of
Theorem~\ref{Metrizable-path-geometries}. Indeed, by a result of
Whitehead~\cite{Whitehead:1932} around any point in a manifold $M$ provided with a
path geometry, there exists a {\sl convex neighborhood\/} $U$ where any two distinct
points in $U$ are joined by unique path segment lying entirely in $U$. This allows
us to identify the space of paths lying in $U$ with the (smooth) quotient of the
manifold
$$
\{(x,y) \in \partial U \times \partial U : x \neq y \}
$$
by the equivalence relation $(x,y) \cong (y,x)$. As a consequence,
Theorem~\ref{Metrizable-path-geometries} has the following corollary:
\begin{corollary}
Any path geometry on a surface $M$ is locally the system of geodesics of a Finsler metric.
\end{corollary}
\noindent{\it Proof of Theorem~\ref{Metrizable-path-geometries}.\ }
The Finsler metric $F := \pi_{1*}(\pi_2^* \, \mu)$, constructed as in
Theorem~\ref{Crofton-construction}, automatically satisfies the Crofton-type
formula~(\ref{Crofton-for-surfaces}). In turn, the following classical argument shows
that the existence of a Crofton-type formula implies that paths are geodesics.
Let $p$ and $q$ be points in a convex neighborhood $U$ and let $\gamma_{pq}$ be unique path segment
that lies entirely in $U$ and joins the two points. Notice that if $\gamma' \in \Gamma$
is a path different from the one containing $\gamma_{pq}$, then any connected component of $\gamma' \cap U$ intersects
$\gamma_{pq}$ at most once. Indeed, if it intersects it at two points we would have two distinct path segments inside $U$ joining
the same two points, which contradicts the convexity of $U$. Furthermore, if $\sigma$ is a piecewise smooth curve in $U$
joining $p$ and $q$, any connected component of $\gamma' \cap U$ must intersect $\sigma$ at least as many times as it
intersects $\gamma_{pq}$. Using the Crofton-type formula, we have
$$
{\rm length}(\sigma) = \int_{\gamma' \in \Gamma} \#(\sigma \cap \gamma') \, d\mu
\geq
\int_{\gamma' \in \Gamma} \#(\gamma_{pq} \cap \gamma') \, d\mu
= {\rm length}(\gamma_{pq})
$$
and, therefore, path segments are locally length-minimizing.
Let us now prove the converse. Assume that the family of geodesics of a Finsler metric $F$ is precisely our family of paths and
let us find a measure $\mu$ on $\Gamma$ for which the Crofton-type formula~(\ref{Crofton-for-surfaces}) holds.
Let $\tilde{\Gamma}$ be the double cover of $\Gamma$ which parameterizes the family of paths with co-orientation. The natural
projection $\pi_2 : PT^*M \rightarrow \Gamma$ lifts to a projection $\tilde{\pi}_2 : S^*M \rightarrow \Gamma$, where $S^*M$ is the
bundle of unit covectors. If $\omega$ is the standard symplectic form on $T^*M$ and $i : S^*M \rightarrow T^*M$ is the inclusion,
there is a unique area-form $\tilde{\omega}$ on $\tilde{\Gamma}$ such that the equality
$i^* \omega = \tilde{\pi}_2^* \, \tilde{\omega}$ holds on $S^*M$. By the general Crofton formula in Finsler spaces
(Theorem~5.2 in \cite{Alvarez-Berck:2006}), for any smooth curve $c$
$$
\int \!\! F(\dot{c}(t)) \ dt = \frac{1}{4} \int_{\gamma \in \tilde{\Gamma}} \!\! \#(\gamma \cap c) \, |\tilde{\omega}| .
$$
It follows that $(1/2)|\tilde{\omega}|$ descends to a smooth, positive measure $\mu$ on $\Gamma$ for which the Crofton-type
formula~(\ref{Crofton-for-surfaces}) holds.
\qed
\section{Horocycles and Crofton formulas on the hyperbolic plane}
One remarkable feature of Theorem~\ref{Metrizable-path-geometries} is that different measures in
Eq.~(\ref{Crofton-for-surfaces}) yield different Finsler metrics with the same geodesics. We shall now see in an
example that this is no longer the case for non-reversible path geometries where different cooriented
paths can be tangent on the condition that their coorientations at the point of tangency be different.
Consider ${\mathbb R}^3$ together with the Lorentzian form $[{\bf x},{\bf y}] = -x_1 y_1 - x_2 y_2 + x_3 y_3$. As is well-known,
the hyperbolic plane can be defined as the hypersurface
$$
{\mathbb H}^2 = \{{\bf x} \in {\mathbb R}^3 : [{\bf x},{\bf x}] = 1, x_3 > 0 \}
$$
with the induced Riemannian metric $g_{\bf x}({\bf v},{\bf w}) = -[{\bf v},{\bf w}]$. The unit tangent bundle of ${\mathbb H}^2$ can be
identified with the submanifold
$$
S{\mathbb H}^2 = \{({\bf x},{\bf v}) : {\bf x} \in {\mathbb H}^2, [{\bf x},{\bf v}] = 0, [{\bf v},{\bf v}] = -1 \} \subset {\mathbb R}^3 \times {\mathbb R}^3 \
$$
and the family of horocycles is parameterized by the upper light-cone
$$
{\mathcal H} = \{\mbox{\boldmath $\xi$} \in {\mathbb R}^3 : [\mbox{\boldmath $\xi$},\mbox{\boldmath $\xi$}] = 0, \xi_3 > 0 \}:
$$
to each $\mbox{\boldmath $\xi$} \in {\mathcal H}$ we associate the horocycle $\{{\bf x} \in {\mathbb H}^2 : [\mbox{\boldmath $\xi$}, {\bf x}] = 1 \}$.
The incidence relation associated to the family of horocycles is described by the double fibration
\begin{equation}\label{horocyclic-double-fibration}
\xymatrix{
& {S{\mathbb H}^2} \ar[dl]_{\pi_1}\ar[dr]^{\pi_2} & \\
{\mathbb H}^2 & & {\mathcal H} ,}
\end{equation}
where the fibrations are given by $\pi_1({\bf x},{\bf v}) = {\bf x}$ and
$\pi_2({\bf x},{\bf v}) = {\bf x} + {\bf v}$. Indeed, it is easy to verify that the horocycle associated to ${\bf x} + {\bf v}$ is precisely
the horocycle that passes through ${\bf x}$ and whose tangent is orthogonal to (and cooriented by) the unit vector ${\bf v}$.
The construction of a measure on ${\mathcal H}$ that is invariant under the induced action of isometry group ${\rm Iso}({\Hyp})$ is also
quite simple: the two-form $\left((d\xi_1 \wedge d\xi_2 \wedge d\xi_3) \haken \mbox{\boldmath $\xi$} \right)/[\mbox{\boldmath $\xi$},\mbox{\boldmath $\xi$}]$ in ${\mathbb R}^3$ is manifestly
invariant under Lorentz transformations and restricts to a well-defined volume form $\Omega$ on the upper light cone. Another
way to describe this volume form is as $(d\xi_1 \wedge d\xi_2)/\xi_3$.
In the two propositions that follow we shall prove that the Finsler metrics associated to the measures $|\Omega|$ and $\xi_3 |\Omega| = |d\xi_1 \wedge d\xi_2|$
do not have the same geodesics.
\begin{proposition}
Consider the measures $\Omega$ and $\xi_3 |\Omega| = |d\xi_1 \wedge d\xi_2|$ on the space of horocycles ${\mathcal H}$. The Finsler
metric $F = \pi_{1*}(\pi_2^* |\Omega|)$ is a positive multiple of the hyperbolic metric and the Finsler metric
$\pi_{1*}(\pi_2^* \, \xi_3|\Omega|)$ is conformal to $F$ with conformal factor $x_3$.
\end{proposition}
\noindent {\it Proof.\ }
The ${\rm Iso}({\Hyp})$-equivariance of the double fibration~(\ref{horocyclic-double-fibration}) and the invariance of the measure $|\Omega|$, immediately
imply that $F$ is invariant and hence a positive multiple of the hyperbolic metric. On the other hand
$$
\pi_{1*}(\pi_2^* \, \xi_3|\Omega|) = \pi_{1*}((x_3 + v_3) \, \pi_2^*|\Omega|) = \pi_{1*}(x_3 \, \pi_2^* |\Omega|) + \pi_{1*}(v_3 \, \pi_2^* |\Omega|).
$$
Since $v_3$ is an odd function on $\pi_1^{-1}(x) = S_{\bf x} {\mathbb H}^2$, the second summand is zero and
$\pi_{1*}(\pi_2^* \, \xi_3|\Omega|) = x_3 F$.
\qed
\begin{proposition}
Two Finsler metrics that are conformal by a non-constant factor do not have the same geodesics.
\end{proposition}
\noindent {\it Proof.\ }
Let $F$ be a Finsler metric on a manifold $M$ and let $\nu$ be a non-constant, positive, smooth function on $M$. It is enough to
show that $F$ and $\nu F$ do not have the same geodesics on some open chart where $\nu$ is not constant and, therefore, we may work
with local coordinates. The Euler-Lagrange equations for $\nu F$ are
$$
\frac{d}{dt}\left(\nu \frac{\partial F}{\partial v}\right) - \frac{\partial (\nu F)}{\partial x} = 0 .
$$
Therefore, a path $x(t)$ that is a geodesic for both $F$ and $\nu F$ must satisfy the equations
$$
\frac{d}{dt}\left(\frac{\partial F}{\partial v}\right) - \frac{\partial F}{\partial x} = 0 \mbox{ and }
\frac{d\nu}{dt}\frac{\partial F}{\partial v} - F \frac{\partial \nu}{\partial x} = 0.
$$
However, if we choose initial conditions $(x,v)$ such that $(\partial \nu / \partial x) (x) \neq 0$ and
$(\partial \nu / \partial x) (x) \cdot v = 0$, we see that the second equation cannot be satisfied
unless $v = 0$. \qed
\section{Circular path geometries on the two-sphere}
Our aim in this section is to construct all path geometries in the two-sphere where the
paths are circles. In fact, such {\sl circular path geometries \/} will be identified
with a certain class of great circle fibrations of the three-sphere. The results of
Gluck and Warner~\cite{Gluck-Warner:1983} will then yield a simple construction.
We begin by using the Hopf fibration $\pi : S^3 \rightarrow S^2$,
${\bf q} \mapsto {\bf q} {\bf i} \bar{{\bf q}}$, to establish a simple correspondence between oriented
circles on the two-sphere of purely imaginary quaternions of unit length and a class of
oriented great circles on the three-sphere of unit quaternions.
Let $({\bf q}, \kappa) \in S^3 \times {\mathbb R}$ and consider the oriented great circle
$C({\bf q},\kappa) \subset S^3$ that passes through ${\bf q}$ in the direction of
${\bf q} {\bf u}$, where
$$
{\bf u} = \frac{\kappa {\bf i}}{\sqrt{1 + \kappa^2}} + \frac{{\bf k}}{\sqrt{1 + \kappa^2}} .
$$
It is easy to see that the oriented great circles $C(-{\bf q},\kappa)$ and $C({\bf q},\kappa)$
coincide and that $C({\bf q},\kappa){\bf i}$ is the same great circle, but with
reversed orientation, as $C({\bf q} {\bf i}, -\kappa)$.
\begin{proposition}\label{circle-construction}
The image $c({\bf q}, \kappa)$ of the great circle $C({\bf q},\kappa)$ under the Hopf
fibration is the oriented circle on the two-sphere that passes through
$\pi({\bf q}) = {\bf q} {\bf i} \bar{{\bf q}}$ in the direction of ${\bf q} {\bf j} \bar{{\bf q}}$ and has geodesic
curvature $\kappa$. Moreover, for any oriented circle $c$ on the two-sphere there is a
unique great circle $C({\bf q},\kappa)$ such that $c = c({\bf q}, \kappa)$.
\end{proposition}
\noindent {\it Proof.\ }
Simply put, we will identify circles in $S^2$ with their Frenet frames in $SO(3)$ and then
use the standard identification of the double cover of the group of rotations with the
group of unit quaternions.
If ${\bf x} : (a {\, . \, . \,} b) \rightarrow S^2$ is a regular curve and
${\bf v}(t) = \dot{{\bf x}}(t)/\nmx{{\bf x}(t)}$ is its normalized velocity vector, then the
matrix ${\bf X}(t)$ whose columns are ${\bf x}(t)$, ${\bf v}(t)$, and ${\bf x}(t) \times {\bf v}(t)$
describes a curve in $SO(3)$. The Frenet equations take the form
$$
\dot{{\bf X}}(t) = {\bf X}(t)
\left( \begin{array}{ccc}
0 & -s(t) & 0 \\
s(t) & 0 & -s(t)\kappa(t) \\
0 & s(t)\kappa(t) & 0 \end{array} \right) ,
$$
where $s(t) = \nmx{\dot{{\bf x}}(t)}$ is the speed of the curve and $\kappa(t)$ is its
geodesic curvature. Notice that the curve ${\bf x}(t)$ describes a circle with constant speed
if and only if the matrix on the right is constant and, therefore, if and only if the
Frenet frame $X(t)$ is the left-translation of a one-parameter subgroup on $SO(3)$.
Consider now the map $\sigma : S^3 \rightarrow SO(3)$ that takes the unit quaternion
${\bf q}$ to the rotation $\sigma({\bf q})({\bf x}) = {\bf q} {\bf x} \bar{{\bf q}}$, ${\bf x} \in S^2$. In matrix
form, $\sigma({\bf q})$ is the matrix whose columns are the vectors ${\bf q} {\bf i} \bar{{\bf q}}$,
${\bf q} {\bf j} \bar{{\bf q}}$, and ${\bf q} {\bf k} \bar{{\bf q}}$. The map $\sigma$ is a
two-to-one covering ($\sigma({\bf q}) = \sigma(-{\bf q})$) and a group homomorphism. It
follows that $\sigma$ defines a bijection between great circles in $S^3$ and
left-translates of one-parameter subgroups in $SO(3)$. Moreover, an easy computation
shows that the image of a great circle $C({\bf q},\kappa)$ under $\sigma$ satisfies the
equation
$$
\dot{{\bf X}}(t) = {\bf X}(t)
\left( \begin{array}{ccc}
0 & -2/\sqrt{1 + \kappa^2} & 0 \\
2/\sqrt{1 + \kappa^2} & 0 & -2\kappa(t)/\sqrt{1 + \kappa^2} \\
0 & 2\kappa(t)/\sqrt{1 + \kappa^2} & 0 \end{array} \right) .
$$
In other words, $\sigma(C({\bf q},k))$ is the Frenet lift of a circle of curvature $\kappa$
described with constant speed $2/\sqrt{1 + \kappa^2}$. \qed
Given a circular path geometry on the two-sphere we may construct a foliation of $S^3$ by
oriented great circles as follows: given a point ${\bf q} \in S^3$, consider the unique
(circular) path $c$ on the two-sphere that passes through ${\bf q} {\bf i} \bar{{\bf q}} = {\bf x}$ and is
tangent to the line spanned by ${\bf q} {\bf j} \bar{{\bf q}} = {\bf v}$ at this point. Orient $c$ so that
${\bf v}$ is its normalized tangent vector at ${\bf x}$. If $\kappa$ is the geodesic curvature
of the oriented circle $c$, take $C({\bf q},\kappa)$ to be the leaf of the the foliation that
passes through ${\bf q}$. Another way of understanding this foliation is to remark that the
Frenet lifts of all circles in the path geometry, where each circle is oriented in two
different ways, form a foliation of $SO(3)$ which can be transferred to $S^3$ via the
double cover $\sigma : S^3 \rightarrow SO(3)$.
The (smooth) great circle foliations of the three-sphere associated to paths geometries on the
two-sphere are easy to characterize. First, if we provide $S^3$ with the contact structure
such that the contact hyperplane at the point ${\bf q}$ is the plane spanned by ${\bf q} {\bf i}$ and
${\bf q} {\bf k}$, then all great circles in the foliation are Legendrian. Second, if $C$ is an
oriented great circle in the foliation and $\dot{{\bf q}}$ is its velocity vector
at a point ${\bf q} \in C$, then $\langle \dot{{\bf q}},{\bf q}{\bf k} \rangle > 0$. Third, since we lift
every circle in the path geometry with its two possible orientations, if $C$ is an
oriented great circle in the foliation, then $C {\bf i}$ {\sl with its orientation reversed \/}
is also in the foliation.
Nevertheless, let us forget for an instant the special character of these great circle
foliations and recall Gluck and Warner's simple characterization of (general) great circle
foliations of the three-sphere. The idea is to characterize all submanifolds of
the Grassmannian of oriented two-planes in ${\mathbb R}^4$ that induce great circle foliations
of $S^3$.
Using coordinates associated to the basis of $\Lambda^2 {\mathbb R}^4$ formed by the bivectors
\begin{eqnarray*}
&{\bf v}_1 = ({\bf 1} \wedge {\bf i} - {\bf j} \wedge {\bf k})/2, \
{\bf v}_2 = ({\bf 1} \wedge {\bf j} + {\bf i} \wedge {\bf k})/2, \
{\bf v}_3 = ({\bf 1} \wedge {\bf k} - {\bf i} \wedge {\bf j})/2, \\
&{\bf v}_4 = ({\bf 1} \wedge {\bf i} + {\bf j} \wedge {\bf k})/2, \
{\bf v}_5 = ({\bf 1} \wedge {\bf j} - {\bf i} \wedge {\bf k})/2, \
{\bf v}_6 = ({\bf 1} \wedge {\bf k} + {\bf i} \wedge {\bf j})/2,
\end{eqnarray*}
the set of simple bivectors of unit length is the product of the spheres
\begin{eqnarray*}
S_{-}^2 &=& \{(x_1,x_2,x_3,0,0,0) : x_1^2 + x_2^2 + x_3^2 = 1/2 \} \quad {\rm and} \\
S_{+}^2 &=& \{(0,0,0,x_4,x_5,x_6) : x_4^2 + x_5^2 + x_6^2 = 1/2 \} .
\end{eqnarray*}
Using this identification of ${\widetilde{\rm G}}_2({\mathbb R}^4)$ with $S_{-}^2 \times S_{+}^2$, Gluck and Warner
proved the following
\begin{theorem}[\cite{Gluck-Warner:1983}]\label{Gluck-Warner-thm}
A submanifold of ${\widetilde{\rm G}}_2({\mathbb R}^4)$ corresponds to a smooth fibration of the three-sphere
by oriented great circles if and only if it is the graph of a smooth map
$f : S_{\pm}^2 \rightarrow S_{\mp}^2$ with $\|df\| < 1$.
\end{theorem}
We are now ready to prove the main result in this section.
\begin{theoremc*}
Let $\kappa : S_{+}^2 \rightarrow {\mathbb R}$ be an smooth function. The graph of the map
$f : S_{+}^2 \rightarrow S_{-}^2$,
$$
f({\bf x}) := \left(\frac{\kappa({\bf x})}{\sqrt{2 + 2\kappa({\bf x})^2}}, 0,
\frac{1}{\sqrt{2 + 2\kappa({\bf x})^2}} \right),
$$
seen as a submanifold of the Grassmannian ${\widetilde{\rm G}}_2({\mathbb R}^4)$, engenders a circular path geometry
if and only if $\kappa$ is odd and satisfies $\nmx{\nabla \kappa} < 1 + \kappa^2$. Moreover, every circular
path geometry can be obtained this way.
\end{theoremc*}
\noindent {\it Proof.\ }
We just need to characterize the type of map $f$ appearing in
Theorem~\ref{Gluck-Warner-thm} that corresponds to the class of oriented great
circle fibrations of $S^3$ associated to circular path geometries on $S^2$.
In the fibrations we are considering, an oriented great circle $C$ passing through
a point ${\bf q}$ does so in the direction of ${\bf q} {\bf u}$, where ${\bf u}$ is of the form
$a{\bf i} + b {\bf k}$ with $a^2 + b^2 = 1$ and $b > 0$. Since the coordinates of the unit
bivector $(1/\sqrt{2}) {\bf q} \wedge {\bf q} {\bf u}$ are easily computed to be of the form
$(1/\sqrt{2})(a,0,b,x,y,z)$, where $x^2 + y^2 + z^2 = 1$ we have that the map
corresponding to the fibration must a map $f : S_{+}^2 \rightarrow S_{-}^2$ of the form
$$
f(x) = (\alpha({\bf x}),0,\beta({\bf x})), \quad \alpha({\bf x})^2 + \beta({\bf x})^2 = \frac{1}{2},
\quad \beta({\bf x}) > 0.
$$
Letting $\kappa({\bf x}) = \alpha({\bf x})/\beta({\bf x})$, we have that
$$
f({\bf x}) := \left(\frac{\kappa({\bf x})}{\sqrt{2 + 2\kappa({\bf x})^2}}, 0,
\frac{1}{\sqrt{2 + 2\kappa({\bf x})^2}} \right).
$$
and the condition that $\|df\| < 1$ translates to $\nmx{\nabla \kappa } < 1 + \kappa^2$.
It remains to show that the function $\kappa$ is odd. To see this recall the symmetry
condition: if $C$ is an oriented great circle in the foliation associated to a path geometry,
then $C {\bf i}$ with its orientation reversed is also in the foliation. In coordinates
$(x_1,\dots,x_6)$, the action of right multiplication by ${\bf i}$ on the space of bivectors
is given by
$$
(x_1,x_2,x_3,x_4,x_5,x_6) \mapsto (x_1,-x_2,-x_3,x_4,x_5,x_6) .
$$
Therefore, we must have that $(-\alpha({\bf x}),0,\beta({\bf x}),-{\bf x})$ is in the graph of $f$
and so $\alpha$ is odd, $\beta$ is even, and $\kappa = \alpha/\beta$ is odd.
\qed
\section{Final Remarks}
It is natural to ask how many of the metrics constructed in Theorem~D are actually Riemannian. The answer is {\it not many.} Indeed,
it follows at once from a result of A.~G. Khovanskii~(\cite{Khovanskii:1980}) that if the geodesics of an affine connection on $S^2$
are circles, then there exists a diffeomorphism that sends the family of geodesics to the family of great circles. It follows
then from Beltrami's theorem that the only Riemannian metrics that can be constructed through Theorem~D are of constant curvature. The difference between Riemannian and Finsler geometry responsible for this phenomenon is that the exponential map of a Finsler metric is not $C^2$ on the zero section.
In a similar vein, we may ask for the Riemannian version of Theorem~B: {\sl when is a two-dimensional path geometry locally the family of (unparameterized) geodesics of a Riemannian metric on $M$?} The preliminary problem of determining whether the paths are locally the (unparameterized) geodesics of an affine connection is classical and has been studied by Tresse, Cartan, Bol, and many others. We refer the reader to the book of Arnold~(\cite{Arnold:1988}) and the paper of Bryant~(\cite{Bryant:1997}) for two excellent modern expositions. Even when the family of paths is locally the family of geodesics of an affine connection, there are obstructions and it is only recently that substantial progress has been made by Bryant, Dunajski, and Eastwood~(\cite{Bryant-Dunajski-Eastwood:2008}) towards the solution of this problem.
Another interesting problem is to ascertain up to what point Theorem~B extends to the non-reversible setting. We distinguish two different
questions.
\begin{question}
Given a two-dimensional reversible path geometry, construct all non-reversible Finsler metrics for which the paths
are geodesics.
\end{question}
\begin{question}
Is every non-reversible two-dimensional path geometry locally the family of geodesics of a non-reversible Finsler metric?
\end{question}
The challenge these questions pose stems from the inability of integral-geometric techniques such as cosine transforms and Crofton-type
formulas to yield non-reversible metrics.
The simplest example of a non-reversible path geometry on the plane is the family of circles of a fixed radius. This case has been elegantly solved by Tabachnikov in~\cite{Tabachnikov:2004}. In addition to the non-reversibility of this path geometry, one encounters an
additional phenomenon: the Finsler metrics whose geodesics are circles of a fixed radius can only be defined locally. Indeed, a version of the Hopf-Rinow theorem holds for non-reversible Finsler metrics (see Theorem~6.6.1 in \cite{Bao-Chern-Shen:2000}) and is a clear obstruction
to the existence of globally-defined Finsler metrics whose geodesics are circles of a fixed radius. Thus it seems that in this and other
non-reversible problems it is better to work with the following generalization of Finsler metrics:
\begin{definition*}
A {\sl magnetic Lagrangian} on a manifold $M$ is a continuous function $L : TM \rightarrow {\mathbb R}$ that is positively
homogeneous of degree one, smooth outside the zero section and satisfies the following {\sl quadratic convexity condition:}
the restriction of $L$ to each tangent space $T_xM$ is the support function of a quadratically-convex body in $T_x^*M$.
\end{definition*}
The quintessential magnetic Lagrangian on the plane is
$$
L(x,y;u,v) = \sqrt{u^2 + v^2} + \frac{k}{2}(xv-yu) ,
$$
which models the motion of a particle under a constant magnetic field and whose extremals are circles of a fixed radius.
To conclude, we clarify the relationship between Theorem~D and Theorem~4.2 in~\cite{Alvarez-Duran:2002} where
\'Alvarez Paiva and Dur\'an construct a large class of examples of Finsler metrics on the sphere whose geodesics are circles.
The idea in~\cite{Alvarez-Duran:2002} is quite simple: let us take a smooth positive measure $\mu$ on the space $S^{3*}$ of cooriented
totally geodesic two-spheres in $S^3$ and require that it be invariant under the circle action induced from the Hopf action on $S^3$. By
Pogorelov's solution of Hilbert's fourth problem, the Crofton-type formula
$$
\int \! F(\dot{c}(t) \, dt = \int_{\xi \in S^{3*}} \! \#(\xi \cap c) \, d\mu(\xi)
$$
defines a Finsler metric that is invariant under the Hopf action and for which great circles are geodesics. If $\pi : S^3 \rightarrow S^2$
is the Hopf fibration, we construct a submersive Finsler metric on the two-sphere by requiring that the unit ball at $T_{\pi({\bf q})}S^2$ be
the image under $D\pi$ of the unit ball at $T_{\bf q} S^3$. We refer to \cite{Alvarez-Duran:2002} for the proof that the geodesics of this metric are circles.
We wish to point out that this construction cannot give rise to every Finsler metric whose geodesics are circles. Indeed, the invariance of
the measure $\mu$ under the induced Hopf action implies that this construction depends on a function of two variables, while
Theorem~D shows that the general construction depends on two functions of two variables.
|
\section{Introduction}
Matrix-algebraic questions have often their roots in quantum information theory.
Mutually unbiased bases (MUB) are considered and investigated because of their relation for example to quantum state tomography \cite{wooters} or quantum cryptography \cite{security}.
A collection of MUB can be viewed as a particular example of a quasi-orthogonal
system of subalgebras of $M_n({\mathbb C})$ (in this work by {\it subalgebra} we shall always mean
a $^*$-subalgebra containing $\mathbbm 1\in M_n({\mathbb C})$; for definition and details on quasi-orthogonality see the next section). In algebraic terms, it is a quasi-orthogonal system of
{\it maximal abelian} subalgebras (MASAs).
Recently research has began in the non-commutative direction \cite{petz,petzkhan,opsz,ohno,pszw}, too.
(Note that in some of these articles instead of ``quasi-orthogonal'' the term ``complementary'' is used.)
Indeed, it should not be the commutativity of subalgebras deciding wether something
deserves to be studied or not.
From the point of view of quantum physics, the interesting quasi-orthogonal systems and
decompositions are those that
contain factors and MASAs only. (Factors are related
to subsystems and MASAs are
related to maximal precision measurements.)
An example for a quantum physics motivated quasi-orthogonal system
which is composed of both abelian and non-abelian algebras
is the collection of following $3$ subalgebras
of $M_2({\mathbb C})\otimes M_2({\mathbb C})\equiv M_4({\mathbb C})$
(i.e.\! the algebra of $2$ quantum bits):
$M_2({\mathbb C})\otimes \mathbbm 1$ (the algebra associated
to the first qbit), $\mathbbm 1\otimes M_2({\mathbb C})$ (the algebra associated
to the second qbit), and the maximal abelian subalgebra associated to
the so-called {\it Bell-basis} (which plays an important role
e.g.\! in the protocol of {\it dense-coding}).
Existential and constructional questions are already difficult in the
abelian case. We know many things when the dimension is a power of
a prime \cite{ivanovic,woofie}, but for example it is still a question,
whether in $6$ dimensions there exists a collection of $7$
MUB or not \cite{h6,matemisi}.
Little is known when not all subalgebras are assumed to be maximal abelian.
What are the existing constructions
and established obstructions (that is, reasons preventing the existence of
certain such systems)? Of course there are some trivial necessary conditions (that will be
discussed later). Considering systems containing not only factors and MASAs,
it is easy to see, that in general these conditions, alone,
cannot be also sufficient (see the example given in section \ref{sec:pre:decomp}).
However, up to the knowledge of the author, previous to
this work, nontrivial obstructions regarding ``interesting'' systems
were only found in very small dimensions
(namely in dimension $4$, see \cite{petzkhan, opsz, pszw}), using --- in part --- some
rather explicit calculations. Moreover, existing constructions such as the ones in
\cite{opsz,ohno} are usually carried out in prime-power dimensions, only.
Thus there is a wide gap between constructions and obstructions where
``anything could happen''.
The aim of this work is to shorten this gap. In particular,
we shall exclude the existence of some interesting systems
(and moreover, we shall do so not only in some low dimensions).
This paper is organized as follows. First, ---
partly for reasons of self-containment, partly for fixing notations --- a quick overview
(including a presentation of the known results) is given about
quasi-orthogonality, quasi-orthogonal systems
and quasi-orthogonal decompositions. Though it is
well-known to experts, certain parts
--- at least, up to the knowledge of the author --- have never been
collected together. In particular, $3$ conditions will be singled out and
listed as ``trivial necessary conditions'' of existence for a system.
Then in section \ref{sec:M_n} we consider decompositions of $M_n({\mathbb C})\otimes M_n({\mathbb C}) =
M_{n^2}({\mathbb C})$. The tensorial product $M_n({\mathbb C})\otimes M_n({\mathbb C}) \equiv M_{n^2}({\mathbb C})$
appears in quantum physics when one deals with a bipartite system composed
of two equivalent parts. Of course
$M_{n^2}({\mathbb C})$ has many subfactors isomorphic to $M_n({\mathbb C})$ --- in physics
such a subfactor may stand for a subsystem; for example $M_n({\mathbb C})\otimes
\mathbbm 1$ stands for the first part of the bipartite system. It seems
therefore a natural question to investigate quasi-orthogonal
decompositions of $M_{n^2}({\mathbb C})$ into subfactors isomorphic to
$M_n({\mathbb C})$ and a number of MASAs. (As was mentioned,
MASAs are related to maximal precision measurements.)
We shall show that there is no such decomposition in
which there would be only $1$ factor (with the other algebras being
maximal abelian) and neither there are decompositions with $3$ factors.
(Note that with $2$ factors there are decompositions, see
\cite[Theorem 6]{pszw}, for example.) As far as
the author knows, this is the first example\footnote{In reality ---
though in a somewhat implicit manner --- another work \cite{lmsweiner} of the
present author has already dealt with the case of a single factor; see the
remark after corollary \ref{1factor}.
However, the non-existence of this kind of decomposition was never stated there ---
that paper had a different aim.}
for excluding the existence of some ``interesting'' quasi-orthogonal systems
(whose existence cannot be ruled out by the trivial necessary conditions)
in an infinite sequence of higher and higher dimensions.
We shall deal with these cases using a recent result
\cite{ohnopetz}, by which if we replace
each subalgebra in such a decomposition with its commutant, we again get
a quasi-orthogonal decomposition.
However, this is something rather
particular: in general, if two subalgebras are quasi-orthogonal, their
commutants will not remain so.
To study the relation of the commutants,
in section \ref{sec:formula} for two subalgebras ${\mathcal A},{\mathcal B} \subset M_n({\mathbb C})$
we shall take the corresponding
trace-preserving expectations $E_{\mathcal A},E_{\mathcal B}$ and consider the quantity
\begin{equation}
c({\mathcal A},{\mathcal B}) : = {\rm Tr}(E_{\mathcal A} E_{\mathcal B})
\end{equation}
where $E_{\mathcal A} E_{\mathcal B}$ is viewed as an $M_n({\mathbb C}) \rightarrow M_n({\mathbb C})$ linear
map (and hence its trace is well-defined). Then $c({\mathcal A},{\mathcal B})\geq 1$ and
equality holds if and only if ${\mathcal A}$ and ${\mathcal B}$
are quasi-orthogonal.
Thus $c({\mathcal A},{\mathcal B})$ measures how much ${\mathcal A}$ and ${\mathcal B}$ are (or: how
much they are {\it not}) quasi-orthogonal. We shall prove,
that if ${\mathcal A}$ and ${\mathcal B}$ satisfy a certain homogenity condition (which is
always satisfied, if they are
factors or maximal abelian subalgebras) then
\begin{equation}
c({\mathcal A}',{\mathcal B}') = \frac{n^2}{{\rm dim}({\mathcal A}) {\rm dim}({\mathcal B})} c({\mathcal A},{\mathcal B}).
\end{equation}
Finally, in the last section we shall show in some concrete examples how the derived
formula can be used to generalize our earlier arguments and thus to exclude the existence of some further quasi-orthogonal systems.
In some sense our examples will fall ``close" to the cases dealt
with in section \ref{sec:M_n}. However, in contrast to those cases, here the commutants will not
remain (exactly) quasi-orthogonal; so instead of ``exact" statements we shall rely on
our quantitative formula.
\section{Preliminaries}
\label{sec:preliminaries}
\subsection{Quasi-orthogonality}
There is a natural scalar product on $M_n({\mathbb C})$ (the so-called
{\it Hilbert-Schmidt} scalar product) defined by the formula
\begin{equation}
\langle A,B\rangle = {\rm Tr}(A^*B)\;\;\;\;\;\;\;\;\; (A,B\in M_n({\mathbb C})).
\end{equation}
Thus if ${\mathcal A}\subset M_n({\mathbb C})$ is a linear subspace, it is
meaningful to consider the ortho-projection $E_{\mathcal A}$ onto ${\mathcal A}$. When ${\mathcal A}$
is actually a $*$-subalgebra containing $\mathbbm 1 \in M_n({\mathbb C})$ (or in
short: a subalgebra), $E_{\mathcal A}$ coincides with the so-called
{\it trace-preserving conditional expectation} onto ${\mathcal A}$.
Two subalgebras ${\mathcal A},{\mathcal B}\subset M_n({\mathbb C})$, as linear subspaces, cannot be
orthogonal, since ${\mathcal A}\cap {\mathcal B}\neq \{0\}$ as $\mathbbm 1\in {\mathcal A}\cap {\mathcal B}$. At
most, the subspaces ${\mathcal A}\cap \{\mathbbm 1\}^\perp$ and ${\mathcal B}\cap \{\mathbbm
1\}^\perp$ can be orthogonal, in which case we say that ${\mathcal A}$ and ${\mathcal B}$
are {\bf quasi-orthogonal}.
Note also that $A\in M_n({\mathbb C})$ is orthogonal to $\mathbbm 1$ if and only
if ${\rm Tr}(A)=0$ and so the subspace ${\mathcal A}\cap \{\mathbbm 1\}^\perp$ is
simply the ``traceless part'' of ${\mathcal A}$. In other words, ${\mathcal A}$ and ${\mathcal B}$ are
quasi-orthogonal if and only if their traceless parts are orthogonal.
For an $X\in M_n({\mathbb C})$ denote its traceless part by $X_0$; that is,
\begin{equation}
X_0=X-\tau(X)\mathbbm 1
\end{equation}
where $\tau = \frac{1}{n}{\rm Tr}$ is the {\bf normalized trace}. (Note
that the normalization is done in such a way that $\tau(\mathbbm 1)=1$.)
Then the traceless parts $A_0,B_0$ of $A,B\in M_n({\mathbb C})$ are orthogonal
if and only if
\begin{equation}
0=\tau(A_0^*B_0)=\tau((A^*-\overline{\tau(A)}\mathbbm
1)(B-\tau(B)\mathbbm 1) = \tau(A^*B)-\overline{\tau(A)}\tau(B),
\end{equation}
that is, if and only if $\tau(A^*B)=\tau(A^*)\tau(B)$. So, since if
$A$ is an element of the subalgebra, then so is $A^*$, we have that
two subalgebras ${\mathcal A},{\mathcal B}$ of $M_n({\mathbb C})$ are quasi-orthogonal if and only if
for all $A\in {\mathcal A}$ and $B\in {\mathcal B}$,
\begin{equation}\label{product-trace}
\tau(A B) = \tau(A)\tau(B).
\end{equation}
\subsection{Factors, abelian subalgebras and MUB}
For any subalgebra ${\mathcal A}\subset M_n({\mathbb C})$ one can consider its {\bf
commutant}
\begin{equation}
{\mathcal A}'\equiv \{X\in M_n({\mathbb C})| \,\forall A\in {\mathcal A}: \, AX-XA = 0\}
\end{equation}
which is again a subalgebra. One has that the {\bf second commutant}
${\mathcal A}''\equiv ({\mathcal A}')'={\mathcal A}$. A subalgebra ${\mathcal A}$ whose center
\begin{equation}
{\mathcal Z}({\mathcal A})={\mathcal A}\cap {\mathcal A}'
\end{equation}
is trivial (i.e.\! such that ${\mathcal Z}({\mathcal A})={\mathbb C}\mathbbm 1$) is called a {\bf
factor}. If ${\mathcal A}\subset M_n({\mathbb C})$ is a factor, then there exist $j,k$ natural
numbers such that $jk=n$, and that up to unitary equivalence, ${\mathcal A}$ is of
the form
\begin{equation}
{\mathcal A}=M_j({\mathbb C})\otimes \mathbbm 1 \equiv \{ A\otimes \mathbbm 1 | A\in
M_j({\mathbb C})\} \subset M_j({\mathbb C})\otimes M_k({\mathbb C})\equiv M_{jk}({\mathbb C}).
\end{equation}
Then ${\mathcal A}'=\mathbbm 1 \otimes M_k({\mathbb C})$ and so if ${\mathcal A}$ is factor, then
${\mathcal A}$ and ${\mathcal A}'$ are always quasi-orthogonal; this follows easily from the
trace-criterion (\ref{product-trace}) and the fact that
\begin{equation}
{\rm Tr}(A\otimes B) = {\rm Tr}(A) {\rm Tr}(B)
\end{equation}
for all $A\in M_j({\mathbb C})$ and $B\in M_k({\mathbb C})$.
Another example of quasi-orthogonal subalgebras comes from mutually
unbiased bases. Two orthonormed bases ${\mathcal E}=(\mathbf e_1, \ldots, \mathbf
e_n)$ and ${\mathcal F}=(\mathbf f_1, \ldots, \mathbf f_n)$ in ${\mathbb C}^n$ such that
\begin{equation}
|\langle \mathbf e_k,\mathbf f_j \rangle| = {\rm constant} =
\frac{1}{\sqrt{n}}
\end{equation}
for all $k,j=1,\ldots, n$, are said to be {\bf mutually unbiased}, or in
short, ${\mathcal E}$ and ${\mathcal F}$ is a pair of MUB.
Clearly, unbiasedness does not depend on the order of vectors in ${\mathcal E}$
and ${\mathcal F}$, nor on their ``phase factors''. (That
is, the MUB property is not disturbed by replacing a basis vector $\mathbf
v$ by $\lambda \mathbf v$, where $\lambda\in {\mathbb C}, |\lambda|=1$.) For
this reason, one often associates subalgebras to
these bases (which do not depend on the order of vectors and their
phases) and then works with them rather than with the actual bases.
Let us see how can we assign a subalgebra to an orthonormed basis ${\mathcal E}$.
For a vector $\mathbf v\neq 0$, denote the ortho-projection onto the
one-dimensional subspace ${\mathbb C}\mathbf v$ by $P_{\mathbf v}$. Then the
linear subspace of $M_n({\mathbb C})$
\begin{equation}
{\mathcal A}_{\mathcal E} \equiv {\rm Span}\{\, P_{\mathbf e_j}\, |j=1,\ldots, n\}
\end{equation}
is actually a subalgebra. Infact it is a {\bf maximal abelian subalgebra}
(in short: a MASA), and every MASA of $M_n({\mathbb C})$ is of this form.
Elementary calculation shows that if $\mathbf v, \mathbf w$ are vectors of
unit length then
\begin{equation}
{\rm Tr}(P_{\mathbf v}P_{\mathbf w}) =
|\langle \mathbf v,\mathbf w\rangle |^2.
\end{equation}
Hence by an application of the trace-criterion (\ref{product-trace}) one
has that ${\mathcal E}$ and ${\mathcal F}$ is a pair of MUB if and only if the associated
maximal abelian subalgebras ${\mathcal A}_{\mathcal E}$ and ${\mathcal A}_{\mathcal F}$ are quasi-orthogonal.
A famous question concerning MUB is: how many orthonormed
bases can be given in $n$ dimensions in such a way that any two of
the given collection is a MUB? There is a simple bound concerning this
maximum number --- which we shall
denote by $N(n)$ --- namely, that if $n>1$ then $N(n)\leq n+1$. Let us
recall now how this bound can be obtained by a use of the above
introduced subalgebras.
The traceless part of $M_n({\mathbb C})$ is $n^2-1$ dimensional, whereas the
traceless part of a maximal abelian subalgebra of $M_n({\mathbb C})$ is
$n-1$ dimensional. If $n>1$, then at most
\begin{equation}
\frac{n^2-1}{n-1}=n+1
\end{equation}
$n-1$-dimensional orthogonal subspaces can be fitted in an $n^2-1$
dimensional space, implying that for $n>1$ we have $N(n)\leq n+1$.
It is known by construction \cite{ivanovic,woofie} that if $n$ is a power of a prime,
then $N(n) = n+1$. However, apart from
$n=p^k$ (where $p$ is a prime), there is no other dimension $n>1$ in
which the value of $N(n)$ would be known. In particular,
already the value of $N(6)$ is an open question with a long literature on
its own; see e.g.\! \cite{h6,matemisi} . All we know is that $3\leq
N(6)\leq 7$ with numerical evidence \cite{numerical} indicating
that $N(6)$ is actually $3$.
\subsection{Quasi-orthogonal systems and decompositions}
\label{sec:pre:decomp}
A collection of pairwise quasi-orthogonal subalgebras
${\mathcal A}_1,{\mathcal A}_2,\ldots$ of $M_n({\mathbb C})$ is said to form a {\bf
quasi-orthogonal system} in $M_n({\mathbb C})$.
If in addition the given subalgebras linearly span the full space
$M_n({\mathbb C})$, we say that the collection is a {\bf quasi-orthogonal
decomposition} of $M_n({\mathbb C})$.
Suppose we are looking for a quasi-orthogonal system in
$M_n({\mathbb C})$ (or quasi-orthogonal {\it decomposition} of $M_n({\mathbb C})$)
${\mathcal A}_1,\ldots, {\mathcal A}_k$ such that ${\mathcal A}_j$ is isomorphic to ${\mathcal B}_j\;
(j=1,\ldots,k)$, where ${\mathcal B}_1,\ldots,{\mathcal B}_k$ are given matrix
algebras. For example, we may look for a quasi-orthogonal system in which
each algebra is a MASA --- as is the case when we want to find a
collection of MUB --- or, motivated by the study of ``quantum bits'' we may look for a system consisting of subalgebras all isomorphic to $M_2({\mathbb C})$ --- as is investigated in
\cite{petzkhan,opsz}.
What can we say about the existence of a specific system? Some necessary
conditions are easy to establish. In particular, the following three
will be referred as the ``trivial neccessary conditions''
for the existence of a specific quasi-orthogonal system in $M_n({\mathbb C})$ (or:
quasi-orthogonal decomposition of $M_n({\mathbb C})$).
\begin{itemize}
\item[(1)] $M_n({\mathbb C})$ must contain {\it some} subalgebras
${\mathcal A}_1,\ldots, {\mathcal A}_k$ isomorphic
to the given algebras ${\mathcal B}_1,\ldots, {\mathcal B}_k$, respectively,
\item[(2)] the product ${\rm dim}({\mathcal B}_i) {\rm dim}({\mathcal B}_j)\leq n^2$
for all $1\leq i<j \leq k$,
\item[(3)] $\sum_{j=1}^k ({\rm dim}({\mathcal B}_j)-1) \leq n^2-1$ and a
corresponding quasi-orthogonal system is a quasi-orthogonal {\it
decomposition} if and only if in the above formula equality holds.
\end{itemize}
The first condition does not require too much explanation. Nevertheless,
it rules out the existence of various quasi-orthogonal systems. For
example, can we have a quasi-orthogonal system in $M_5({\mathbb C})$ consisting of
$3$ subalgbebras each of which is isomorphic to $M_2({\mathbb C})$? Clearly no:
simply, $M_5({\mathbb C})$ does not contain any subalgebra that would be
isomorphic to $M_2({\mathbb C})$ since $2$ does not divide $5$.
The second condition, at first sight, is perhaps less evident;
let us see now why is it necessary. Suppose ${\mathcal A}$ and ${\mathcal B}$ are
quasi-orthogonal subalgebras of $M_n({\mathbb C})$. Let $A_1,\ldots A_{d_{\mathcal A}}$
and $B_1,\ldots, B_{d_{\mathcal B}}$ be orthonormed bases in ${\mathcal A}$ and ${\mathcal B}$
(with $d_{\mathcal A},d_{\mathcal B}$ standing for the dimensions of ${\mathcal A}$ and ${\mathcal B}$),
respectively. Then, by definition of the (Hilbert-Schmidt) scalar
product and by the trace property (\ref{product-trace}) we have that
\begin{eqnarray}\label{product-basis}
\nonumber
n \langle A_iB_j, A_{i'}B_{j'}\rangle &=&
n^2 \tau( (A_iB_j)^* A_{i'}B_{j'})=
n^2 \tau(A_i^*A_{i'}B_{j'}B_j^*)
= n^2 \tau(A_i^*A_{i'})\tau(B_{j'}B_j^*) \\
&=&
n^2 \tau(A_i^*A_{i'})\tau(B_j^*B_{j'})=
\langle A_i, A_{i'}\rangle\,
\langle B_j, B_{j'}\rangle,
\end{eqnarray}
showing that $\sqrt{n}A_iB_j\; (i=1,\ldots,d_{\mathcal A}; \, j=1,\ldots,d_{\mathcal B})$ is an
orthonormed system in $M_n({\mathbb C})$. Hence the number of members in this system
must be less or equal than the dimension of the full space $M_n({\mathbb C})$; that is,
$d_{\mathcal A} d_{\mathcal B}\leq n^2$.
The third condition is necessary simply because if ${\mathcal A}_1,\ldots, {\mathcal A}_k$ are
quasi-orthogonal, then their traceless parts are orthogonal
subspaces in the traceless part of $M_n({\mathbb C})$. We argue exactly
like we did at discussing the maximum number of MUB
(which, for us, is just a particular case): we have $k$ orthogonal
subspaces of dimensions ${\rm dim}({\mathcal A}_j)-1\; (j=1,\ldots,k)$ in a
${\rm dim}(M_n({\mathbb C}))-1=n^2-1$ dimensional space, implying the claimed
inequality. Moreover, the subspaces span the full space (i.e.\!
we have a quasi-orthogonal {\it decomposition}) if and only if the
dimensions add up exactly to $n^2-1$.
So these conditions are necessary for existence. But are they also sufficient?
The answer, in general, is not.
\bigskip
\noindent
{\it Example.}
Can we find a quasi-orthogonal system in $M_{2n}({\mathbb C})$ consisting of
an abelian subalgebra ${\mathcal A}$ of
dimension $n+1$ and a factor ${\mathcal B}$ isomorphic to $M_n({\mathbb C})$?
If $n>2$, the answer is: not. Indeed, assume by contradiction that ${\mathcal A},{\mathcal B}$
is such a pair. Let $P_1,\ldots ,P_{n+1}$ be the minimal projections of ${\mathcal A}$.
Since we are in a $2n$-dimensional space, at least one of them is a projection onto
a one-dimensional space. So suppose $P_k$ is the orthogonal projection onto the subspace
generated by the unit-length vector $x$. Then by the trace property (implied by quasi-orthogonality)
\begin{equation}
\langle x, Bx\rangle = {\rm Tr}(P_k B) = \frac{1}{2n} {\rm Tr}(P_k){\rm Tr}(B) = \frac{1}{2n}{\rm Tr}(B)
\end{equation}
for all $B\in {\mathcal B}$. This shows that the linear map $B\mapsto Bx$ is injective
on ${\mathcal B}$. Indeed, if $Bx=0$ then $0=\|Bx\|^2= \langle Bx,Bx\rangle =
\langle x, B^*Bx\rangle$, which by the above equation would mean that
${\rm Tr}(B^*B)=0$, implying that $B=0$. However, this is a contradiction, as
the dimension of ${\mathcal B}$ is bigger than the dimension of the full
space: $n^2>2n$. Yet the listed necessary conditions would allow the existence of
such a system. Indeed, the first condition is trivially satisfied, the second is satisfied
as ${\rm dim}({\mathcal A}) {\rm dim}({\mathcal B}) = (n+1)n^2< (2n)^2$, whereas the third is satisfied
since $({\rm dim}(A)-1)+({\rm dim}({\mathcal B})-1) = n+n^2-1 < (2n)^2-1$.
\smallskip
\noindent
Since our motivation is quantum information theory, we are mainly interested
by quasi-orthogonal systems formed by maximal abelian subalgebras and factors.
For such systems it is somewhat more difficult to show that the trivial necessary
conditions are not also sufficient. Let us continue now by discussing
the known examples of ``interesing'' systems.
The trivial necessary conditions allow the existence of a
quasi-orthogonal system in $M_n({\mathbb C})$ composed of $k$ MASAs
as long as $k\leq n+1$ (see the third condition). Moreover,
a quasi-orthogonal system composed of exactly $n+1$ MASAs
would give a quasi-orthogonal decomposition of $M_n({\mathbb C})$.
As was mentioned, the existence of such systems is a popular research
theme (though the problem is usualy considered rather in terms of
MUB than MASA), and little is known when $n$ is not a power of a prime.
Another, more recent problem is to find a system of
quasi-orthogonal subalgebras in $M_{2^k}({\mathbb C})$
in which all subalgebras are isomorphic to $M_2({\mathbb C})$. Here there
is a more direct motivation: a quantum bit, in some sense, is
a subalgebra isomorphic to $M_2({\mathbb C})$, whereas the full algebra
$M_{2^k}({\mathbb C})\simeq
M_2({\mathbb C})\otimes M_2({\mathbb C})\otimes\ldots $ is used in the description of
the register of a quantum computer containing $k$ quantum bits.
In this case too, the first two trivial necessary conditions are
automatically satisfied, whereas the third one says that
such a system can consists of at most $(2^{2k}-1) / 3 =:S(k)$
subalgebras. Again, exactly $S(k)$ such subalgebras would give a
quasi-orthogonal decomposition. (It is easy to see that $S(k)$
is an integer.) In \cite{opsz} $S(k)-1$ such subalgebras are
presented by a construction using induction on $k$. For
$k>2$ it is not known whether the construction is {\it optimal};
that is, whether the upper bound $S(k)$ could be realized or
not. However, it is proved \cite{petzkhan} that for $k=2$ --- i.e.\! in
$M_4({\mathbb C})$ --- the construction is indeed optimal: there is no
quasi-orthogonal system consisting of $S(2)=5$ subalgebras
isomorphic to $M_2({\mathbb C})$. This shows that the listed trivial
necessary conditions, even in the special case of our interest, are not always
sufficient, too. (As far as the author of this
work knows, this was the first example of an ``interesting'' quasi-orthogonal
system satisfying the trivial conditions, whose existence was
{\it disproved}.)
The case of $M_4({\mathbb C})$ has received quite a bit of attention
\cite{petzkhan,opsz,pszw}. Indeed, this is the smallest dimension in which
--- at least from our point of view --- something nontrivial
is happening. As we are interested by factors and MASAs,
let us consider a quasi-orthogonal decomposition
of $M_4({\mathbb C})$ consisting of a collection of MASAs
(so subalgebras isomorphic to ${\mathbb C}^4$) and proper subfactors (so
subalgebras isomorphic to $M_2({\mathbb C})$). Again, the
first two trivial necessary conditions are automatically satisfied,
whereas dimension counting (third condition) says that for such a
decomposition we need $5$ subalgebras. The trivial necessary conditions
do not give anything more. However, in \cite{pszw} it was proved that such
a decomposition exists if and only if an even number of these $5$
subalgebras are factors. So for example one can construct such a
decomposition with $3$ factors and $2$ MASAs, but
not with $2$ factors and $3$ MASAs. This again shows that
the trivial necessary conditions are not always sufficient, too.
The problem with quasi-orthogonal copies of $M_2({\mathbb C})$ in $M_{2^k}({\mathbb C})$
can be also generalized in the sense that one may look for
quasi-orthogonal copies of $M_n({\mathbb C})$ in $M_{n^k}({\mathbb C})$. If $n$ is
a power of a non-even prime, then $M_{n^k}({\mathbb C})$ admits a
quasi-orthogonal decomposition into subalgebras isomorphic to
$M_n({\mathbb C})$. (As was mentioned, the same does {\it not} hold for $n=k=2$.)
The proof is constructional and relies on the existence of finite fields
and in some sense it is carried out in a similar manner to
the construction of $n+1$ MUB in dimension $n=p^\alpha$ (where $p>2$ is a
prime and $\alpha$ is a natural number).
\section{Decompositions of $M_n({\mathbb C})\otimes
M_n({\mathbb C}) \equiv M_{n^2}({\mathbb C})$}
\label{sec:M_n}
We shall now consider quasi-orthogonal
decompositions of $M_{n^2}({\mathbb C})$ into subfactors isomorphic to
$M_n({\mathbb C})$ and a number of MASAs.
Such decompositions of $M_2({\mathbb C})\otimes M_2({\mathbb C})\equiv
M_4({\mathbb C})$ are well studied in \cite{pszw}. However, there the achieved
results relay on explicit calculations carried out in $4$ dimensions. What
can we do in higher dimensions?
Note that the trivial necessary conditions do not rule out the
existence of a decomposition of the mentioned type;
all they say that such decompositions must consists of
\begin{equation}
\frac{n^4-1}{n^2-1} = n^2+1
\end{equation}
subalgebras (since both a maximal abelian subalgebra of $M_{n^2}({\mathbb C})$
and the factor $M_n({\mathbb C})$ is $n^2$-dimensional).
Decomposition into MASAs is of course interesting,
but it is known to be a hard question which is usualy studied in terms of
MUB and it is out of the scope of this article. Actually, there is a certain mathematical (or
more precisely: operator algebraic) advantage of having not only MASAs: it is often
helpful to consider the commutant of a subalgebra. (The commutant
of a MASA is itself, so it does not give anything ``new''.)
Infact, the result in \cite{petzkhan} concerning quasi-orthogonal
copies of $M_2({\mathbb C})$ in $M_4({\mathbb C})$ is achieved exactly by considering
commutants.
In \cite{ohnopetz} an important result is deduced about the quasi-orthogonality
of commutants. We shall
now recall this result (stating it in a sightly different form).
\begin{lemma}
Let ${\mathcal A}_1$ and ${\mathcal A}_2$ be quasi-orthogonal subalgebras of $M_n({\mathbb C})$. Then
the commutants ${\mathcal A}_1'$ and ${\mathcal A}_2'$ are quasi-orthogonal if and only if
${\rm dim}({\mathcal A}_1)\, {\rm dim}(A_2)=n^2$.
\end{lemma}
\begin{proof}
The observation established by calculation (\ref{product-basis})
shows that the required equality holds if and only if
the set $\{A_1A_2|\,A_1\in{\mathcal A}_1,A_2\in{\mathcal A}_2\}$ spans
$M_n({\mathbb C})$. Hence our lemma is a simple reformulation of one of the
claims of the original statement \cite[Prop.\! 2]{ohnopetz}
\end{proof}
\begin{corollary}\label{1factor}
There is no quasi-orthogonal decomposition of $M_{n^2}({\mathbb C})$ into maximal
abelian subalgebras and a (single) factor isomorphic to $M_n({\mathbb C})$.
\end{corollary}
\begin{proof}
Suppose the maximal abelian algebras ${\mathcal A}_1,\ldots,{\mathcal A}_{n^2}$ together
with the factor ${\mathcal B}$ form such a decomposition. Then, since both
${\rm dim}({\mathcal B})={\rm dim}(M_n({\mathbb C}))=n^2$
and also the dimension of a maximal abelian subalgebra of $M_{n^2}({\mathbb C})$
is $n^2$, by the previous lemma we have that ${\mathcal B}'$ is quasi-orthogonal
to ${\mathcal A}'_k={\mathcal A}_k\; (k=1,\ldots,n^2)$. But since ${\mathcal B}$ is a factor, ${\mathcal B}'$ is
also quasi-orthogonal to ${\mathcal B}$. Hence ${\mathcal B}'$ should be quasi-orthogonal to
each member of a quasi-orthogonal decomposition, implying that ${\mathcal B}'$
should be equal to ${\mathbb C} \mathbbm 1$ and in turn, that ${\mathcal B}={\mathcal B}''$ should be
the full matrix algebra $M_{n^2}({\mathbb C})$ (which is clearly a contradiction).
\end{proof}
\noindent
{Remark.} In \cite{lmsweiner} the author of the present work has shown that
if ${\mathcal A}_1,\ldots,{\mathcal A}_d$ is a system of $d$
MASAs in $M_d({\mathbb C})$, then any pair of elements in
the orthogonal subspace $({\mathcal A}_1+\ldots +{\mathcal A}_d)^\perp$ must commute. In particular,
if ${\mathcal A}_1,\ldots,{\mathcal A}_d,{\mathcal B}$ is a quasi-orthogonal system in $M_d({\mathbb C})$ where
${\mathcal A}_1,\ldots,{\mathcal A}_d$ are MASAs, then ${\mathcal B}$ must be a
commutative algebra. This is of course a much stronger affirmation than the
above corollary. However, that article uses a much longer proof
and the method presented here has the further advantage that
--- as we shall shortly see --- it can be applied to cases when the number of
MASAs is less than $d$. In any case, the aim of the cited
work was to study mutually unbiased bases
(and not quasi-orthogonal decompositions in general);
the nonexistence of the above discussed system was not
stated explicitly there.
\bigskip
\noindent
Now how about decompositions of $M_{n^2}({\mathbb C})$ into MASAs
and {\it two} factors isomorphic to $M_n({\mathbb C})$? Such
decompositions, in general, cannot be ruled out. Indeed, as was mentioned,
in \cite{pszw} the case of $n=2$ was treated and in particular an example was
given for such a decomposition. Moreover, it was shown
that there are no decompositions of $M_4({\mathbb C})$ into
MASAs and factors isomorphic to $M_2({\mathbb C})$ in which the number of
factors would be $1,3$ or $5$. For general $n>1$, we shall now prove that
there is no decompositions of $M_{n^2}({\mathbb C})$ into MASAs
and factors isomorphic to $M_n({\mathbb C})$ in which the number
of factors is $3$ (and we have already seen that nor it can be $1$). We
will need some preparatory steps.
\begin{lemma}
\label{A1A2}
Let ${\mathcal A}_1$ and ${\mathcal A}_2$ be quasi-orthogonal subalgebras of $M_n({\mathbb C})$ and
$A_1\in {\mathcal A}_1$ and $A_2\in {\mathcal A}_2$ two traceless operators. Then
$A_jA_k \in {\mathcal A}_1+{\mathcal A}_2$ if $j=k$ whereas if $j\neq k$ then
$A_jA_k$ is orthogonal to the subspace ${\mathcal A}_1+{\mathcal A}_2$.
\end{lemma}
\begin{proof}
Apart from trivial affirmations, all we have to check that
is that the ``cross-terms'' $A_jA_k$ (where $j\neq k$) are orthogonal
to the subspace ${\mathcal A}_1+{\mathcal A}_2$. If $X\in {\mathcal A}_1$, then
\begin{equation}
\langle X, A_1A_2\rangle =
{\rm Tr}(X^*A_1A_2)=
{\rm Tr}((A_1^* X)^*W_2)
=
\langle (A_1^* X), A_2\rangle = 0
\end{equation}
since $(A_1^* X)\in {\mathcal A}_1$ whereas $A_2$ is a traceless operator in ${\mathcal A}_2$
which is supposed to be quasi-orthogonal to ${\mathcal A}_1$. If $X\in {\mathcal A}_2$ then using the
invariance of trace under cyclic permutations we still get that
\begin{equation}
\langle X, A_1A_2\rangle =
{\rm Tr}(X^*A_1A_2)=
{\rm Tr}(A_1A_2 X^*)
=
\langle A_1^*, (A_2 X^*)\rangle = 0
\end{equation}
as the traceless element $A_1^*$ of ${\mathcal A}_1$ is orthogonal to any element of ${\mathcal A}_2$ (and in particular,
to $A_2X^*$). The rest (the orthogonality of the other cross-term: $A_2A_1$) follows by symmetry of the argument.
\end{proof}
\begin{lemma}
\label{BcapA=1}
Let ${\mathcal A}_1$ and ${\mathcal A}_2$ be quasi-orthogonal subalgebras of $M_n({\mathbb C})$ and
suppose that ${\mathcal B}$ is a third subalgebra of $M_n({\mathbb C})$ such that ${\mathcal B}\subset {\mathcal A}_1+{\mathcal A}_2$.
Then either ${\mathcal B}\subset {\mathcal A}_j$ for some $j=1,2$ or ${\mathcal B}\cap {\mathcal A}_1 = {\mathcal B}\cap {\mathcal A}_2 = {\mathbb C} \mathbbm 1$.
\end{lemma}
\begin{proof}
Suppose there exists a $B\in {\mathcal B}$ which is neither in ${\mathcal A}_1$ nor in ${\mathcal A}_2$. Then its traceless part
\begin{equation}
B_0 = B - \tau(B)\mathbbm 1 = B- ({\rm Tr}(B)/n)\mathbbm 1
\end{equation}
is still an element of ${\mathcal B}$ which is neither in ${\mathcal A}_1$ nor in ${\mathcal A}_2$, so
\begin{equation}
B_0=B_1 + B_2
\end{equation}
for some $B_j\in {\mathcal A}_j$ nonzero traceless operators $(j=1,2)$.
If $X\in {\mathcal B} \cap {\mathcal A}_1$ then again its traceless part $X_0$ is still in the intersection ${\mathcal B}\cap {\mathcal A}_1$.
Thus $X_0 B_1\in {\mathcal A}_1$ whereas
by the previous lemma $X_0B_2$ is orthogonal to ${\mathcal A}_1+{\mathcal A}_2$ since
$X_0\in {\mathcal A}_1$. On the other hand, as $X_0\in {\mathcal B}$, we have that
$X_0 B_1 + X_0 B_2 = X_0 B_0 \in {\mathcal B}\subset {\mathcal A}_1+ {\mathcal A}_2$. These two things imply that
$X_0 B_2 = 0$.
Of course the fact that $B_2\neq 0$ is not enough for showing that $X_0=0$.
However, the argument presented at equation (\ref{product-basis}) shows that --- apart from a
factor depending on the dimension $n$ --- the {\it trace-norm} of a product of two elements belonging to two quasi-orthogonal subalgebras is simply the product of norms. Hence in our case $X_0 B_2 = 0$
actually {\it does} imply that one of the terms in the product must be zero, and so that $X_0=0$.
Thus the arbitrary element $X$ of the intersection ${\mathcal B}\cap {\mathcal A}_1$ is a multiple of the identity;
that is ${\mathcal B}\cap {\mathcal A}_1 = {\mathbb C}\mathbbm 1$. The rest of the claim follows by repeating the argument with
${\mathcal A}_1$ and ${\mathcal A}_2$ exchanged.
\end{proof}
\begin{lemma}
\label{anti-comm}
Let ${\mathcal A}_1$ and ${\mathcal A}_2$ be quasi-orthogonal subalgebras of $M_n({\mathbb C})$ and
suppose that $A_j\in {\mathcal A}_j$ are traceless operators ($j=1,2$)
such that $A^2\in {\mathcal A}_1+{\mathcal A}_2$ where $A=A_1+A_2$. Then $A_1$ and $A_2$ must
anti-commute.
\end{lemma}
\begin{proof}
The claim is evident because by the previous lemma, in the expansion
$A^2= A_1^2+A_2^2 + A_1A_2+A_2 A_1$,
the first two terms are in ${\mathcal A}_1+{\mathcal A}_2$ whereas the last two terms are orthogonal to this subspace.
\end{proof}
\begin{theorem}
Let ${\mathcal A}_1$ and ${\mathcal A}_2$ be quasi-orthogonal subalgebras of $M_n({\mathbb C})$ and
suppose that ${\mathcal B}$ is a third subalgebra of $M_n({\mathbb C})$ such that ${\mathcal B}\subset {\mathcal A}_1+{\mathcal A}_2$.
Then either ${\mathcal B}\subset {\mathcal A}_j$ for some $j=1,2$ or ${\mathcal B}\simeq {\mathbb C}^2$.
\end{theorem}
\begin{proof}
Assume by contradiction that ${\mathcal B}$ is not included in neither of the two given quasi-orthogonal subalgebras, but ${\mathcal B}$ is not isomorphic to ${\mathbb C}^2$. Then ${\rm dim}(B)>2$ (since up to isomorphism,
there is only one two dimensional star-algebra: ${\mathbb C}^2$) and so the
traceless part of ${\mathcal B}$,
\begin{equation}
{\mathcal B}_0:= {\mathcal B}\cap \{\mathbbm 1\}^\perp
\end{equation}
is at least 2-dimensional.
Let $E_j$ be the trace-preserving expectation onto ${\mathcal A}_j$ for $j=1,2$. For any traceless element
$X\in {\mathcal A}_1+{\mathcal A}_2$ we have that $X=E_1(X)+E_2(X)$, so ${\mathcal B}_0\subset E_1({\mathcal B}_0)+E_2({\mathcal B}_0)$.
Moreover, this inclusion cannot be an equality, since in that case ${\mathcal B}_0$ would nontrivially
intersect ${\mathcal A}_1$ or ${\mathcal A}_2$, contradicting to our previous lemma.
Thus at least one out of the subspaces: $E_1({\mathcal B}_0), E_2({\mathcal B}_0)$ must be more than
1-dimensional; we may assume that ${\rm dim}(E_2({\mathcal B}_0))>1$.
Let $B\in {\mathcal B}$ be a traceless self-adjoint element. Then $B=B_1+B_2$ where $B_j=E_j(B)$ $(j=1,2)$,
and since ${\rm dim}(E_2({\mathcal B}_0))>1$, there exists a $\tilde{B}\in {\mathcal B}_0$ such that
$\tilde{B}_2=E_2(\tilde{B})$ is nonzero and orthogonal to $B_1$. As ${\mathcal B}$ is an algebra, we have that
$B\tilde{B}\in {\mathcal B}\subset {\mathcal A}_1+{\mathcal A}_2$. But $B\tilde{B} = B_1\tilde{B}_1+ B_2\tilde{B}_2 +
B_1\tilde{B_2}+ B_2\tilde{B}_1$, and according to lemma \ref{A1A2}, the first two
terms in this sum are in ${\mathcal A}_1+{\mathcal A}_2$ whereas the last two terms are orthogonal to this subspace, so
actually $B_1\tilde{B_2}+ B_2\tilde{B}_1 = 0$. On the other hand, by using the product-property
(\ref{product-trace}), the anti-commutativity of $B_1$ and $B_2$ (assured by lemma \ref{anti-comm}),
and the fact that $B_2= E_2(B)$ is self-adjoint (as so is $B$), we have that
\begin{eqnarray}
\nonumber
\langle B_1\tilde{B_2}, B_2\tilde{B}_1 \rangle &=&
{\rm Tr}((B_1\tilde{B_2})^*B_2\tilde{B}_1) =
{\rm Tr}(\tilde{B_2}^*B_1 B_2\tilde{B}_1)
= - {\rm Tr}(\tilde{B_2}^*B_2 B_1\tilde{B}_1 )
\\
&=& n \tau(\tilde{B_2}^*B_2 B_1\tilde{B}_1 )
=
n \tau(\tilde{B_2}^*B_2) \tau(B_1\tilde{B}_1) = 0,
\end{eqnarray}
since by assumption $0=\langle \tilde{B_2}, B_2\rangle= {\rm Tr}(\tilde{B_2}^*B_2)=
n \tau(\tilde{B_2}^*B_2)$. Thus $B_1\tilde{B_2}=B_2\tilde{B}_1=0$ since they are orthogonal but
their sum is zero. As $\tilde{B}_2\neq 0$, this implies (by the argument already explained towards the end of the proof of lemma \ref{BcapA=1}) that ${\mathcal B}_1=0$. That is, $B=B_2$ is actually an element of ${\mathcal A}_2$. But our assumption, together with lemma \ref{BcapA=1} imply that
${\mathcal B}\cap {\mathcal A}_2 = {\mathbb C}\mathbbm 1$. It should then further follow that $B=0$; that is, we have shown that any self-adjoint traceless element in ${\mathcal B}$ is zero and hence that ${\mathcal B}={\mathbb C}\mathbbm 1$ which contradicts to the assumption that ${\mathcal B}$ is {\it not} a subalgebra of ${\mathcal A}_1$ or ${\mathcal A}_2$.
\end{proof}
\begin{theorem}
There are no quasi-orthogonal decompositions of $M_{n^2}({\mathbb C})$ into
maximal abelian subalgebras and factors isomorphic to $M_n({\mathbb C})$ in which
the number of factors would be $1$ or $3$.
\end{theorem}
\begin{proof}
We have already proved the case in which the number of factors is $1$, so now assume by
contradiction that we have a quasi-orthogonal decomposition containing three factors: ${\mathcal B}_1,{\mathcal B}_2,{\mathcal B}_3$
(all isomorphic to $M_n({\mathbb C})$) and $n^2-2$ MASAs ${\mathcal A}_1,\ldots, {\mathcal A}_{n^2-2}$. Then, as was already noted and applied, considering the commutants: ${\mathcal B}_1',{\mathcal B}_2',{\mathcal B}_3',{\mathcal A}_1,\ldots,
{\mathcal A}_{n^2-2}$
(where we have used that the commutant of a MASA is itself), we
still have a quasi-orthogonal decomposition. Thus, ${\mathcal B}_1'$ is quasi-orthogonal to both ${\mathcal B}_1$ (since it is a factor) and the algebras ${\mathcal A}_1,\ldots, {\mathcal A}_{n^2-2}$ and hence ${\mathcal B}_1\subset
{\mathcal B}_2+{\mathcal B}_3$. By our previous theorem it then follows that ${\mathcal B}_1'$ is either equal to ${\mathcal B}_2$ or to
${\mathcal B}_3$. Repeating our argument for ${\mathcal B}_2$ and ${\mathcal B}_3$, we see that the ${\mathcal B}'_j={\mathcal B}_{\sigma(j)}$
for some $\sigma:\{1,2,3\}\to\{1,2,3\}$ such that:
\begin{itemize}
\item $\sigma^2= {\rm id}$ (since the second commutant gives back the original algebra),
\item $\sigma$ has no fixed points (${\mathcal B}'_j\neq {\mathcal B}_j$ as ${\mathcal B}_j$ is not a MASA).
\end{itemize}
However, these two properties are evidently contradicting.
\end{proof}
\section{The trace formula}
\label{sec:formula}
Suppose $P$ and $Q$ are the ortho-projections onto the subspaces $N$ and $K$,
respectively. By elementary arguments involving traces and positive
operators, one has that ${\rm Tr}(PQ)$ is a nonnegative real number,
\begin{equation}
{\rm dim}(N\cap K) \leq {\rm Tr}(PQ)\leq {\rm min}\{{\rm dim}(N), \, {\rm dim}(K)\},
\end{equation}
and moreover that ${\rm dim}(N\cap K) = {\rm Tr}(PQ)$ if and only if
$N\cap (N\cap K)^\perp$ and $K\cap (N\cap K)^\perp$ are orthogonal.
Thus we may say that the nonnegative number ${\rm Tr}(PQ)-{\rm dim}(N\cap K)$
measures ``how much'' the subspaces $N\cap (N\cap K)^\perp$ and
$K\cap (N\cap K)^\perp$ are {\it not} orthogonal. This number is zero if and only
if they are orthogonal, and in some sense the bigger it is, the further away they
are from orthogonality. Let us see now what this has to do with quasi-orthogonal subalgebras.
A subalgebra
${\mathcal A}\subset M_n({\mathbb C})$ is in particular a linear subspace. As was
discussed, $M_n({\mathbb C})$ has a natural scalar product, so it is meaningful to talk
about orthogonality. Thus we may consider the ortho-projection $E_{\mathcal A}$ onto ${\mathcal A}$.
Note that this map is usually referred as the {\it unique trace-preserving
expectation} onto ${\mathcal A}$. For two subalgebras ${\mathcal A},{\mathcal B} \subset M_n({\mathbb C})$ we shall now
introduce the quantity
\begin{equation}
c({\mathcal A},{\mathcal B}):={\rm Tr}(E_{\mathcal A} E_{\mathcal B}).
\end{equation}
Note that here ${\rm Tr}$ is the trace of the set of linear operators {\it acting}
on $M_n({\mathbb C})$, and not the trace of $M_n({\mathbb C})$.
Recall that by ``subalgebra'' we always mean a $^*$-subalgebra containing the
identity, so ${\mathcal A} \cap {\mathcal B}$ is at least one-dimensional. Thus
$c({\mathcal A},{\mathcal B})$ is a nonnegative real and infact
\begin{equation}
1\leq c({\mathcal A},{\mathcal B}) \leq {\rm min}\{{\rm dim}({\mathcal A}), {\rm dim}({\mathcal B})\}
\end{equation}
with $c({\mathcal A},{\mathcal B})=1$ if and only if ${\mathcal A}$ and ${\mathcal B}$ are quasi-orthogonal.
We are interested by the relation between the quantities
$c({\mathcal A},{\mathcal B})$ and $c({\mathcal A}',{\mathcal B}')$ where ${\mathcal A}'$ and ${\mathcal B}'$ are the commutants of ${\mathcal A}$ and
${\mathcal B}$, respectively. The next example shows that in general, $c({\mathcal A}',{\mathcal B}')$ cannot
be determined by the value of $c({\mathcal A},{\mathcal B})$ and the (unitary) equivalence
classes\footnote{Two isomorphic subalgebras of $M_n({\mathbb C})$ (that is:
$^*$-subalgebras containing $\mathbbm 1\in M_n({\mathbb C})$) are are not
necessarily unitarily equivalent. For example, if $P$ and $Q$ are
ortho-projections in $M_4({\mathbb C})$ onto subspaces of dimensions $2$ and $3$,
respectively, then ${\mathcal A} \simeq {\mathcal B} \simeq {\mathbb C}^2$ where ${\mathcal A}={\mathbb C} P + {\mathbb C} \mathbbm 1$
and ${\mathcal B}={\mathbb C} Q + {\mathbb C} \mathbbm 1$. However, clearly there is no unitary $U\in
M_n({\mathbb C})$ such that $U{\mathcal A} U^*$ would coincide with ${\mathcal B}$.} of the subalgebras ${\mathcal A}$ and ${\mathcal B}$.
\bigskip
\noindent
{\it Example.} Let ${\mathcal A}:=M_4({\mathbb C})\otimes \mathbbm 1_4 \subset M_4({\mathbb C})\otimes
M_4({\mathbb C})\equiv M_{16}({\mathbb C})$ and $\tilde{{\mathcal A}}:={\mathcal A}'=\mathbbm 1_4 \otimes M_4({\mathbb C})$.
Then clearly, ${\mathcal A}$ and $\tilde{{\mathcal A}}$ are unitarily equivalent.
Let further $P_1\in M_4({\mathbb C})$ be an ortho-projection onto a
one-dimensional subspace and $P_2\in M_4({\mathbb C})$ an ortho-projection onto a
two-dimensional subspace. Finally, let ${\mathcal B}:={\mathcal D}_1 \otimes {\mathcal D}_2$ where
${\mathcal D}_1,{\mathcal D}_2\subset M_4({\mathbb C})$ are the abelian subalgebras generated by the single
ortho-projections $P_1$ and $P_2$, respectivly. Then, as all of the
algebras ${\mathcal A}, {\mathcal A}', \tilde{{\mathcal A}}, \tilde{{\mathcal A}}', {\mathcal B},{\mathcal B}'$ have a product-form, it
is easy to see that
\begin{equation}
E_{\mathcal A} E_{\mathcal B} = (id_4 \otimes {\rm Tr}_4) (E_{{\mathcal D}_1}\otimes
E_{{\mathcal D}_2}) = E_{{\mathcal D}_1}\otimes {\rm Tr}_4 = E_{{\mathcal D}_1\otimes \mathbbm 1_4},
\end{equation}
so
$c({\mathcal A},{\mathcal B}) = {\rm Tr}(E_{\mathcal A} E_{\mathcal B}) = {\rm Tr}(E_{{\mathcal D}_1\otimes \mathbbm 1_4}) =
{\rm dim}({\mathcal D}_1) = 2$
and similarly, that $c(\tilde{{\mathcal A}},{\mathcal B})$ is also equal to $2$. However, as
${\mathcal B}'={\mathcal D}_1'\otimes {\mathcal D}_2'$ while the commutants of ${\mathcal A}$ and $\tilde{{\mathcal A}}$
are $\tilde{{\mathcal A}}$ and ${\mathcal A}$, respectively, we have that
\begin{equation}
c({\mathcal A}',{\mathcal B}') = {\rm dim}({\mathcal D}_2') = 2^2 + 2^2 \neq 1^2 + 3^2 = {\rm dim}({\mathcal D}_1') =
c(\tilde{{\mathcal A}},{\mathcal B}').
\end{equation}
Thus the value of $c({\mathcal A},{\mathcal B})$, even together with the knowledge of the unitary
equivalence classes of ${\mathcal A}$ and ${\mathcal B}$, is insufficient for determining $c({\mathcal A}',{\mathcal B}')$.
\bigskip
\noindent
A subalgerba, up to unitary equivalence, is always of the form
\begin{equation}
{\mathcal A}= \oplus_{k} \left(
M_{n_k}({\mathbb C}) \otimes \mathbbm 1_{m_k}
\right) \, \subset M_n({\mathbb C})
\end{equation}
where $n=\sum_k n_k m_k$
(and $\mathbbm 1_x$ is the unit of $M_x({\mathbb C})$)
with commutant
\begin{equation}
{\mathcal A}'= \oplus_{k} \left(\mathbbm 1_{n_k}({\mathbb C}) \otimes M_{m_k}({\mathbb C})
\right)\, \subset M_n({\mathbb C}).
\end{equation}
In case the ratios $n_k/m_k$ are independent of the index $k$,
we shall say that the subalgebra ${\mathcal A}$ is {\bf homogeneously balanced}.
Note that if $n_k/m_k = \lambda$ for all indices $k$, then
$n= \sum_k n_k m_k = \lambda \sum_k n_k^2 =\lambda {\rm dim}({\mathcal A})$ and
${\rm dim}({\mathcal A}) = \sum_k n_k^2 = \lambda^2 \sum_k m_k^2 = \lambda^2 {\rm dim}({\mathcal A}')$.
Some evident, but important consequences
of our definition and this last remark are:
\begin{itemize}
\item
${\mathcal A}$ is homogeneously balanced if and only if so is ${\mathcal A}'$,
\item
if ${\mathcal A}\subset M_n({\mathbb C})$ and ${\mathcal B}\subset M_m({\mathbb C})$ are homogeneously balanced
then so is the tenzorial product
${\mathcal A}\otimes {\mathcal B} \subset M_n({\mathbb C})\otimes M_m({\mathbb C}) \equiv M_{nm}({\mathbb C})$,
\item
factors and MASAs are automatically homogeneously balanced,
\item
if ${\mathcal A}$ is homogeneously balanced then ${\rm dim}({\mathcal A}){\rm dim}({\mathcal A}') = n^2$
\item
this homogeneity, in general, is not only a condition about the isomorphism class of
${\mathcal A}$, but also a condition about the way it ``sits'' in $M_n({\mathbb C})$: of two
isomorphic subalgebras, one may be homogeneously balanced while the other is not.
\end{itemize}
Note that the algebra ${\mathcal B}$ in the previous example was {\it not} homogeneously balanced.
We shall now recall a simple, but important fact; for the more general statement and its proof
see \cite[Prop.\! 1]{ohnopetz}.
\begin{lemma}
If ${\mathcal A}\subset M_n({\mathbb C})$ is homogeneously balanced and
$A_1,\ldots A_N$ is an
ortho-normed basis of ${\mathcal A}$, then
$E_{\mathcal A}(X) = \frac{n}{N}\sum_{j=1}^N A_k X A_k^*$
for all $X\in M_n({\mathbb C})$.
\end{lemma}
\begin{theorem}
If ${\mathcal A},{\mathcal B}\subset M_n({\mathbb C})$ are homogeneously balanced then
$$
c({\mathcal A}',{\mathcal B}') = \frac{n^2}{{\rm dim}({\mathcal A}){\rm dim}({\mathcal B})}c({\mathcal A},{\mathcal B}).
$$
\end{theorem}
\begin{proof}
Let $A_1,\ldots A_N\in {\mathcal A}$ and $B'_1,\ldots B'_{\tilde{N}}\in {\mathcal B}'$ be two
ortho-normed bases, where $N:={\rm dim}({\mathcal A})$ and
$ \tilde{N}:={\rm dim}({\mathcal B}')= n^2/{\rm dim}({\mathcal B})$. Using the previous lemma
\begin{eqnarray}
\nonumber
\sum_{j,k}{\rm Tr}(A_j B'_k A_j^* {B'_k}^*)
&=& \frac{n}{N}
\sum_k {\rm Tr}(E_{{\mathcal A}'}(B'_k){B'_k}^*)
= \frac{n}{N}\sum_k
{\rm Tr}(E_{{\mathcal A}'}(E_{{\mathcal A}'}(B'_k){B'_k}^*))
\\
\nonumber
&=& \frac{n}{N}\sum_k
{\rm Tr}(E_{{\mathcal A}'}(B'_k)E_{{\mathcal A}}({B'_k}^*))
= \frac{n}{N}\sum_k \| E_{{\mathcal A}'}(B'_k) \|_{{\rm Tr}}^2
\\
\label{eq:ccomp}
&=&\frac{n}{{\rm dim}({\mathcal A})} Tr (E_{{\mathcal A}'}E_{{\mathcal B}'})
\end{eqnarray}
where we have used the simple fact that if $P,Q$ are ortho-projections then
${\rm Tr}(PQ)= \sum_k \|Pq_k\|^2$ where $q_1,\ldots, q_s$ is an ortho-normed basis
in the image of $Q$. However, we could have carried out the above
calculation in a similar way but with the role of ${\mathcal A}$ and ${\mathcal B}'$ exchanged.
Confronting the obtained form to the one appearing in the previous equation one can
easily obtain the claimed formula.
\end{proof}
\section{Example applications of the formula}
\label{sec:application}
The so-far presented results relied on the result of Petz and Ohno which ensured
that if ${\mathcal A},{\mathcal B}\subset M_n({\mathbb C})$ are quasi-orthogonal and
${\rm dim}({\mathcal A}){\rm dim}({\mathcal B}) = n^2$ then also ${\mathcal A}'$ and ${\mathcal B}'$ form a quasi-orthogonal pair.
Note that if ${\mathcal A}$ and ${\mathcal B}$ are homogeneously balanced, then this fact is a simple consequences of our formula obtained in the last section. In some sense, our formula gives a {\it quantitative generalization} of this fact.
To make use of this quantitative information, all we need is the following observation.
\begin{lemma}
\label{appliformula}
Let ${\mathcal C}\subset M_n({\mathbb C})$ be a subalgebra and ${\mathcal A}_1,\ldots,{\mathcal A}_k\subset M_n({\mathbb C})$ be a system of quasi-orthogonal subalgebras. Then
\begin{itemize}
\item[(i)]
${\rm dim}({\mathcal C})\geq 1-k + \sum_{j=1}^k c({\mathcal A}_j,{\mathcal C})$, and
\item[(ii)]
${\rm dim}({\mathcal C}) \leq n^2-1+ \sum_{j=1}^k (c({\mathcal A}_j,{\mathcal C})-{\rm dim}({\mathcal A}_j))$
\end{itemize}
with equality holding in (i) if and only if ${\mathcal C}\subset {\mathcal A}_1+{\mathcal A}_2+\ldots + {\mathcal A}_k$ (which is automatically satisfied if in particular
${\mathcal A}_1,\ldots,{\mathcal A}_k$ is a quasi-orthogonal system of $M_n({\mathbb C})$).
\end{lemma}
\begin{proof}
The subalgebra ${\mathbb C}\mathbbm 1$ is contained in every subalgebra, so
$E_{{\mathcal A}_j}-E_{{\mathbb C}\mathbbm 1}$ is a projection; in fact it is the ortho-projection onto
the ``traceless part'' of ${\mathcal A}_j$. Quasi-orthogonality of ${\mathcal A}_1,\ldots {\mathcal A}_k$ is then equivalent to the fact
that the projections $(E_{{\mathcal A}_j}-E_{{\mathbb C}\mathbbm 1})$ are mutually orthogonal for $j=1,\ldots ,k$. Moreover, in this case
\begin{equation}
F:=E_{{\mathbb C}\mathbbm 1} + \sum_{j=1}^k (E_{{\mathcal A}_j}-E_{{\mathbb C}\mathbbm 1})
\end{equation}
is nothing else than the ortho-projection onto the subspace ${\mathcal A}_1+\ldots+{\mathcal A}_k$. Hence
\begin{eqnarray}
\label{eq:FE_C}
\nonumber
{\rm Tr}(FE_{{\mathcal C}}) &=& {\rm Tr}(E_{{\mathbb C}\mathbbm 1}E_{\mathcal C}) + \sum_{j=1}^k
({\rm Tr}(E_{{\mathcal A}_j}E_{{\mathcal C}}) - {\rm Tr}(E_{{\mathbb C}\mathbbm 1}E_{\mathcal C}) =
\\
& =&
1 + (\sum_{j=1}^k c({\mathcal A}_,{\mathcal C}) - 1) = 1-k + \sum_{j=1}^k c({\mathcal A}_j,{\mathcal C})
\end{eqnarray}
as $E_{{\mathbb C}\mathbbm 1}E_{\mathcal C} = E_{{\mathbb C}\mathbbm 1}$ is a projection onto
a one-dimensional subspace. Thus (i) follows as
\begin{equation}
{\rm Tr}(FE_{\mathcal C})\leq {\rm Tr}(E_{\mathcal C}) = {\rm dim}({\mathcal C})
\end{equation}
with equality holding if and only if $E_{\mathcal C}$ is a smaller projection than $F$; i.e.\! when
${\mathcal C}\subset ({\mathcal A}_1+\ldots + {\mathcal A}_k)$.
The inequality (ii) follows by considering $\mathbbm 1$
as the sum of the two orthogonal projections: $\mathbbm 1 = F + (\mathbbm 1-F)$. Now
$F$ is an ortho-projection onto a $d:={\rm dim}(\sum_{j}{\mathcal A}_j)$ dimensional space,
where by quasi-orthogonality
\begin{equation}
d = 1+\sum_j({\rm dim}(A_j)-1),
\end{equation}
whereas $(\mathbbm 1-F)$ is the ortho-projection onto the orthogonal of ${\mathcal A}_1+\ldots+{\mathcal A}_k$, which is an
$n^2-d$ dimensional subspace. Thus
\begin{equation}
{\rm dim}({\mathcal C})={\rm Tr}(E_{\mathcal C}) = {\rm Tr}(FE_{\mathcal C})+ {\rm Tr}((\mathbbm 1-F)E_{\mathcal C})
\end{equation}
where ${\rm Tr}((\mathbbm 1-F)E_{\mathcal C})\leq n^2-d =n^2 -1-\sum_j({\rm dim}(A_j)-1)$. This, together
with (\ref{eq:FE_C}) expressing the term ${\rm Tr}(FE_{\mathcal C})$, concludes our proof.
\end{proof}
So let us see now how we can use our formula in practice. We begin with a fairly
simple case; we shall
consider a quasi-orthogonal system in $M_6({\mathbb C})$ containing $6$ maximal abelian subalgebras
${\mathcal A}_1,\ldots {\mathcal A}_6$ and a subalgebra ${\mathcal B}$ isomorphic to $M_2({\mathbb C})$.
The trivial necessary conditions would allow the existence of such a system.
Of course, as was explained in the remark made after
Corollary \ref{1factor}, by using the strong result of \cite{lmsweiner}, it is easy to show that such a system cannot exists.
But how could we rule out its existence in a more direct manner?
Now we cannot use Corollary \ref{1factor}: $6$ is not a square number and more in
particular ${\rm dim}(M_2({\mathbb C}))= 2^2\neq 6$, so ${\mathcal B}'$ would not remain quasi-orthogonal to
the subalgebras ${\mathcal A}_1,\ldots, {\mathcal A}_6$.
So assume the existence of such a system. Then by the fact that ${\mathcal A}_j$ is a MASA $(j=1,\ldots ,6)$, and
by an application of our formula
\begin{equation}
c({\mathcal A}_j,{\mathcal B}') = c({\mathcal A}_j',{\mathcal B}') = \frac{6^2}{6*4}c({\mathcal A}_j,{\mathcal B}) = \frac{3}{2},
\end{equation}
since ${\rm dim}({\mathcal B})= {\rm dim}(M_2({\mathbb C})) =4$ and $c({\mathcal A}_j,{\mathcal B}) = 1$ by the assumed quasi-orthogonality
of ${\mathcal B}$ and ${\mathcal A}_j$. Moreover, $c({\mathcal B},{\mathcal B}')=1$ because ${\mathcal B}$ was assumed to be a factor.
So considering the quasi-orthogonal system ${\mathcal A}_1,\ldots,{\mathcal A}_6,{\mathcal B}$ and the algebra ${\mathcal C}:={\mathcal B}'\simeq M_3({\mathbb C})$, we have
${\rm dim}({\mathcal A}_j) = 6, \, {\rm dim}({\mathcal B}) = 2^2 = 4$, and
\begin{eqnarray}
\nonumber
n^2-1 + (c({\mathcal B},{\mathcal C})-{\rm dim}({\mathcal B}))
+ \sum_{j=1}^6 (c({\mathcal A}_j,{\mathcal C})-{\rm dim}(A_j))
&= &
\\
6^2-1 \; +\;\;\;\;\;\;\;\;(1-4)\; \;\;\;\;\;\;\;\;\; +\;\;\; \;\;\;\; 6 \;*\;\;(\frac{3}{2}-6)\;\;\; \;\;\;\;\;\;\;\;&=& 5,
\end{eqnarray}
which is in conflict with (ii) of lemma \ref{appliformula}, as
${\rm dim}({\mathcal C})={\rm dim}({\mathcal B}')=3^2 =9\nleq 5$.
This is nice, but --- as was mentioned --- it is a fairly simple case in which we have already known the nonexistence. So we shall finish by considering a somewhat more complicated example.
\begin{proposition}
There is no quasi-orthogonal system in $M_6({\mathbb C})$ consisting of $5$ maximal abelian subalgebras and $3$ factors isomorphic to $M_2({\mathbb C})$.
\end{proposition}
\begin{proof}
Again, note that the existence of such a system cannot be ruled out by the trivial necessary conditions.
We assume ${\mathcal A}_1,\ldots,{\mathcal A}_5,{\mathcal B}_1,{\mathcal B}_2{\mathcal B}_3$ is such a system (with the ``${\mathcal A}$'' algebras being the maximal
abelian ones, and the ``${\mathcal B}$'' algebras the factors isomorphic to $M_2({\mathbb C})$).
To apply our formula, we will need to consider the commutants as well as the original algebras.
If ${\mathcal B}_1,{\mathcal B}_2,{\mathcal B}_3\simeq M_2({\mathbb C})$ and ${\mathcal B}_1,{\mathcal B}_2,{\mathcal B}_3\subset M_6({\mathbb C})$, then
their commutants ${\mathcal B}'_1, {\mathcal B}'_2, {\mathcal B}'_3 \simeq M_3({\mathbb C})$. Since $M_2({\mathbb C})$ cannot be embedded in $M_3({\mathbb C})$ in an identity preserving way,
we have that ${\mathcal B}_j$ is not contained in ${\mathcal B}'_k$ and consequently that
\begin{equation}
c({\mathcal B}_j,{\mathcal B}'_k) < {\rm dim}({\mathcal B}_j) = 4
\end{equation}
for every $j,k=1,2,3$. However, we shall need a better estimate. The fact is that ${\mathcal B}_j$ is not only not contained in ${\mathcal B}'_k$, but actually
we can say something about their ``minimal distance''. We shall shortly interrupt our proof with a lemma concerning this issue.
\begin{lemma}
$c({\mathcal B}_j,{\mathcal B}'_k) \leq 3$.
\end{lemma}
\noindent
{\it Proof (of lemma).}
Let $X,Y,Z, W\in {\mathcal B}_j$ be an orthogonal basis such that $W=\mathbbm 1$ and
$X,Y,Z$ correspond to the Pauli-matrices in a suitable identification ${\mathcal B}_j\simeq M_2({\mathbb C})$.
Let us further denote the trace-preserving expectation onto ${\mathcal B}'_k$ by $E$. Then
$E(X)$ (and similarly $E(Y)$ and $E(Z)$, too) remains self-adjoint,
so it is unitarily equivalent with a diagonal matrix. Moreover,
as it belongs to ${\mathcal B}'_k\simeq M_3({\mathbb C})$, we may actually assume it
is unitarily equivalent with the diagonal matrix
${\rm diag}(\lambda_1,\lambda_1,\lambda_2,\lambda_2,\lambda_3,\lambda_3)\in M_6({\mathbb C})$.
We have that
\begin{itemize}
\item $\lambda_1+ \lambda_2 + \lambda_3 = 0$,
\item $\lambda_1,\lambda_2,\lambda_3 \in [-1,1]$.
\end{itemize}
Indeed, the first equation follows as ${\rm Tr}(E(X))= {\rm Tr}(X)=0$, whereas the second follows from the fact
$\mathbbm 1 \pm E(X) = E(\mathbbm 1 \pm X)$ --- just as $\mathbbm 1 \pm X$ --- is a positive operator.
Now elementary calculus shows that in the region determined by the two equation, we have
\begin{equation}
{\rm Tr}(E(X)^2) = 2(\lambda_1^2 + \lambda_2^2 + \lambda_3^2) \leq 4.
\end{equation}
On the other hand, ${\rm Tr}(X^2) = {\rm Tr}(\mathbbm 1)=6$; actually,
$X,Y,Z,\mathbbm 1$ is an orthogonal basis whose each member has
(trace)norm-square equal to $6$. Thus, using the arguments explained at
and after equation (\ref{eq:ccomp}), we have that
\begin{equation}
c({\mathcal B}_j,{\mathcal B}'_k) =
\frac{1}{6}({\rm Tr}(E(X)^2+E(Y)^2+E(Z)^2+E(\mathbbm 1)^2))
\leq \frac{1}{6}(4 + 4+ 4+ 6) = 3
\end{equation}
which is just what we wanted to prove.
\quad
\smallskip
To finish the proof, we consider the algebra ${\mathcal C}:={\mathcal B}'_1\simeq M_3({\mathbb C})$ and the quasi-orthogonal system
${\mathcal A}_1,\ldots , {\mathcal A}_5,{\mathcal B}_1,{\mathcal B}_2,{\mathcal B}_3$. As ${\mathcal A}'_k={\mathcal A}_k$. By an application of our formula have that
\begin{equation}
c({\mathcal A}_k,{\mathcal C}) = c({\mathcal A}_k,{\mathcal B}'_1) = \frac{6^2}{6*2^2 } c({\mathcal A}_k,{\mathcal B}_1) = \frac{6^2}{2^2*6} = \frac{3}{2}
\end{equation}
since $c({\mathcal A}_k,{\mathcal B}_1)=1$ by quasi-orthogonality. Now
$c({\mathcal B}_1,{\mathcal C}) = c({\mathcal B}_1,{\mathcal B}'_1)=1$ since
${\mathcal B}$ is a factor, and finally, for $c({\mathcal B}_2,{\mathcal C})$ and $c({\mathcal B}_3,{\mathcal C})$ we can use the estimate provided
by the lemma we have just made.
To sum it up: we have $n^2-1=6^2-1=35$,
\begin{eqnarray}
\nonumber
\sum_{j=1}^5(c({\mathcal A}_j,{\mathcal C}) - {\rm dim}({\mathcal A}_j)) &= & 5 * (\frac{3}{2}-6) = -\frac{45}{2}, \;\;\;\; {\rm and}
\\
\sum_{j=1}^3(c({\mathcal B}_j,{\mathcal C}) - {\rm dim}({\mathcal B}_j)) &\leq& (1-4) + (3-4) +(3-4) = - 5
\end{eqnarray}
which gives $35-(45/2)-5 = 15/2 \ngeq 9 ={\rm dim}({\mathcal C})$, in contradiction with point (ii) of lemma
\ref{appliformula}.
\end{proof}
|
\section{Introduction}
In Paper I of this series (Machalski, Koziel-Wierzbowska \& Jamrozy, 2007) a problem
of the cosmological evolution of the intergalactic medium (IGM) was recalled and a
necessity to find distant (z$>$0.5) "giant"-sized radio sources (hereafter referred to
as GRSs) with a very low energy densities in their extended lobes, to solve this problem,
was emphasised.
Extended large-sized double radio sources are not easy to recognise because of their
relatively low radio brightness and a difficulty to detect eventual bridge connecting
brighter parts (lobes) of a common radio structure. Several observational efforts show that
most of known GRSs lie at low redshifts of z$<$0.25. For a long time this caused
a presumption that such extragalactic double radio sources, especially those of FRII-type,
did not exist at redshifts higher than about 1 because of the expected strong evolution
of a uniform IGM, $\rho_{\rm IGM}\propto (1+z)^{3}$, confining the lobes of sources (e.g.
Kapahi 1989).
The situation changed over 10 years ago when Cotter, Rawlings \& Saunders (1996) and
Cotter (1998) presented an unbiased sample of giant radio sources selected from the 7C
survey (McGilchrist et al. 1990). Their sample comprised 12 large-size sources with
0.3$<$z$<$0.9; four of them having D$>$1 Mpc have been included in Table 1 of Paper I,
where there was shown that a list of known GRSs with z$>$0.5 and D$>$1 Mpc is very short.
The undertaken search for such GRSs on the southern sky hemisphere with the 11m SALT
telescope during the Performance Verification (P--V) phase has resulted in the detection
of 21 GRSs with the projected linear size greater than 1 Mpc.
However, we found that their redshifts do not exceed the value of 0.4 and the energy
density in only two of them is less than $10^{-14}$ J/m$^{3}$. We serendipitously
discovered that one of them (J1420$-$0545) is, in fact, the largest GRS in the Universe
(cf. Machalski et al. 2008).
Dynamical considerations relating to propagation of a bow shock at the head of powerful
supersonic jet through the galactic and/or intergalactic medium (e.g.
Arnaud et al. 1984; Begelman \& Cioffi 1989; Falle 1991) predict that the lengthening of
a channel along which the jet flows is governed by the balance between the jet's thrust
and the ram pressure of the ambient medium. This balance implies that if the channel
elongates at a rate faster than the rate at which the jet delivers energy, the motion of
the jet's head would slow down until it become subsonic. This should cause a drastic
reduction in both the speed and radiation of the head, thus effectively limiting further
growth of the radio structure both in size and luminosity. Such a scenario is incorporated
in the analytical model for the evolution of FRII-type radio sources of Kaiser,
Dennett-Thorpe \& Alexander (1997; hereafter referred to as KDA) which combines the dynamical
model of Kaiser \& Alexander (1997) with the model for expected radio emission under the
influence of energy loss processes.
The dynamical evolution of a FRII radio source strongly depends on characteristics of the
ambient medium. Gopal-Krishna \& Wiita (1987) proposed the two-medium model consisting of
an X-ray halo around the parent galaxy with gas density decreasing with radial distance
from the galaxy and a much hotter intergalactic medium (IGM) with constant density.
These two media were conceived to be pressure-matched at their interface.
It is worth noting that their model (hereafter referred to as G-KW model)
does not account for the warm/hot intergalactic medium (WHIM) frequently
surrounding massive galaxies, as it was constructed before the first
assessment of the WHIM by Cen \& Ostriker (1999). The original G-KW
model
allows a prediction of the limiting (maximum) values for the source's age and
linear size
depending on the environment conditions, the jet power, and the cosmic epoch characterised
by the source's redshift. However, our recent detections of very large-sized radio sources
with z$>$1 and exceeding the limits predicted by their model suggests that some of its free parameters should be modified.
In this paper an observational constraint for the G-KW model is analysed. For this purpose, an
effort in determining the largest sizes and dynamical ages of FRII-type radio sources at
redshifts
1$<$z$<$2 is undertaken. In Section 2, the
original G-KW model is briefly described and modified adopting modern (contemporary) values for
thermodynamic temperature and gas density of the two media. Then, the predicted relations between
the sources' linear size and the age, as well as the expansion speed of the radio lobes'
heads
and the age are calculated. The observational data used to constrain the two-medium model are
presented in Section 3. The small sample of the most distant giant-sized radio sources (Table 1
in Paper I) is revised and supplemented with two other limited samples of FRII-type sources
comprising: (i) sources larger than 400 kpc within the redshift range 1$<$z$<$2, most of them
found in this paper, and (ii) selected 3CRR sources in majority smaller than 400 kpc at z$>$0.5
forming a comparison sample of "normal"-sized radio sources.
Physical parameters of the sample sources: the dynamical age, the jet power, the central
radio-core density and the IGM density, and others, are derived using the DYNAGE algorithm
(Machalski et al. 2007a). The application of this algorithm to the sample sources and the
resulting values of the sources' parameters are described in Section 4. A comparison of
the model predictions with the observational data is presented and discussed in Section 5.
For the purpose of calculating the linear size, volume and luminosity of the sample sources we
use a $\Lambda$CDM model with cosmological parameters $\Omega_{\rm m}$=0.27, $\Omega_{\rm \Lambda}$=0.73
and $H_{0}$=71 km\, s$^{-1}$Mpc$^{-1}$.
\section{The revised G-KW model}
\subsection{Base of the model}
The jet's dynamics is governed by a balance between its thrust, $\Pi_{\rm jet}\approx
Q_{\rm jet}/v_{\rm jet}$ and the ram pressure force of the IGM, $\rho_{\rm a}v_{\rm hs}^{2}A_{\rm hs}$,
where $Q_{\rm jet}$ is the jet power, $v_{\rm jet}$
is it's velocity, $\rho_{\rm a}$ is the ambient
medium gas density, $v_{\rm hs}$ is the speed of the jet's head
(hot spot) with which it advances into the ambient medium, and $A_{\rm hs}$ is the
cross-sectional area of the bow shock at the end of the jet. In the G-KW model, the jet
propagates into a two-component medium comprised of:
-- the gaseous halo with a power-law density profile $\rho_{\rm h}(d)=\rho_{0}\left[ 1+(d/a_{0})^{2}
\right]^{-\delta}$ bound to the parent optical galaxy, where $\rho_{0}$ and $a_{0}$ are the density
and the radius of the central radio core, respectively, and $\delta$=5/6. This distribution is assumed
to be invariant with redshift. It is also assumed that this halo has nearly uniform electron temperature
$(kT)_{\rm h}$[keV] (medium 1), and
-- the surrounding hotter IGM of uniform density, $\rho_{\rm IGM}$, with the temperature $(kT)_{\rm IGM}
(1+z)^{2}$[keV] (medium 2).
Similarily to Gopal-Krishna \& Wiita (1987) we have to assume characteristic values for the
density and temperature of the considered media. The values adopted hereafter for the two
components are based on the following data:
(1) The radio core radius, $a_{0}$=3 kpc is based on the fitted X-ray surface-brightness profile
of nine nearby, low-luminosity radio galaxies recently observed by Croston et al. (2008). This
value of the radius is derived from the observed angular radius of about 10 arcsec.
(2) The halos' gas temperature have been determined in a number of papers. A uniform temperature
$(kT)_{\rm h}$=0.7 keV was measured for a few nearby, X-ray luminous elliptical galaxies with the
{\sl Chandra Observatory} by Allen et al. (2006). Using {\sl XMM-Newton} and {\sl Chandra}
observations, the values from 1 to 5 keV with a median of about 2.1 keV was found by Belsole et al.
(2007) for the X-ray clusters surrounding 20 luminous 3CRR radio sources. For the low-luminosity
radio galaxies analysed by Croston et al. (2008), a median of the fitted temperatures is about
1.4 keV.
(3) The halos'gas (proton) density of (1 -- 2)$\times 10^{4}$ m$^{-3}$ is fitted to X-ray counts
by Belsole et al. (2007).
The interface between the X-ray halo and IGM is determined balancing the IGM pressure against the
pressure distribution in the halo. A non-relativistic gas in thermal equilibrium that has an electron
density $n_{\rm e}$[m$^{-3}$] and temperature $(kT)_{\rm e}$[keV] will have an electron pressure
$p_{\rm e}$=$n_{\rm e}\,(kT)_{\rm e}$[Pa]. Expressing electron density by the mass density,
$\rho$=$n\,\mu\,m_{\rm H}$, this balance will have place at the halo's radius $R_{\rm h}$ calculated
from
\begin{equation}
\frac{\rho_{0}}{\mu_{\rm h}m_{\rm H}}\left[1+(R_{\rm h}/a_{0})^{2}\right]^{-\delta}(kT)_{\rm h}=
\frac{\rho_{\rm IGM}}{\mu_{\rm IGM}m_{\rm H}}(kT)_{\rm IGM},
\end{equation}
\noindent
where $\mu$ and $m_{\rm H}$ are the mean molecular weight and the mass of hydrogen atom,
respectively. In this paper we assume $\mu_{\rm h}$=0.5 and $\mu_{\rm IGM}$=1.4. Besides,
for the halo (medium 1) we adopt $n_{p}$=$1.5\times 10^{4}$ m$^{-3}$, i.e. a mean proton
density of the values given by Belsole et al. (2007), which corresponds to $\rho_{0}$=$10^{-22.6}$
kg\,m$^{-3}$, and the temperature $(kT)_{\rm h}$=1.4 keV. For the IGM density we take 50\% of the
cosmic matter density, i.e. $\rho_{\rm IGM}$=$0.5\Omega_{\rm m}h^{2}\rho_{\rm clos}$=
$0.5\times 0.27\times 0.71^{2}\times (3\,H_{0}^{2})/(4\pi\,G)$, which gives $\rho_{\rm IGM}$=
$10^{-26.9}$ kg\,m$^{-3}$. For the IGM temperature we adopt $(kT)_{\rm IGM}$=25 keV. Substituting
the above values into Eq.\,(1) we find $R_{\rm h}$=642 kpc. This radius of X-ray halo is
compatible with the radii determined by Cassano et al. (2007) for 15 Abell cluster radio halos
with the mean of $\sim 560\pm$170 kpc. This is worth to notice that our radius of 642 kpc is much
larger than 171 kpc used by Gopal-Krishna \& Wiita. In an expanding and uniform IGM this radius
should evolve as $R_{\rm h}(z)=642(1+z)^{-5/(2\delta)}$ kpc, i.e. $642(1+z)^{-3}$ kpc for
$\delta$=5/6.
\vspace{2mm}
Fig.1 (a), (b), (c) present the basic characteristics of the two-media model: the mass density
$\rho(d)$, the electron temperature $kT(d)$, and the resulting electron gas pressure $p(d)$, as
functions of the distance from the host galaxy (compact radio core), respectively. Note that
the balance $p_{\rm h}(R_{\rm h})=p_{\rm IGM}$ at $d=R_{\rm h}$ corresponds to a rapid
transition between $\rho(R_{\rm h})$ and $\rho_{\rm IGM}$, as well as between $kT(R_{\rm h})$ and $kT_{\rm IGM}$, and causes an unphysical effect shown in Section 5.1. The dashed curves in Fig.1 (a) and (b) indicate desired smooth transitions between the relevant parameters which would introduce a better physical scenario into the model.
\begin{figure}[t]
\begin{center}
\includegraphics[width=13cm]{Kuligowska1.eps}
\FigCap{Properties of the two-media environment surrounding the center of radio galaxy.}
\end{center}
\end{figure}
\vspace{2mm}
\subsection{Predictions of the model}
For $d\leq R_{\rm h}$ it is assumed that the jet propagate (through the medium 1) with a
constant opening angle, $\theta$. Under this condition, the ram pressure balance results in the
following dependence for the jet length (the radio lobe size, $D_{\ell}$) on time (the lobe's age, $t$)
and the jet's head expansion velocity, v$_{\rm h}$, on $D_{\ell}$ or $t$:
\begin{equation}
D_{\ell}(t)=\left[(2-\delta)\,A\,t\right]^{\frac{1}{2-\delta}},
\end{equation}
\begin{equation}
{\rm v}_{\rm h}(D_{\ell})=A\,D_{\ell}^{(\delta -1)},
\end{equation}
\begin{equation}
{\rm v}_{\rm h}(t)=\left[(2-\delta)\,A^{\frac{1}{\delta -1}}t\right]^{\frac{\delta -1}{2-\delta}},
\end{equation}
\noindent
where
\[A\equiv\left(\frac{4\,c_{1}Q_{\rm jet}}{\pi\theta^{2}c\,\rho_{0}a_{0}^{2\delta}}\right)^{1/2}.\]
\noindent
Here $c_{1}$ is a constant with a value between 1.5 and 3.8 depending on the source's (lobe's)
geometry described by its axial ratio $R_{\rm T}$ (cf. Kaiser \& Alexander 1997), while
$c\approx$ v$_{\rm jet}$ is the speed of light. The jet's opening angle is also described by
$R_{\rm T}$, $\theta^{2}=c_{2}/(4\,R_{\rm T})$, where $c_{2}$ is a constant with a value between
3.6 to 4.1 depending on specific heats for the material in the jet and the lobe (cocoon),
(Eq. 17 in Kaiser \& Alexander 1997).
At $d=R_{\rm h}(z)$ the jet enters the hotter IGM (medium 2) at least an order of magnitude
less dense but pressure-matched, as shown in Fig.\,1. In order to analyse the jet's propagation
over this regime, Gopal-Krishna \& Wiita have considered two likely extreme scenarios for the
lobe's expansion:
\vspace{2mm}
--- {\sl Scenario A} where the jet opening angle, $\theta$, is conserved. Due to a rapid
decrease of the ambient density at the interface, $\rho_{\rm IGM}\ll \rho_{\rm h}(R_{\rm h})$,
a sufficient ram-pressure will be provided only if the jet's head velocity, v$_{\rm hs}$,
increases abruptly at $d$=$R_{\rm h}$ and then gradually approaches the v$\propto d^{-1}$ law
expected for a constant density medium. In this scenario:
\begin{equation}
D_{\ell}(t)=\left\{2\left(K(z)+a_{0}^{\delta}\,A\left[\frac{\rho_{0}}{\rho_{\rm IGM}(1+z)^{3}}
\right]^{1/2}t\right)\right\}^{1/2}\hspace{10mm}{\rm and}
\end{equation}
\begin{equation}
{\rm v}_{\rm h}(t)=a_{0}^{\delta}\,A\left[\frac{\rho_{0}}{\rho_{\rm IGM}(1+z)^{3}}\right]^{1/2}/D_{\ell}(t),
\end{equation}
\noindent
where
\[K(z)=\frac{1}{2}R_{\rm h}^{2}(z) - R_{\rm h}^{(2-\delta)}(z)\frac{a_{0}^{\delta}}{2-\delta}
\left[\frac{\rho_{0}}{\rho_{\rm IGM}(1+z)^{3}}\right]^{1/2}\]
\noindent
is a redshift-dependent constant providing that the time corresponding to $D_{\ell}$=$R_{\rm h}$
in Eqs.\,2 and 5 is the same.
--- {\sl Scenario B} where the jet's head velocity across the interface remains continuous
and therefore matched to the value given by Eq.\,(3) for $D_{\ell}$=$R_{\rm h}$. This can be achieved
only with an abrupt flaring of the jet's opening angle. Under this condition the model predicts:
\begin{equation}
D_{\ell}(t)=\left\{2\left(A\,R_{\rm h}^{\delta}(z)\,t-R_{\rm h}^{2}(z)\frac{\delta}{2(2-\delta)}
\right)\right\}^{1/2}\hspace{10mm}{\rm and},
\end{equation}
\begin{equation}
{\rm v}_{\rm h}(t)=A\,R_{\rm h}^{\delta}(z)/D_{\ell}(t).
\end{equation}
\noindent
The sound speed in the IGM is given by
\begin{equation}
s_{\rm IGM}(z)=\left[\frac{\Gamma(kT)_{\rm IGM}}{\mu\,m_{\rm H}}\right]^{1/2}(1+z).
\end{equation}
\noindent
Given the values $\Gamma$=5/3, $\mu$=1.4 and $(kT)_{\rm IGM}$=25 keV, the sound speed
limiting the source (lobe) axial expansion velocity is $s_{\rm IGM}\approx 0.0056\,c$
at $z\approx 0$, and $s_{\rm IGM}\approx 0.0169\,c$ at $z\approx 2$.
The time dependence of the source size $D_{\ell}(t)$ calculated from Eqs.\,2, 5 and 7, as well
as of the axial expansion velocity v$(t)$ calculated from Eqs.\,4, 6 and 8 for the three
cases: $Q_{\rm jet}$=$10^{37.5}$ W and $z$=0.5, $Q_{\rm jet}$=$10^{38.5}$ W and $z$=1.0,
and $Q_{\rm jet}$=$10^{39.5}$ W and $z$=2.0, are shown in Fig.~5 and Fig.~6, respectively.
To compare these predictions with the observations, we select three different samples of
FRII-type radio sources.
\section{Selection of the samples}
\subsection{Sample 1 (GRS sample)}
The revised sample of FRII-type radio sources with 0.5$<$z$<$1 and the projected linear
size larger than 1 Mpc, compiled from the literature, is presented in Table~1. Columns (1)--(5)
are self explanatory. Entries in columns (6)--(10) are the data used to determine the age
and other physical parameters of the sample sources in Section 4.
These are: $R_{\rm T}$ -- axial ratio of the source, $\phi$ -- assumed
orientation angle of the jet axis, and $P_{\nu}$ -- monochromatic luminosity
of the source at the given low, medium, and high frequency, respectively (cf.
Section 4.1). Some necessary references are given in the Notes.
\MakeTable{lllccrccccl}{12.5cm}{The sample of FRII-type sources with 0.5$<$z$<$1 and D$>$1 Mpc.
The radio luminosities are given in units of W/(Hz$\cdot$sr). The low-frequency luminosity, $P_{lf}$,
is determined either at 74 MHz (marked {\sl a}) or 151 MHz ({\sl b}) or 325 MHz ({\sl c}).}
{\hline
IAU name & Survey & z & Id. &$D$&$R_{\rm T}$&$\phi$ & log &log &log &Notes\\
& & & & [kpc] & &[$^{o}$]& $P_{lf}$ & $P_{1400}$ & $P_{4850}$ \\
(1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9) & (10) & (11)\\
\hline
J0037$+$0027 & & 0.5908& G & 1976 & 5.0 & 90 & 26.103$^{a}$ & 25.149 & 24.606 & (4, 5)\\
B0654$+$482 & 7C & 0.776 & G & 1002 & 4.0 & 90 & 26.120$^{b}$ & 25.262 & 24.733 & (1)\\
J0750$+$656 & & 0.747 & Q & 1606 & 4.0 & 90 & 26.189$^{a}$ & 25.255 & 24.715 & (3)\\
B0821$+$695 & 8C & 0.538 & G & 2580 & 3.0 & 90 & 26.142$^{b}$ & 25.106 & 24.525 & (2)\\
B0854$+$399 & B2 & 0.528 & G & 1014 & 2.7 & 70 & 26.385$^{b}$ & 25.627 & 25.098 &\\
B1058$+$368 & 7C & 0.750 & G & 1100 & 4.2 & 90 & 26.192$^{b}$ & 25.412 & 24.809 & (1)\\
J1130$-$1320& PKS& 0.6337& Q & 2033 & 3.8 & 90 & 27.275$^{a}$ & 26.198 & 25.613 & (4, 5)\\
B1602$+$376 & 7C & 0.814 & G & 1376 & 2.6 & 90 & 26.292$^{b}$ & 25.486 & 24.936 & (1)\\
B1636$+$418 & 7C & 0.867 & G & 1004 & 3.9 & 90 & 26.006$^{b}$ & 25.273 & 24.701 & (1)\\
B1834$+$620 & WNB & 0.5194& G & 1384 & 4.1 & 90 & 26.361$^{b}$ & 25.555 & 25.067 & (6)\\
J1951$+$706 & & 0.550 & G & 1300 & 4.5 & 90 & 25.974$^{a}$ & 24.894 & 24.448 & (3)\\
J2234$-$0224& & 0.55 & Q & 1266 & 4.5 & 90 & --- & 24.840 & --- & (4, 5)\\
\hline
\multicolumn{11}{p{12.5cm}}{
Notes: (1)--Cotter et al. (1996); (2)--Lara et al. (2000); (3)--Lara et al. (2001);
(4)--Koziel-Wierzbowska (2008); (5)--Machalski et al. (2007b); (6)-- Schoenmakers et al. (2000)}
}
\subsection{Sample 2 (Distant-source sample)}
The second sample of FRII-type sources larger than 400 kpc and having z$>$1 consists of
a few sources known from the literature and the sources found in this paper. The latter
part of this sample results from the dedicated research project attempting to determine
how large linear size FRII-type radio sources can achieve at redshifts 1$<$z$<$2.
This part was preselected using the modern Sloan Digital
Sky Survey (hereafter referred to as SDSS: Adelman-McCarthy et al. 2007) as the finding survey.
The optical objects extracted from the SDSS, fulfilling the above redshift criterion and
classified either as a galaxy (G) or a QSO (Q), were then cross-correlated with the radio
1400-MHz sky survey FIRST (Becker, White \& Helfand 1995). In the second step, all optical
objects strictly coinciding (within an angular separation less than 0.5 arcsec) with a compact
radio source, i.e. with a potential radio core, were subject to the further selection. In the
third step, we have checked whether the (core) component is surrounded by a pair of nearly
symmetric, possibly extended radio structures (lobes) with an angular separation providing
a projected linear size $D^{>}_{\sim}$400 kpc. All candidates were verified by an inspection
of their images (if exist) in other available radio surveys, namely the 1400-MHz NVSS
(Condon et al. 1998), 325-MHz WENSS (Rengelink et al. 1997), and 74-MHz VLSS (Cohen et al.
2007). The assumed cosmology (cf. Introduction) predicts that a maximum of the linear size/angular
size quotient at z$>$1 is $\sim$8.55 kpc/arcsec which, in turn, implies that a source larger
than 400 kpc should have an angular size $LAS^{>}_{\sim}$47 arcsec. All radio
sources selected this way from the
SDSS are classified as QSO. The final sample is presented in Table~2. All columns give the
similar data as those in Table~1.
The radio images of exemplary sample sources, made using combined data from the NVSS and
FIRST surveys, are shown in Figs.\,2--4. The precise coordinates, angular sizes and 1400 MHz total flux
densities, taken from NVSS cataloque, are given in the Appendix (Table~8).
However, in order to use this sample to constrain the analytical G-KW model, besides the redshift
and linear size of the member sources, we have to determine their age and other physical
parameters which is not attainable without an information about the radio spectrum within
wide-enough spectral range and providing the radio luminosity of a given source at least three
or more different observing frequencies. For the large part of the Sample~2 we were able to
use flux densities from the radio catalogues: VLSS (74 MHz), 6C and 7C (151 MHz; Hales et al.
1988 and Riley et al. 1999, respectively), WENSS (325 MHz), B3 (408 MHz; Ficarra et al. 1985,
NVSS (1400 MHz) and GB6 (4850 MHz; Gregory et al. 1996). It is worth to notice that three
QSOs in Sample~2 have (projected) linear size larger than 1 Mpc!
\MakeTable{lllcrrccccl}{12.5cm}{The sample of FRII-type sources with 1$<$z$<$2 and D$>$400 kpc. The low-frequency luminosity marked {\sl a}, {\sl b}, {\sl c}
-- as in Table~1. The values in parenthesis are approximated because of less certain
subtraction of the core contribution}
{\hline
IAU name & Survey & z & Id. &$D$&$R_{\rm T}$&$\phi$ & log & log & log & Notes\\
& & & & [kpc] & &[$^{o}$]& $P_{lf}$ & $P_{1400}$ & $P_{4850}$ \\
(1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9) & (10) & (11)\\
\hline
J0245$+$0108 & & 1.537 & Q & 456 & 3.6 & 70 & 27.822$^{a}$ & 26.658 & 26.087 & \\
J0809$+$2015 & & 1.129 & Q & 465 & 3.1 & 70 & --- & (25.61)& (25.02)&
(1)\\
J0809$+$2912 & & 1.481 & Q & 1120 & 3.0 & 70 & 27.234$^{a}$ & 26.476 & 26.015 & \\
J0812$+$3031 & & 1.312 & Q & 1240 & 3.0 & 70 & 25.504$^{c}$ & 25.032 & --- &
(2)\\
J0819$+$0549 & & 1.701 & Q & 985 & 3.0 & 70 & --- & 25.578 & --- & (1,
2)\\
J0839$+$2928 & & 1.136 & Q & 417 & 2.3 & 45 & 26.813$^{a}$ & 25.882 & 25.364 & \\
J0842$+$2147 & & 1.182 & Q & 1080 & 3.0 & 70 & --- & 25.425 & --- & (1,
2)\\
J0857$+$0906 & & 1.688 & Q & 506 & 4.5 & 70 & 27.357$^{a}$ & 26.276 & 25.763 & \\
J0902$+$5707 & & 1.595 & Q & 862 & 4.0 & 70 & 25.905$^{c}$ & 25.325 & --- &
(2)\\
J0906$+$0832 & & 1.617 & Q & 682 & 3.6 & 70 & --- & 25.750 & --- & (1,
2)\\
J0947$+$5154 & & 1.063 & Q & 478 & 3.9 & 70 & 27.002$^{a}$ & 25.767 & 25.157 & \\
J0952$+$0628 & & 1.362 & Q & 551 & 4.3 & 70 & 27.081$^{a}$ & 25.988 & 25.508 & \\
B1011$+$365 & 6C & 1.042 & G & 416 & 3.5 & 70 & 26.902$^{b}$ & 26.204 & 25.680 & \\
J1030$+$5310 & & 1.197 & Q & 835 & 3.0 & 90 & 26.336$^{b}$ & 25.481 & 24.945 & \\
J1039$+$0714 & & 1.536 & Q & 501 & 3.0 & 70 & --- & 25.450 & --- & (1,
2)\\
B1108$+$359 &3C252& 1.105 & G & 493 & 3.0 & 70 & 28.164$^{a}$ & 26.921 & 26.282 & \\
B1109$+$437 & B3 & 1.664 & Q & 488 & 4.5 & 70 & 28.092$^{a}$ & 27.357 & 26.812 & \\
J1130$+$3628 & & 1.072 & Q & 422 & 3.0 & 50 & 25.992$^{b}$ & 25.277 & (24.67)& \\
J1207$-$0244 & & 1.100 & Q & 444 & 3.0 & 50 & 26.590$^{a}$ & 25.407 & --- & (2)\\
J1434$-$0123 & & 1.020 & Q & 490 & 3.0 & 50 & 26.894$^{a}$ & 25.763 & --- & (2)\\
J1550$+$3652 & & 2.061 & Q & 675 & 3.0 & 70 & 26.771$^{b}$ & 26.041 & 25.527 & \\
J1706$+$3214 & & 1.070 & Q & 438 & 3.0 & 50 & 26.757$^{a}$ & 25.682 & 25.150 & \\
B1723$+$510 &3C356& 1.079 & G & 614 & 4.0 & 70 & 27.875$^{b}$ & 26.916 & 26.304 & \\
J2345$-$0936 & & 1.275 & Q & 513 & 3.2 & 90 & 27.318$^{a}$ & 26.266 & (25.62)& \\
B2352$+$796 &3C469.1&1.336& G & 626 & 4.4 & 70 & 28.269$^{a}$ & 27.167 & 26.628 & \\
\hline
\multicolumn{11}{p{12.5cm}}{
Notes: (1)--off the 7C and WENSS surveys; (2)--off the 4850-MHz GB6 or PMN6 (Griffith et
al.
1995) surveys or below their flux density limit}
}
\begin{figure}[t]
\begin{center}
\includegraphics[width=12 cm]{j0809+2915}
\includegraphics[width=12 cm]{J1039+0714}
\FigCap{1400 MHz VLA maps of the sources J0809+2912 and J1039+0714. The NVSS images (gray
scale) are combined
with the FIRST contour maps. Total intensity, logarithmic contours are spaced by a factor of 2, starting with a value of 1.0 mJy/beam (except J1434-0123 starting with 0.2 mJy/beam).}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[width=12 cm]{j1130+3628}
\includegraphics[width=12 cm]{J1207-0244}
\FigCap{The same as in Fig.2 but for J1130+3628 and J1207-0244}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[width=12 cm]{J1434-0123}
\includegraphics[width=12 cm]{J1706+3214}
\FigCap{The same as in Fig.2 but for J1434-0123 and J1706+3214}
\end{center}
\end{figure}
\subsection{Sample 3 (3CRR sample)}
The third observational sample comprises 3CRR FRII-type sources with z$>$0.5 but size
smaller than 400 kpc (except of 3C265 and 3C292 which are larger). For the reason to have
homogeneous and complete data for radio spectra
of the sample sources necessary for estimation of their dynamical age and corresponding
physical parameters (jet power, mean expansion velocity, etc.), this sample is limited
to sources with the Galactic latitude of $\delta_{\rm II}>30^{o}$ where good spectral data
are available from the low-frequency radio surveys: VLSS, 6C, 7C and WENSS. The sample
sources are presented in Table~3.
\MakeTable{llcrrcccc}{12.5cm}{3CRR sample of FRII-type sources with z$>$0.5 and $\delta$ > 30$^{o}$}
{\hline
Source & z & Id. & $D$ & $R_{\rm T}$ & $\phi$ & log & log & log \\
& & & [kpc] & &[$^{o}$]& $P_{lf}$ & $P_{1400}$ & $P_{4850}$ \\
(1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9) \\
\hline
3C6.1 & 0.8404 & G & 199 & 2.7 & 70 & 27.706$^{a}$ & 26.938 & 26.494 \\
3C13 & 1.351 & G & 237 & 3.0 & 70 & 28.261$^{a}$ & 27.259 & 26.682 \\
3C22 & 0.937 & Q & 190 & 3.0 & 70 & 27.780$^{a}$ & 26.920 & 26.403 \\
3C34 & 0.689 & G & 341 & 4.3 & 90 & 27.656$^{a}$ & 26.474 & 25.815 \\
3C41 & 0.794 & G & 175 & 2.4 & 70 & 27.404$^{b}$ & 26.851 & 26.491 \\
3C54 & 0.8284 & G & 401 & 3.9 & 70 & 27.670$^{a}$ & 26.644 & 26.168 \\
3C65 & 1.174 & G & 142 & 2.3 & 70 & 28.077$^{b}$ & 27.296 & 26.761 \\
3C68.1 & 1.238 & Q & 386 & 2.4 & 50 & 27.967$^{b}$ & 27.239 & 26.767 \\
3C68.2 & 1.575 & G & 191 & 2.3 & 70 & 28.290$^{b}$ & 27.259 & 26.548 \\
3C169.1 & 0.633 & G & 315 & 2.7 & 70 & 27.287$^{a}$ & 26.214 & 25.700 \\
3C184 & 0.99 & G & 35 & 4.0 & 70 & 27.832$^{a}$ & 27.017 & 26.514 \\
3C196 & 0.871 & Q & 39 & 1.5 & 30 & 28.522$^{a}$ & 27.605 & 27.117 \\
3C204 & 1.112 & Q & 257 & 3.4 & 70 & 28.032$^{a}$ & 26.857 & 26.267 \\
3C205 & 1.534 & Q & 134 & 4.2 & 70 & 28.329$^{a}$ & 27.469 & 26.903 \\
3C217 & 0.898 & G & 86 & 4.0 & 70 & 27.796$^{a}$ & 26.853 & 26.318 \\
3C220.1 & 0.610 & G & 86 & 4.0 & 70 & 27.541$^{a}$ & 26.469 & 25.875 \\
3C220.3 & 0.685 & G & 52 & 3.0 & 70 & 27.425$^{b}$ & 26.710 & 26.045 \\
3C239 & 1.786 & G & 96 & 1.4 & 70 & 28.483$^{b}$ & 27.541 & 26.914 \\
3C247 & 0.749 & G & 95 & 2.4 & 70 & 27.350$^{b}$ & 26.746 & 26.284 \\
3C254 & 0.734 & Q & 96 & 2.0 & 50 & 27.974$^{a}$ & 26.791 & 26.198 \\
3C263 & 0.6563 & Q & 307 & 2.7 & 50 & 27.616$^{a}$ & 26.634 & 26.182 \\
3C263.1 & 0.824 & G & 41 & 1.7 & 70 & 27.696$^{b}$ & 26.911 & 26.369 \\
3C265 & 0.8108 & G & 587 & 3.9 & 70 & 27.775$^{b}$ & 26.859 & 26.310 \\
3C266 & 1.275 & G & 36 & 2.4 & 70 & 27.981$^{b}$ & 27.137 & 26.509 \\
3C268.4 & 1.400 & Q & 83 & 2.8 & 50 & 28.123$^{a}$ & 27.273 & 26.774 \\
3C270.1 & 1.519 & Q & 102 & 1.3 & 50 & 28.314$^{a}$ & 27.478 & 26.917 \\
3C272 & 0.944 & G & 461 & 3.3 & 70 & 27.417$^{b}$ & 26.708 & 26.135 \\
3C280 & 0.996 & G & 113 & 1.5 & 70 & 27.990$^{b}$ & 27.288 & 26.829 \\
3C280.1 & 1.659 & Q & 169 & 4.0 & 70 & 28.267$^{a}$ & 27.343 & 26.771 \\
3C289 & 0.967 & G & 81 & 1.3 & 70 & 27.686$^{b}$ & 26.966 & 26.430 \\
3C292 & 0.71 & G & 960 & 4.2 & 90 & 27.406$^{b}$ & 26.547 & 26.075 \\
3C294 & 1.779 & G & 135 & 1.9 & 70 & 28.413$^{b}$ & 27.431 & 26.840 \\
3C322 & 1.681 & G & 283 & 2.5 & 70 & 28.171$^{b}$ & 27.485 & 26.954 \\
3C324 & 1.207 & G & 85 & 1.8 & 70 & 27.999$^{b}$ & 27.257 & 26.712 \\
3C325 & 0.860 & G & 122 & 2.2 & 70 & 27.657$^{b}$ & 27.011 & 26.481 \\
3C330 & 0.549 & G & 395 & 5.4 & 90 & 27.432$^{b}$ & 26.771 & 26.334 \\
3C337 & 0.635 & G & 295 & 2.5 & 70 & 27.522$^{a}$ & 26.599 & 26.125 \\
3C352 & 0.8057 & G & 75 & 2.3 & 70 & 27.695$^{a}$ & 26.699 & 26.108 \\
3C427.1 & 0.572 & G & 153 & 2.2 & 70 & 27.437$^{b}$ & 26.634 & 26.071 \\
3C437 & 1.48 & G & 316 & 3.9 & 70 & 28.086$^{b}$ & 27.481 & 27.030 \\
3C441 & 0.707 & G & 236 & 2.5 & 70 & 27.395$^{b}$ & 26.636 & 26.166 \\
3C470 & 1.653 & G & 205 & 4.0 & 70 & 28.269$^{a}$ & 27.444 & 26.952 \\
\hline
}
\section{Ageing analysis of the samples' sources}
\subsection{Application of the DYNAGE algorithm}
The dynamical age analysis is performed using the DYNAGE algorithm of Machalski
et al. (2007a). It is based on the analytical KDA model (cf. Introduction).
The original KDA model, assuming values for a number of free parameters of the
model (e.g. the jet power $Q_{\rm jet}$, radius $a_{0}$ and density $\rho_{0}$
of the radio core, exponent $\beta$ in the power-law density distribution in
the external medium $\rho(d/a_{0})^{-\beta}$, and the effective injection
spectral index $\alpha_{\rm inj}$, which approximates the initial relativistic
particle continuum at a head of the jet $n(\gamma_{i})=n_{0}(\gamma_{i}^{-p})$,
where $\gamma_{i}$ is the Lorentz factor of these particles and
$p=(\alpha_{\rm inj}+1)/2)$, allows prediction of a time evolution of the
source's parameters, e.g. its length, $D(t)$, and radio luminosity,
$P_{\nu}(t)$, at a given observing frequency -- as a function of its assumed
age $t$.
Oppositely, the DYNAGE algorithm enables solving a reverse problem, i.e.
finding the values of $t$, $\alpha_{\rm inj}$, $Q_{\rm jet}$, and $\rho_{0}$
for a real source. Determining the values of these four free parameters of the
model is possible by a fit to the observational parameters of a source: its
projected linear size $D$, the volume $V$, the radio monochromatic luminosity
$P_{\nu}$, and the radio spectrum $\alpha_{\nu}$ that provides $P_{\nu,i}$ at
a number of observing frequencies $i=1, 2, 3,...$ To determine these
luminosities for the sample sources, we have had to fix the flux density in
their lobes at a number of frequencies, i.e. to subtract a flux contribution
from the compact components such as the radio core and hot spots in these
lobes. This was especially important for the sample sources identified with
quasars. The resulting luminosities at three observing frequencies: the lowest
frequency used (74, 151, or 325 MHz), the medium one of 1400 MHz, and the
highest one of 4850 MHz, are given in columns (8)--(10) in Tables 1 and 2, as
well as in columns (7)--(9) in Table~3.
As in KDA, we adopted a cylindrical geometry of the source's cocoon (radio
lobes) where its volume is determined by the deprojected length $D/\sin\phi$
and the axial ratio $R_{\rm T}$. The values of $R_{\rm T}$ are estimated from
the low-frequency radio images, and an angle of orientation of the jet axis to
the observer's line of sight, $\phi$, is subjectively estimated from the
observed asymmetry in the lobes' arms and brightness. The valus of these two
parameters are given in columns (6) and (7) of Tables 1 and 2, and in columns
(5) and (6) in Table~3.
Unfortunately, the values of other free parameters of the model have to be assumed. These
are: $\gamma_{i,{\rm min}}=1$ and $\gamma_{i,{\rm max}}=10^{7}$
that are the Lorentz factors determining the energy range of the relativistic particles
used in integration of their initial power-law distribution; $\Gamma_{\rm j}$,
$\Gamma_{\rm x}$,
$\Gamma_{\rm B}$, and $\Gamma_{\rm c}$, are all equal to 5/3, are the adiabatic indices in the equation of state
for the jet material, the unshocked medium surrounding the lobes, a "magnetic" fluid, and
the source (cocoon) as a whole, respectively; and $k^{\prime}=0$ is the ratio of the
energy density of thermal particles to that of the relativistic particles. The more
detailed description of application of the above algorithm to observed radio structures is
published in Machalski et al. (2009).
\subsection{Results of the modelling}
The resulting values of the age and other physical parameters of the sample sources are
given in Tables 4, 5 and 6 (for the Sample 1, 2 and 3, respectively). Columns (2) and
(3) give the dynamical age, $t$, and the effective, initial slope of the radio spectrum
$\alpha_{\rm inj}$. Columns (4) -- (7): (in the logarithmic scale) the jet power, $Q_{\rm jet}$,
the central core density, $\rho_{0}$, the pressure along the jet axis, $p_{\rm h}$, and the
total energy radiated out during the age of source, $U_{\rm c}$. The age $t$ and the fitted
values of $Q_{\rm jet}$ and $\rho_{0}$ fulfil the dynamical equation
\begin{equation}
t=\left(\frac{D}{2\,c_{1}\sin\phi}\right)^{\frac{5-\beta}{3}}\left(\frac{\rho_{0}a_{0}^{\beta}}
{Q_{\rm jet}}\right)^{1/3},
\end{equation}
\noindent
Where $D/(2\,\sin \phi)\equiv D_{\ell}$ is the deprojected lenght of a given lobe treated
here as one
half of the total, observed size of the source, $D$. The energy density fulfilling the energy equipartition condition is calculated from
\begin{equation}
u_{\rm eq}(t)=\frac{18\,c_{1}^{(2-\beta)}}{(\Gamma_{\rm x}+1)(\Gamma_{\rm c}-1)(5-\beta)^{2}
{\cal P}_{\rm hc}}\left(\rho_{0}a_{0}^{\beta}\right)^{\frac{3}{5-\beta}}
Q_{\rm jet}^{\frac{2-\beta}{5-\beta}}t^{-\frac{4+\beta}{5-\beta}},
\end{equation}
\noindent
where ${\cal P}_{\rm hc}$=$(2.14-0.52\beta)R_{\rm T}^{2.04-0.25\beta}$ is the empirical formula
for the pressure ratio along the jet axis and the transverse direction taken from Kaiser (2000).
The total radiated energy is simply $U_{\rm c}$=$u_{\rm eq}\times V_{\rm c}$, where the source
(cocoon) volume attained at the age $t$ is $V_{\rm c}$=$(\pi/4)\,D^{3}/(2R_{\rm T})^{2}$.
Column (8) gives the magnetic field strength estimate derived from
\begin{equation}
B=\left(\frac{24}{7}10^{11}\pi\,u_{\rm B}\right)^{1/2}\hspace{2mm}[{\rm nT}],
\end{equation}
\noindent
where the magnetic field energy density is $u_{\rm B}$=$u_{\rm eq}(1+p)/(5+p)$. The last
column (9) gives the actual expansion velocity of the lobe's head along the jet axis which is the derivative
of the $D_{\ell}$ function
\begin{equation}
{\rm v}_{\rm h}(t)=\frac{dD_{\ell}}{dt}=\frac{3\,c_{1}}{5-\beta}\left(\frac{Q_{\rm
jet}}{\rho_{0}a_{0}^{\beta}}
\right)^{\frac{1}{5-\beta}}t^{\frac{\beta -2}{5-\beta}}.
\end{equation}
\section{Observational constrain for the model}
\subsection{Comparison of the model's prediction with the observational data}
The model's predictions, i.e. the dependence of the deprojected linear size and the jet's head
velocity on the source's age (${D}/{sin\phi}$ vs $t$ and $v_{\rm h}/c$ vs $t$,
respectively), are compared with the data determined for the sampled sources and given in
Tables 4, 5 and 6. Fig. 5 shows the dependence of D/sin$\phi$ on $t$. The model's
predictions are presented for three different sets of values of the jet power and redshift:
$z=0.5$ and $Q_{\rm jet}=10^{37.5}$ W, $z=1$ and $Q_{\rm jet}=10^{38.5}$ W, $z = 2$ and
$Q_{\rm jet}=10^{39.5}$ W, marked "1", "2", and "3", respectively. The above dependence
predicted in the frame of scenario A are drawn with the solid lines, while these for scenario B
-- with the dashed lines. The points of bifurcation of the model's predictions into scenario A
and scenario B correspond to the halo diameter dependent on redshift, $2R_{\rm h}(z)$ (on the ordinate axis)
and to the related age (on the abscissa axis). The ends of both solid and dashed lines
indicate a maximum size and age at which the predicted expansion velocity reaches the speed of sound. The crosses show the distribution of the GRSs from Sample 1,
the full circles -- the distant sources (mostly QSOs) from Sample 2, and the open circles -- the 3CRR sources from Sample 3.
The dependence of v$_{\rm h}/c$ on $t$ is shown in Fig. 6. The model's predictions and the sample sources are indicated with the lines and symbols as in Fig. 5. The rapid increase of v$_{\rm h}/c$ in the frame of scenario A results from the discontinuities of ${\rho(d)}$ and $kT(d)$ at the halo-IGM interface (cf. Fig. 1). The three horizontal dotted lines indicate the predicted lower values of the axial expansion velocity,
v$_{\rm h,min}/c$, limited by the sound speed at the given values of redshift.
The maximum size which any FRII-type radio source can reach with a given value of v$_{\rm h,min}/c$ , (i.e. at the corresponding redshift) depends on the jet power. The size $D_{\rm max}$ in the frame of scenario A, as a function of v$_{\rm h,min}/c$ (or $z$), is plotted on Fig. 7 for three values of $Q_{\rm jet}$ considered above.
\MakeTable{lclcccrccccll}{12.5cm}{Age and physical parameters of the sources in Sample~1}
{\hline
IAU name &&t&$\alpha_{\rm inj}$&log\,$Q_{\rm jet}$&log\,$\rho_{0}$&log\,$p_{\rm h}$ &log\,U$_{\rm c}$&
B$_{\rm eq}$ & v$_{\rm h}$/c \\
&&[Myr]& & [W] & [kg/m$^{3}$] & [N/m$^{2}$] & [J] & [nT] \\
(1) & & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9)\\
\hline
J0037+0027 &&$\;\:$59.0 & 0.506 & 38.474 & $-$23.311 & $-$12.13 & 53.00 & 0.16 & 0.047 \\
B0654+482 &&$\;\:$57.0 & 0.555 & 38.296 & $-$22.741 & $-$11.68 & 52.92 & 0.33 & 0.024 \\
J0750+656 &&$\;\:$39.0 & 0.507 & 38.563 & $-$23.670 & $-$12.20 & 53.02 & 0.18 & 0.057 \\
B0821+695 && 135.0 & 0.645 & 38.710 & $-$22.968 & $-$12.45 & 53.84 & 0.17 & 0.027 \\
B0854+399 &&$\;\:$69.0 & 0.522 & 38.362 & $-$22.992 & $-$12.04 & 53.24 & 0.29 & 0.023 \\
B1058+368 &&$\;\:$57.0 & 0.545 & 38.384 & $-$22.741 & $-$11.66 & 52.98 & 0.32 & 0.027 \\
J1130$-$1320&& 100.0& 0.537 & 39.005 & $-$22.431 & $-$11.71 & 53.90 & 0.33 & 0.028 \\
B1602+376 &&$\;\:$47.0 & 0.524 & 38.621 & $-$23.648 & $-$12.35 & 53.35 & 0.22 & 0.041 \\
B1636+418 &&$\;\:$38.0 & 0.526 & 38.367 & $-$23.229 & $-$11.82 & 52.83 & 0.29 & 0.037 \\
B1834+620 &&$\;\:$75.0 & 0.533 & 38.455 & $-$22.688 & $-$11.80 & 53.18 & 0.28 & 0.026 \\
J1951+706 &&$\;\:$73.0 & 0.525 & 38.061 & $-$22.921 & $-$12.02 & 52.73 & 0.25 & 0.025 \\
J2234$-$0224&&$\;$(91.0)& & (38.1) &($-$22.55) & & & & 0.019 \\
\hline
}
\MakeTable{lllccrccccl}{12.5cm}{Age and physical parameters of the sources in Sample~2}
{\hline
IAU name &&t&$\alpha_{\rm inj}$&log\,$Q_{\rm jet}$&log\,$\rho_{0}$&log\,$p_{\rm h}$ &log\,U$_{\rm c}$&
B$_{\rm eq}$ & v$_{\rm h}$/c \\
&&[Myr]& & [W] & [kg/m$^{3}$] & [N/m$^{2}$] & [J] & [nT] \\
(1) & & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9)\\
\hline
J0245+0108 && 18.5 & 0.603 & 39.295 & $-$22.153 & $-$10.30 & 53.45 & 1.74 & 0.038 \\
J0809+2912 &&$\;\:$7.5 & 0.496 & 39.715 & $-$24.555 & $-$11.65 & 53.59 & 0.41 & 0.233 \\
J0839+2928 && 21.5 & 0.535 & 38.655 & $-$23.031 & $-$11.44 & 52.93 & 0.62 & 0.031 \\
J0857+0906 && 14.6 & 0.564 & 39.111 & $-$22.629 & $-$10.48 & 53.08 & 1.13 & 0.054 \\
J0947+5154 && 50.5 & 0.624 & 38.508 & $-$21.616 & $-$10.62 & 53.06 & 1.12 & 0.015 \\
J0952+0628 && 23.0 & 0.554 & 38.806 & $-$22.453 & $-$10.75 & 52.98 & 0.89 & 0.037 \\
B1011+365 && 18.1 & 0.542 & 38.752 & $-$22.686 & $-$10.76 & 52.94 & 1.00 & 0.036 \\
J1030+5310 && 28.0 & 0.548 & 38.605 & $-$23.408 & $-$11.77 & 53.05 & 0.38 & 0.042 \\
B1108+359 && 23.3 & 0.655 & 39.560 & $-$21.913 & $-$10.22 & 53.90 & 2.18 & 0.031 \\
B1109+437 &&$\;\:$4.3 & 0.544 & 40.028 & $-$23.238 & $-$10.04 & 53.47 & 1.87 & 0.175 \\
J1130+3628 && 29.5 & 0.536 & 38.180 & $-$22.898 & $-$11.39 & 52.53 & 0.47 & 0.027 \\
J1550+3652 &&$\;\:$6.6 & 0.511 & 39.340 & $-$24.327 & $-$11.42 & 53.16 & 0.53 & 0.160 \\
J1706+3214 && 36.6 & 0.572 & 38.460 & $-$22.340 & $-$11.00 & 52.90 & 0.72 & 0.022 \\
B1723+510 && 23.5 & 0.661 & 39.676 & $-$21.881 & $-$10.09 & 53.92 & 1.98 & 0.041 \\
J2345$-$0936 && 29.6 & 0.564 & 38.797 & $-$22.308 & $-$10.82 & 53.23 & 1.05 & 0.024 \\
B2352+796 && 16.0 & 0.578 & 39.747 & $-$22.155 & $-$10.08 & 53.74 & 1.83 & 0.058 \\
\hline
}
\MakeTable{lccccrcrc}{12.5cm}{Age and physical parameters of the sources in Sample~3.The age solution for the sources which name is followed by the asterisk is provided by J. Machalski (unpublished).}
{\hline
IAU name &t&$\alpha_{\rm inj}$&log\,$Q_{\rm jet}$&log\,$\rho_{0}$&log\,p$_{\rm h}$&log\,U$_{\rm c}$&
B$_{\rm eq}$ & v$_{\rm h}$/c \\
&[Myr]& & [W] & [kg/m$^{3}$] & [N/m$^{2}$] & [J] & [nT] \\
(1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) & (9)\\
\hline
3C6.1 &$\;\;$ 5.12 & 0.564 & 39.298 & $-$22.891 & $-$10.08 & 53.03 & 2.75 & 0.057 \\
3C13 &$\;\;$ 5.87 & 0.583 & 39.606 & $-$22.549 & $-$9.81 & 53.35 & 3.41 & 0.060 \\
3C22 &$\;\;$ 4.68 & 0.559 & 39.316 & $-$22.799 & $-$9.92 & 52.96 & 3.03 & 0.060 \\
3C34 & 24.0 & 0.630 & 39.012 & $-$21.434 & $-$9.86 & 53.22 & 2.56 & 0.018 \\
3C41 &$\;\;$ 1.52 & 0.534 & 39.580 & $-$24.250 & $-$10.34 & 52.85 & 2.21 & 0.182 \\
3C54* & 15.2 & 0.562 & 39.141 & $-$22.280 & $-$10.26 & 53.17 & 1.64 & 0.039 \\
3C65 &$\;\;$ 3.40 & 0.588 & 39.604 & $-$22.881 & $-$9.74 & 53.23 & 4.65 & 0.065 \\
3C68.1* &$\;\;$ 7.94 & 0.564 & 39.784 & $-$23.053 & $-$10.43 & 53.66 & 1.70 & 0.089 \\
3C68.2 &$\;\;$ 4.35 & 0.771 & 40.209 & $-$22.406 & $-$9.41 & 53.94 & 6.97 & 0.076 \\
3C169.1 & 24.7 & 0.575 & 38.611 & $-$22.230 & $-$10.68 & 53.02 & 1.37 & 0.019 \\
3C184 &$\;\;$ 0.37 & 0.577 & 39.471 & $-$23.395 & $-$8.46 & 52.01 & 12.83& 0.142 \\
3C196 &$\;\;$ 0.60 & 0.579 & 40.042 & $-$23.374 & $-$8.91 & 52.82 & 10.14& 0.188 \\
3C204 & 11.6 & 0.625 & 39.350 & $-$21.969 & $-$9.76 & 53.36 & 3.28 & 0.035 \\
3C205 &$\;\;$ 1.45 & 0.584 & 39.949 & $-$22.840 & $-$8.98 & 52.95 & 6.76 & 0.143 \\
3C217 &$\;\;$ 1.87 & 0.588 & 39.301 & $-$22.481 & $-$8.97 & 52.41 & 7.08 & 0.068 \\
3C220.1 & 11.0 & 0.610 & 38.900 & $-$21.653 & $-$9.53 & 52.78 & 3.74 & 0.024 \\
3C220.3 &$\;\;$ 1.92 & 0.580 & 38.955 & $-$22.345 & $-$8.97 & 52.21 & 9.01 & 0.040 \\
3C239* &$\;\;$ 2.90 & 0.673 & 39.903 & $-$22.904 & $-$9.68 & 53.60 & 7.61 & 0.052 \\
3C247* &$\;\;$ 1.70 & 0.547 & 39.182 & $-$23.182 & $-$9.88 & 52.50 & 3.77 & 0.087 \\
3C254 &$\;\;$ 5.60 & 0.627 & 39.172 & $-$22.240 & $-$9.60 & 52.98 & 5.18 & 0.031 \\
3C263 & 12.8 & 0.557 & 39.126 & $-$22.584 & $-$10.44 & 53.15 & 1.53 & 0.044 \\
3C263.1* &$\;\;$ 1.30 & 0.600 & 39.133 & $-$23.110 & $-$9.40 & 52.43 & 8.83 & 0.050 \\
3C265* & 27.0 & 0.609 & 39.415 & $-$21.905 & $-$10.25 & 53.72 & 1.66 & 0.034 \\
3C266* &$\;\;$ 0.86 & 0.639 & 39.462 & $-$22.674 & $-$8.64 & 52.48 & 16.12& 0.065 \\
3C268.4 &$\;\;$ 0.95 & 0.574 & 39.809 & $-$23.280 & $-$9.15 & 52.70 & 6.56 & 0.159 \\
3C270.1* &$\;\;$ 2.35 & 0.568 & 39.708 & $-$23.458 & $-$10.07 & 53.26 & 4.30 & 0.079 \\
3C272* & 19.0 & 0.557 & 39.132 & $-$22.484 & $-$10.57 & 53.37 & 1.32 & 0.038 \\
3C280* &$\;\;$ 2.60 & 0.560 & 39.520 & $-$23.578 & $-$10.23 & 53.16 & 3.76 & 0.068 \\
3C280.1 &$\;\;$ 2.50 & 0.574 & 39.795 & $-$22.635 & $-$9.23 & 53.03 & 5.24 & 0.100 \\
3C289* &$\;\;$ 3.60 & 0.571 & 39.080 & $-$23.296 & $-$10.30 & 52.88 & 3.92 & 0.035 \\
3C292* & 38.0 & 0.566 & 39.180 & $-$22.266 & $-$10.86 & 53.60 & 0.81 & 0.035 \\
3C294* &$\;\;$ 3.65 & 0.683 & 39.950 & $-$22.611 & $-$9.54 & 53.67 & 6.95 & 0.058 \\
3C322* &$\;\;$ 4.00 & 0.557 & 39.908 & $-$23.310 & $-$10.16 & 53.58 & 2.66 & 0.111 \\
3C324* &$\;\;$ 2.20 & 0.584 & 39.471 & $-$23.117 & $-$9.71 & 52.98 & 5.91 & 0.061 \\
3C325* &$\;\;$ 2.90 & 0.566 & 39.320 & $-$23.197 & $-$9.95 & 52.89 & 3.77 & 0.066 \\
3C330* &$\;\;$ 6.60 & 0.543 & 39.404 & $-$22.699 & $-$9.98 & 52.93 & 1.82 & 0.084 \\
3C337 & 11.3 & 0.553 & 38.999 & $-$22.853 & $-$10.64 & 53.10 & 1.53 & 0.039 \\
3C352 &$\;\;$ 2.96 & 0.595 & 38.986 & $-$22.631 & $-$9.55 & 52.53 & 5.78 & 0.038 \\
3C427.1* &$\;\;$ 9.10 & 0.592 & 38.903 & $-$22.498 & $-$10.18 & 52.98 & 2.91 & 0.026 \\
3C437* &$\;\;$ 2.42 & 0.546 & 40.113 & $-$23.405 & $-$9.80 & 53.37 & 2.78 & 0.205 \\
3C441* & 10.0 & 0.556 & 38.984 & $-$22.764 & $-$10.45 & 53.05 & 1.91 & 0.037 \\
3C470 &$\;\;$ 1.97 & 0.550 & 39.973 & $-$23.063 & $-$9.41 & 53.10 & 4.25 & 0.155 \\
\hline
}
\subsection{Age and physical parameters of the sample sources}
The three samples used differ significantly in the distribution of fitted age of their members.
A median of age in the Sample 1, 2 and 3 is $65\pm 8$ Myr, $22\pm 2.5$ Myr, and $4\pm 1$ Myr,
respectively. A median linear size in these samples is $\sim$1330 kpc, $\sim$510 kpc, and $\sim$145
kpc, respectively. As the Samples 1 and 2 comprise the largest sources (of FRII type only) each
of them in different redshift range, and this range in the Sample 2 is twice the range of the
Sample 1 -- we compare the $D-z$ dependence for these sources with that found in other samples
unlimited in linear size of their members (e.g. Eales 1985; Barthel \& Miley 1988). Since the
Sample 1, i.e. the sample of known high-redshift GRSs, is very small and
consists of 12 sources only, below we also consider 12 of the largest ones from the
Sample 2.
The relevant medians are: $D_{\rm med}=1330\pm 120$ kpc at $z_{\rm med}=0.60\pm 0.10$, and
$D_{\rm med}=760\pm 80$ kpc at $z_{\rm med}=1.35\pm 0.20$, respectively. Assuming that
$D_{\rm med}\propto (1+z)^{-n}_{\rm med}$, we find $n\approx 1.5$ which is compatible with
a value of this power $\sim$(1--2) determined in several samples of radio galaxies and quasars
(e.g. Kapahi 1989).
Therefore the above result agree with the trend observed in much more abundant samples of sources
and confirms that the largest linear size, which a source can achieve before dimming
below the detection limit, inevitably decreases with redshift irrespective of
a cosmological evolution of the IGM (cf. Nilsson et al. 1993).
In order to enlighten the latter problem, we pay an attention to the pressure at the head of
lobes (along the jet's axis), $p_{\rm h}$, given in column 6 of Tables 4, 5 and 6. This
pressure is determined in the DYNAGE as
\[p_{\rm h}={\cal P}_{\rm hc}(R_{\rm T})p_{\rm c}(R_{\rm T},t),\]
\noindent
where ${\cal P}_{\rm hc}$ is the pressure ratio (cf. Section 4.2), and $p_{\rm c}=(\Gamma_{\rm c}-1)u_{\rm eq}$, where $u_{\rm eq}$ is given by Eq.\,11.
The median of this pressure, $p_{\rm h,med}$, for sources in the Sample 1 and Sample 2 is
$\sim 10^{-12.0}$ N/m$^{2}$ and $\sim 10^{-10.8}$ N/m$^{2}$, respectively, thus a (mean) pressure at the head of lobes of the sources within the redshift range $z$[1, 2] is about 16 times higher
than that for the sources within $z$[0.5, 1]. If so, it is more than twice higher than a ratio
implied from the formula giving pressure of a non-relativistic, homogeneous IGM in thermal
equilibrium, $p(z)=p_{0}(1+z)^{5}$, where $p_{0}$ is the present-day pressure (Subrahmanyan \&
Saripalli 1993). We return to this point in the next section.
\begin{figure}[t]
\begin{center}
\includegraphics[width=12cm]{Kuligowska2.ps}
\FigCap{ $D$ vs $t$ diagrams for three sets of the jet power and redshift. The solid lines indicate
the model predictions in the frame of Scenario A, the dashed lines -- in Scenario B. The sources
from the Samples 1, 2, and 3 are plotted with the crosses, full dots, and open circles, respectively.}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[width=12cm]{Kuligowska3.ps}
\FigCap{$v_{\rm h}/c$ vs $t$ diagrams for the same sets of the model parameters as in Fig.\,5. The
three horizontal dotted lines indicate the lower limits for the expansion velocity determined by
the sound speed at a given redshift.}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[width=10cm]{Kuligowska4.ps}
\FigCap{ $D_{\rm max}$ vs $v_{\rm min}/c$ diagrams for three values of the jet power.}
\end{center}
\end{figure}
\begin{center}
\section{Discussion of the results and conclusions}
\end{center}
The distribution of sources on the planes ${D}/{sin\phi}$ -- $t$ and $V_{\rm h}/c$ -- $t$
(in Fig. 5 and 6, respectively) is more or less compatible with model's prediction, however the samples used to constrain the model may be to small to be decisive. So, much larger samples of sources with reliably determined ages would support or impair the inferences drawn as follows:
(i) Scenario B is rather excluded because both the maximum size, $D_{\rm max}\sim 600$ kpc,
and the corresponding age, $t_{\rm max}\sim 70$ Myr allowed by the model are evidently smaller
and younger then the observed values of those parameters for the sample sources.
Also high-dynamics radio observations show
no evidence for a flaring of the jets in FRII-type sources; oppositely the jets in these sources
are rather recollimated in vicinity of the radio core.
(ii) As expected, the inferred expansion velocity, v$_{\rm h}/c$, of all the sources used to
constrain the G-KW model are higher than the limiting sound speed marked in Fig.\.6.
This result confirms a common believe that the heads of lobes of FRII type
sources are overpressured with respect to the external gaseous medium. We note that the lowest
expansion velocities of the largest sources, i.e. these in the Samples 1 and 2, are comparable
to those of much smaller sources in the Sample 3.
(iii) The age of the three sources in Sample 2 is probably underestimated and the corresponding
expansion velocity -- overestimated (the three full dots with $t<10$ Myr and v$_{\rm h}>0.1\,c$
in Figs.\,5 and 6) due to possible relativistic effects (cf. Arshakian \& Longair 2000). All
are classified as quasars, and the two of them (J0809+2912 and J1550+3652) are highly
asymmetric in their lobes' brightness. If a probable anisotropic radiation is discerned and
the proper luminosity of the cocoon is taken into calculations, the age of these sources would
be older and the expansion velocity -- lower than the values in Table 6.
(iv) The pressure in the diffuse lobes of the largest radio sources seems to offer a tool useful for an
estimation of the IGM pressure. If the axial pressure in lobes of the largest observed sources,
especially these in the Sample 2, is close to equilibrium with the
IGM pressure, the inferred values of $p_{\rm h}$ would suggest even stronger density evolution
of the IGM than $\rho_{\rm IGM}\propto (1+z)^{3}$ (cf. Eqs.\,5 and 6). However, radio images
of the sources in the Sample 2 evidently indicate a presence of hot spots which, in turn,
may confirm that the hot spot regions are highly overpressured with respect to the IGM, and
the actual age of sources cannot be considered as a lifetime. Moreover, the DYNAGE fits show that
the jet powers of the most distant sources are significantly higher than that for sources of a
similar age but at low redshifts $z<0.2$. As it was shown in Section 5.2, the median
age in the Sample 2 is $\sim 22$ Myr. An
inspection of the compilation of over 200 radio
galaxies and quasars with the dynamical ages fitted using the DYNAGE method (the data
unpublished yet) resulted in only 7 radio galaxies of age of about 25 Myr and lying within the
redshift range $z$[0.1, 0.2]. These galaxies and their observational and dynamical parameters
are listed in Table 7.
\begin{table}
\caption{Observational and dynamical parameters of seven FRII-type radio galaxies with $0.1^{<}_{\sim} z ^{<}_{\sim}0.2$, for which their
dynamical age determined with the DYNAGE is 12 Myr$<$t$<$36 Myr.}
\begin{tabular}{llrcccc}
\hline
Name & z & $D$[kpc] & log\,$P_{1400}$ & $t$[Myr] & log\,$Q_{\rm jet}$ & log\,$\rho_{0}$\\
& & & [W/(Hz$\cdot$sr)] & & [W] & [kg/m$^{3}$]\\
\hline
3C332 & 0.1515 & 229 & 25.07 & 36 & 37.45 & $-$22.60\\
3C349 & 0.205 & 287 & 25.46 & 29 & 37.91 & $-$22.44\\
3C357 & 0.1664 & 296 & 25.20 & 24 & 37.77 & $-$22.96\\
3C381 & 0.1605 & 199 & 25.29 & 25 & 37.66 & $-$22.76\\
B0908+376 & 0.1047 & 229 & 24.12 & 20 & 36.50 & $-$23.59\\
B1130+339 & 0.2227 & 99 & 24.98 & 12 & 37.38 & $-$23.01\\
B1457+292 & 0.146 & 172 & 24.15 & 28 & 36.80 & $-$23.99\\
\hline
\end{tabular}
\end{table}
The median jet power, $Q_{\rm jet,med}$, in the Sample 2 (cf. Table 5) is $\sim 1.8\times 10^{39}$ W,
while that for the galaxies in Table 8 is $\sim 2.8\times 10^{37}$ W. We argue that this is very
unlikely to find so young FRII type radio sources at redshift below 0.2 and driven by jets more
powerful than $10^{38}$ W. Although this may be partly caused by the selection effect. In fact, searching
for high-redshift sources we probe a much larger space volume than the volume
corresponding to a low redshift. Probably therefore this is why the only
low-redshift and very powerful radio galaxy known is Cygnus A with $t\approx 8$
Myr, $D=135$ kpc, and $Q_{\rm jet}\approx 1.4\times 10^{39}$ W (cf. Machalski
et al. (2007a), the above result strongly suggests that the dynamical evolution of FRII
type radio sources at high redshifts (in earlier cosmological epochs) is different and
faster than that of sources at low redshifts. The model constrained in this paper implies
that giant-sized sources (with $D>1$ Mpc) can exist at redshifts as high as $\sim 2$,
however only if their jets are powerful enough.
(v) A differentiation of the jet's (highest) power at high and low redshifts perhaps would be a strong
argument for the theoretical speculations about its dependence on the properties of a black hole
(BH) in AGN: the spin of the BH (Blandford \& Znajek 1977) and/or the accretion process and its rate
(Sikora, Stawarz \& Lasota 2007; Sikora 2009). This is very likely that the accretion of matter
onto the BH must play a dominant role in the jets' production. This led to a presumption that more
powerful jets can be due to a larger amount of material available for the accretion processes inside
AGN formed at earlier cosmological epochs.
Reasuming the above we can conclude as follows:
--- An observational quest for the largest radio sources of FRII type at high redshifts
$1<z<2$ resulted in the sample of 25 sources listed in Table 2, where 20 of 25 are found
in this paper. Because the finding sky survey used was the optical SDSS survey, all the
newly radio-identified optical objects appear to be quasars with the (projected) linear
size from $\sim$400 kpc to $\sim$1200 kpc. The above bias precluded a detection of large,
most distant and low-luminosity radio galaxies whose parent optical counterpart will likely
be of $\sim$(22--24) R mag.
--- Though the samples used to constrain the model are small, the observational data seem
to be concordant with the predictions in the frame of Scenario A. However much larger samples
of distant sources with reliably determined ages will be more decisive for the above aim.
--- The derived lowest values of the lobes' head pressure, $p_{\rm h}$, evidently higher
than the limiting sound speed at different redshifts, strongly suggest that heads of even
the largest sources at high redshifts are still overpressured with respect to the IGM.
-- An existence of giant-sized radio sources at high redshifts is possible due to extremely
high power of their jets up to $\sim 10^{40}$ W. Such the highest jet powers are likely
related to the accretion processes onto massive black holes in the central AGN, which might
be very efficient in the nuclei formed at earlier cosmological epochs.
\Section{Appendix}
Table 8 gives precise sky coordinates of the parent optical object selected from the SDSS survey and identified
with the given radio source, its (largest) angular size
measured in the FIRST images, and the 1400-MHz total flux density taken from
the NVSS catalogue.
\MakeTable{lrcccrccccl}{12.5cm}{The newly discovered FRII-type sources with 1$<$z$<$2 and
D$>$400 kpc in Sample 2 and their observational parameters.}
{\hline
IAU name & RA (J2000) & DEC (J2000) & LAS & $S_{\rm 1400 MHz} $ \\
& [h m s] & [$^{\circ}$ ' ''] & [arcsec] & [mJy]
\\
\hline
J0245$+$0108 & 02 45 34.07 & $+$01 08 13.7 & 53.4 & 327 \\
J0809$+$2015 & 08 09 20.81 & $+$20 15 38.6 & 56.3 & 84.8 \\
J0809$+$2912 & 08 09 06.22 & $+$29 12 35.5 & 129 & 304 \\
J0812$+$3031 & 08 12 40.08 & $+$30 31 09.4 & 146 & 24.5 \\
J0819$+$0549 & 08 19 41.13 & $+$05 49 42.7 & 115 & 27.4 \\
J0839$+$2928 & 08 39 51.75 & $+$29 28 18.2 & 50.5 & 113 \\
J0842$+$2147 & 08 42 39.96 & $+$21 47 10.4 & 130 & 49.2 \\
J0857$+$0906 & 08 57 48.57 & $+$09 06 48.1 & 59.1 & 130 \\
J0902$+$5707 & 09 02 07.20 & $+$57 07 37.9 & 101 & 29.8 \\
J0906$+$0832 & 09 06 49.99 & $+$08 32 55.9 & 79.8 & 78.7 \\
J0947$+$5154 & 09 47 40.01 & $+$51 54 56.8 & 58.6 & 111 \\
J0952$+$0628 & 09 52 28.46 & $+$06 28 10.5 & 65.0 & 127 \\
J1030$+$5310 & 10 30 50.91 & $+$53 10 28.6 & 100 & 56.4 \\
J1039$+$0714 & 10 39 36.67 & $+$07 14 27.4 & 58.6 & 24.7 \\
J1130$+$3628 & 11 30 26.18 & $+$36 28 36.9 & 51.6 & 45.2 \\
J1207$-$0244 & 12 07 08.02 & $-$02 44 44.2 & 54.6 & 62.1 \\
J1434$-$0123 & 14 34 10.77 & $-$01 23 41.7 & 60.6 & 119 \\
J1550$+$3652 & 15 50 02.01 & $+$36 52 16.8 & 80.4 & 47.4 \\
J1706$+$3214 & 17 06 48.07 & $+$32 14 22.9 & 54.2 & 125 \\
J2345$-$0936 & 23 45 40.45 & $-$09 36 10.2 & 60.9 & 223 \\
\hline
\multicolumn{11}{p{8cm}}{}
}
\Acknow{This project was supported by the MNiSW with funding for scientific research in years
2009--2012 under contract No. 3812/B/H03/2009/36.
Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the
Participating Institutions, the National Science Foundation, the U.S. Department of Energy, the
National Aeronautics and Space Administration, the Japanese Monbukagakusho, the Max Planck
Society, and the Higher Education Funding Council for England. The SDSS Web Site is
http://www.sdss.org/.}
\vspace{2mm}
The authors are grateful to the anonymous refree for
valuable remarks and suggestions that helped to improve the paper.
\vspace{3mm}
|
\section*{The Bibliography}
\bibliographystyle{elsarticle-num}
|
\section{Introduction}
Over the past two years, the DAMA/LIBRA collaboration has repeated their claim to have observed an annual modulation in their event rate, and have interpreted this as evidence of dark matter particles interacting with their detector at a confidence level of 8.2$\sigma$~\cite{dama}. Such an annual modulation in the rate of dark matter scattering events is predicted to arise as a result of the variation in the velocity of the Earth with respect to the Milky Way's dark matter halo. Both the reported amplitude and phase of DAMA's signal are consistent with a dark matter interpretation.
In typical dark matter models capable of producing the signal reported by DAMA, however, one would have expected corresponding signals to appear in other direct detection experiments. To the contrary, several such experiments, including CDMS~\cite{cdms}, CRESST~\cite{cresst}, CoGeNT~\cite{cogent}, XENON~\cite{xenon}, COUPP~\cite{coupp}, ZEPLIN~\cite{zeplin}, Edelweiss~\cite{edelweiss}, and KIMS~\cite{kims}, have reported null results, strongly constraining the dark matter interpretation of DAMA's annual modulation~\cite{compare}. This tension between DAMA and other direct detection experiments has motivated a number of scenarios in which dark matter particles scatter preferentially with the sodium iodide (NaI) detectors of DAMA/LIBRA relative to the materials used in other experiments (such as germanium and silicon used in CDMS). Such scenarios feature dark matter particles which interact with nuclei through a resonance~\cite{resonance}, interact with nuclei with a momentum dependence causing them to scatter more efficiently with NaI than other targets~\cite{ffdm}, or which interact with nuclei largely through inelastic processes~\cite{inelastic}. In particular, dark matter particles which, instead of elastically scattering with nuclei, scatter only through processes in which they are excited to a slightly ($\sim$$100$ keV) heavier state have generated a great deal of interest, and have been shown to be capable of reconciling the claims of the DAMA collaboration with the null results of other direct detection experiments~\cite{inelastic}.
In this article, we consider inelastic dark matter and study the role that such particle would play in compact stars, such as white dwarfs. In the case of dark matter candidates which scatter elastically with nuclei, their capture by and annihilation in stars is expected to provide an observable quantity of energy only in the most extreme environments, such as in a density spike of dark matter in the proximity of the supermassive black hole at the center of the Milky Way~\cite{burners}, or in the early phases of evolution of very massive population III stars~\cite{early}. Inelastic dark matter, in contrast, can potentially scatter with nuclei far more efficiently, but only if in possession of sufficient momentum to overcome the mass splitting between the ground and excited states. Because of the depth of the gravitational wells surrounding white dwarfs, inelastic dark matter particles will be accelerated to high velocities as they fall toward such a star, allowing them to easily accumulate enough momentum to scatter inelastically with the carbon or oxygen nuclei which make up the majority of the star's mass. As a result, inelastic dark matter particles will scatter in and be captured by white dwarf stars at very high rates. In their subsequent annihilations, inelastic dark matter particles can deposit a significant amount of energy into these stars, potentially increasing their luminosity and/or temperature. In particular, old and cool white dwarf stars that reside in regions of high dark matter density (such as in dwarf spheroidal galaxies, or in the inner regions of the Milky Way) may provide a valuable test of the inelastic dark matter scenario.
\section{Capture of Inelastic Dark Matter In White Dwarf Stars}
More than two decades ago, the rate at which elastically scattering dark matter particles are captured in the Sun was calculated~\cite{capture}. In many respects, the capture rate of inelastically scattering dark matter particles is similar. The key difference, however, is that in the inelastic scenario the dark matter-nucleus scattering cross section depends on the velocity of the incoming particle. Above the threshold for inelastic scattering, the cross section is given by
\begin{equation}
\sigma_{\rm inelastic} = \sigma_{\rm elastic} \sqrt{1-\frac{2 \delta}{\mu v^2}},
\end{equation}
where $\delta$ is the mass splitting between the dark matter and its excited state, $\mu$ is the dark matter-nucleus reduced mass, and $v$ is dark matter particle's velocity. The mass splitting, $\delta$, plays an important role in both direct detection experiments, and in the capture of dark matter particles onto the Sun. In particular, dark matter particles with a splitting of $\delta \approx 100$ keV and a mass of 100 GeV must have a velocity of $v\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 575$ km/s if they are to inelastically scatter off of an iodine nucleus ($A=127$). Including the motion of the Solar System through the dark matter halo, a non-negligible fraction of the dark matter particles are expected to exceed this velocity, especially in the summer months, leading to the prediction of a high event rate with strong annual modulation in an experiment such as DAMA. In an experiment such as CDMS, which uses Germanium ($A=73$) and Silicon ($A=28$) target nuclei, however, the velocity threshold is significantly higher ($v\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 665$ km/s), leading to considerably suppressed event rates. This relative suppression of the scattering rate with light target nuclei is the essence of the motivation for inelastic dark matter.
In order for inelastic dark matter to avoid exceeding the limits of CDMS and XENON10, while also generating the annual modulation reported by DAMA, the scatterings must excite the dark matter into a state $\delta\approx 80-150$ keV more massive that the ground state~\cite{cdms,xenoninelastic}. As a result of this mass splitting, inelastic dark matter scattering of off nuclei in the Sun is highly suppressed. In particular, essentially no scattering occurs in the Sun between inelastic dark matter and nuclei lighter than carbon. Although iron nuclei constitute only 0.16\% of the Sun's mass, they dominate the capture rate of inelastic dark matter in the Sun.
This suppression is not found, however, in the case of more dense stars, such as white dwarfs or neutron stars. The velocity of a dark matter particle that falls into a white dwarf from rest will exceed 5000 km/s at the surface, and 8000 km/s near the core (in contrast to approximately 1300 km/s near the core of the Sun). In this situation, $2 \delta/\mu v^2 \ll 1$, and the effects of the dark matter's mass splitting are negligible. For example, an inelastic dark matter particle falling from rest into a white dwarf consisting mostly of carbon or oxygen will inelastically scatter with a cross section within a few percent of the value found in the elastic limit.
To calculate the capture rate of inelastic dark matter particles in a white dwarf, we follow the approach described in Refs.~\cite{weiner,wang} for the case of capture in the Sun. For the Sun, we find results which are in good agreement with these previous studies (see the dashed lines in the right frame of Fig.~\ref{fig}).
White dwarf stars can be well modeled as a system in hydrostatic equilibrium with an equation of state of a cold fermi gas. Following Refs.~\cite{kippenhahn,webpage}, we describe the equation of state of a white dwarf as
\begin{eqnarray}
P&=&C_1 f(x),\nonumber \\
\rho&=&C_2 x^3, \nonumber \\
x&=&k_f/(m_e c),
\end{eqnarray}
where $P$ is the pressure, $\rho$ is the density of the star, and $k_f$ is the Fermi momentum. The numerical coefficients and the function $f(x)$ are given by
\begin{eqnarray}
C_1&=&\frac{\pi m_e^4 c^5}{3h^3}, \nonumber \\
C_2&=&\frac{ 8\pi m^3_e c^3}{3h^3},\nonumber \\
f(x)&=&x(2x^2-3)(x^2+1)^{1/2}+3 \ln [x+(1+x^2)^{1/2}], \,\,\,\,\,\,\,\,\,
\end{eqnarray}
where $h$ is the Planck constant. Substituting into the above equations for pressure and density, the equation of hydrostatic equilibrium becomes
\begin{equation}
\frac{C_1}{C_2} \frac{1}{r^2} \frac{d}{dr} \bigg(\frac{r^2}{x^3}\frac{df(x)}{dr}\bigg)=-4\pi G C_2 x^3,
\end{equation}
which we solve numerically using a publicly available code~\cite{webpage} that also includes corrections due to electrostatic~\cite{Hamada} and general relativistic effects. This approach gives an accurate description of the stellar structure of white dwarfs and fits the data exquisitely (see, for example, Fig.~3 of Ref.~\cite{webpage}).
One way in which the calculation of inelastic dark matter capture by white dwarfs departs from that described in Refs.~\cite{weiner,wang} is in the treatment of the star's optical depth. In particular, for cross sections as large as are required to produce the DAMA annual modulation ($\sigma$\,$\sim$\,$10^{-40}$ cm$^2$), dark matter particles are generally expected to scatter many (typically several tens of) times as they pass through a white dwarf. In this optically thick limit, the rate at which dark matter particles are captured simplifies considerably to~\cite{thick,bertone}
\begin{equation}
\Gamma_c \approx \bigg(\frac{8 \pi}{3}\bigg)^{1/2} \, \frac{3G \, R_{\rm WD}\, M_{\rm WD}\, \rho_{\rm DM}}{m_{\rm DM}\, \bar{v}},
\end{equation}
where $R_{\rm WH}$ and $M_{\rm WD}$ are the radius and mass of the white dwarf, respectively, $\rho_{\rm DM}$ is the density of dark matter in the star's environment, and $\bar{v}$ is the dark matter's velocity dispersion (prior to infall into the gravitational potential of the star). Throughout our study, we will adopt $M_{\rm WD}=0.7 M_{\odot}$ and $R_{\rm WD}=0.0093 R_{\odot}$ as our default values (the mass distribution of white dwarfs is strongly peaked near 0.5-0.7 $M_{\odot}$~\cite{mass}). Note that as long as we remain within the optically thick limit, the capture rate does not depend on the interaction cross section of the dark matter, but only on its number density and velocity distribution.
\begin{figure*}[!]
\begin{center}
{\includegraphics[angle=0,width=0.49\linewidth]{P_of_rho_v20.eps}}
\hspace{0.1cm}
{\includegraphics[angle=0,width=0.49\linewidth]{P_of_rho_v250.eps}}\\
\vspace{-0.3cm}
\caption{The rate at which energy is injected into white dwarf stars (solid) and into the Sun (dashed) by annihilating inelastic dark matter, as a function of the environment's dark matter density. In the left frame, we show results for a dark matter velocity dispersion of 20 km/s (appropriate for globular clusters~\cite{clusteresc}), whereas in the right frame we use a velocity dispersion of 250 km/s and a star's velocity of 220 km/s (approprite for the Solar System). In each frame, upper (black) and lower (red) curves use dark matter-nucleon scattering cross sections of $10^{-40}$ cm$^2$ and $10^{-44}$ cm$^2$, respectively ($\sigma \approx 10^{-40}$ cm$^2$ is required to generate the DAMA annual modulation in the inelastic dark matter scenario, whereas $\sigma \approx 10^{-44}$ cm$^2$ is approximately the maximum value allowed for elastically scattering dark matter candidates). In each case, we have adopted a dark matter mass of 100 GeV, and a mass splitting of $\delta=100$ keV, although our results are not strongly sensitive to these values. The horizontal dotted lines denote the energy deposition rate above which observations of cool (3000-4000 K) white dwarfs would conflict with the predictions of inelastic dark matter.}
\label{fig}
\end{center}
\end{figure*}
The large cross section required in the inelastic dark matter scenario quickly leads to an equilibrium between the rates of capture and annihilation in a white dwarf, assuming that the dark matter particles are capable of self-annihilation (which might not be true if the dark matter carries a conserved quantum number rather being stabilized by a parity, for example). As a result, the capture rate can be converted into the rate at which additional energy is injected into the star's core. In a typical white dwarf star ($R_{\rm WD} \approx 0.0093 R_{\odot}$ and $M_{\rm WD}\approx 0.7 M_{\odot}$), inelastic dark matter annihilations provide a contribution to the star's luminosity that is given by:
\begin{eqnarray}
L &\approx& \Gamma_c \, m_{\rm DM} \nonumber \\
&\approx& 2.5\times 10^{28} \,{\rm GeV/s} \,\bigg(\frac{\rho_{\rm DM}}{1 \,{\rm GeV/cm}^3}\bigg) \bigg(\frac{220 \, {\rm km/s}}{\bar{v}}\bigg)\nonumber \\
&\approx& 4\times 10^{25} \,{\rm erg/s} \,\bigg(\frac{\rho_{\rm DM}}{1 \,{\rm GeV/cm}^3}\bigg) \bigg(\frac{220 \, {\rm km/s}}{\bar{v}}\bigg).
\label{power}
\end{eqnarray}
In Fig.~\ref{fig}, we show the rate at which inelastic dark matter deposits energy into a white dwarf (or into the Sun), as a function of the environment's dark matter density. We show results for dark matter with velocity dispersions of 20 km/s (left) and 250 km/s (right), and use two values for the dark matter-nucleon cross section: $10^{-44}$ cm$^2$ (lower red curves) and $10^{-40}$ cm$^2$ (upper black curves). In the case of the larger (smaller) cross section choice, a white dwarf is within the optically thick (thin) limit. As a cross section of $\sigma \approx 10^{-40}$ cm$^2$ is required to generate the DAMA modulation, we will focus on the optically thick case in the remainder of the study.
The coldest and least luminous observed white dwarfs have temperatures as low as 3000 K and luminosities of $(2-3) \times 10^{28}$ erg/s (following from the Stefan-Boltzmann law, a 3000 K blackbody of radius 0.01 $R_{\odot}$ is predicted to have a luminosity of $2.8\times 10^{28}$ erg/s). From these observations, we can conclude that any additional source of energy (such as dark matter annihilations) must not exceed approximately $2\times 10^{28}$ erg/s. Within the context of inelastic dark matter, this translates to an approximate limit of $\rho_{\rm DM} \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 500$ GeV/cm$^3$ ($\rho_{\rm DM} \mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 40$ GeV/cm$^3$) for $\bar{v}=250$ km/s ($\bar{v}=20$ km/s). Although these density limits are orders of magnitude higher than the average dark matter density in the local region of the Milky Way ($\rho_{\rm DM,local}\approx$ 0.3-0.5 GeV/cm$^3$~\cite{ullio}), there exist regions that are thought to contain sufficiently high quantities of dark matter (and/or with sufficiently slow velocity distributions) to make white dwarf stars useful probes of inelastic dark matter. We discuss some of these possibilities in the following section.
\section{Dark Matter Densities In Regions Containing White Dwarf Stars}
Many of the oldest (and coldest) known white dwarf stars are located in the dense collections of stars known as globular clusters. If high densities of dark matter were also present in such systems, these very old and cool white dwarfs would provide an ideal environment to search for the effects of inelastic dark matter. Despite the very high densities of baryonic matter in globular clusters (often containing hundreds of thousands of stars within a volume smaller than a cubic parsec), however, there is no compelling dynamical evidence that these systems contain significant quantities of dark matter.
Within the modern paradigm of globular cluster formation, these objects form in the disk of galaxies rather than in the center of dark matter halos (in contrast to dwarf spheroidal galaxies, for example). Indeed, observations of the Antennae Galaxy dramatically demonstrate the formation of globular clusters within its disk. Regardless, the formation of a globular cluster is generally expected to enhance the dark matter density through adiabatic contraction. In particular, the orbits of dark matter particles that are gravitationally bound to the cluster become contracted as the baryons ({\it ie.}~the stars) themselves gather more densely. This can lead to a contracted dark matter density in the globular cluster's core approximately given by
\begin{equation}
\rho_{\rm DM,core} \approx \rho_{\rm b,core}\, \bigg(\frac{\rho_{\rm DM,i}}{\rho_{b,i}}\bigg) \bigg(\frac{\rho_{\rm ISM}}{\rho_{\rm core}}\bigg)^{1/4},
\end{equation}
where $\rho_{\rm b,core}$ is the density of baryons in the globular cluster's core, $\rho_{\rm DM,i}$ is the initial density of {\it gravitationally bound} dark matter particles prior to adiabatic contraction, $\rho_{b,i}$ is the initial density of baryons prior to adiabatic contraction, $\rho_{\rm ISM}$ is the total density of the interstellar medium, and $\rho_{\rm core}$ is the total density in the core of the cluster. This result was found analytically using the standard adiabatic contraction model of Blumenthal {\it et al.}~\cite{blumenthal}, and was verified numerically using the adiabatic contruction code, \emph{Contra}~\cite{contra}.
Considering as an example the nearby globular cluster M4, we would expect that within an initial distribution of dark matter with a velocity dispersion of $\bar{v}\sim 220$ km/s, only $\sim 10^{-3}-10^{-4}$ or less of the particles will be gravitationally bound, thus significantly limiting the impact of adiabatic contraction. After adiabatic contraction, we estimate that the dark matter density in the core of M4 could be enhanced to on the order of a few GeV/cm$^3$ which, although an order of magnitude greater than prior to the formation of the cluster, is well below the density needed to significantly effect white dwarf stars.
As an alternative to globular clusters, we also consider white dwarfs within the dark matter halos of dwarf spheroidal galaxies. Dwarf spheroidals are highly dense, dark matter dominated systems, and contain very old stars~\cite{old}, making them excellent environments in which to study the impact of inelastic dark matter on white dwarfs. Ref.~\cite{list}, for example, describes seven dwarf spheroidals (Carina, Draco, Fornax, Leo I, Leo II, Sculptor, and Sextans) which have dark matter densities of 40-150 GeV/cm$^3$ and velocity dispersions of $\sim$10-20 km/s within their inner 10-20 parsecs. If old and cold white dwarf stars are present in such environments, we predict that inelastic dark matter will inject energy into them at a rate rivaling their luminosity. As a result, the observation of a white dwarf in such an environment with a temperature below approximately 4000 K would conflict with the inelastic dark matter hypothesis. Although dwarf spheroidals are typically very distant, and thus challenging to observe, there are 3-4 known dwarf spheroidals (Segue I, Segue II, Ursa Major II, and possibly Wilman I) within $\sim$30 kpc of the Solar System, making them within the approximate range of a 30 meter telescope~\cite{30m}. In the more distant future, a 100 meter telescope could place white dwarf stars in all Milky Way dwarf spheroidals within reach.
Lastly, the dark matter density is generally predicted to increase as one approaches the center of the Milky Way. According to the frequently used Navarro-Frenk-White (NFW) profile~\cite{nfw}, for example, the smooth component of the dark matter density in the Milky Way can be parameterized by
\begin{equation}
\rho_{\rm DM} \propto \frac{1}{r [1+(r/R_s)]^2},
\end{equation}
where $r$ is the distance to the Galactic Center, and $R_s\approx 20$ kpc is the scale radius. In addition, the velocity dispersion is predicted to decrease according to $\sigma_v \propto (r/R_s)^{1/2}$. Combining these effects, we estimate that the rate at which inelastic dark matter will accumulate in a white dwarf located at 1 kpc (100 pc) from the Galactic Center is $\sim$$40$ ($\sim$$1250$) times greater than the rate predicted in the vicinity of the Solar System. Adopting an NFW profile, we expect a sizable fraction of the luminosity of cold white dwarf stars within the inner $\sim$100 pc of the Milky Way to originate from inelastic dark matter annihilations. This effect can be considerably enhanced if the Milky Way's dark matter halo were adiabatically contracted by baryons~\cite{ac}. We therefore expect future infrared surveys of the inner Milky Way to provide a valuable test of inelastic dark matter.
\section{Conclusions}
In this paper, we have studied the capture and subsequent annihilation of inelastic dark matter particles in white dwarf stars. We find that inelastic dark matter (as motivated to generate the annual modulation claimed by DAMA/LIBRA without exceeding the limits from other direct detection experiments) is predicted to be captured by compact stars at a rate multiple orders of magnitude higher than elastically scattering dark matter. This makes inelastic dark matter a very efficient source of energy for white dwarf stars, potentially leading to observable consequences.
The role of inelastic dark matter is most pronounced in those stars that reside in regions with a high density (and/or low velocity dispersion) of dark matter. In particular, we find that very old white dwarf stars in the inner tens of parsecs of dwarf spheroidal galaxies will be powered primarily by inelastic dark matter annihilations, preventing their otherwise expected cooling. This is also predicted to be the case for old white dwarfs found within the inner kiloparsec of the Milky Way. If future surveys were to discover very cold (T $\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}}$ 4000 K) white dwarfs in environments such as these, it would disfavor the inelastic dark matter hypothesis. In contrast, the lack of such stars discovered in future observations could be used to bolster the case for inelastic dark matter.
\bigskip
{\it Acknowledgements:} While we were in the final stages of this project, Ref.~\cite{fair} appeared on the LANL archive. While our numerical results are in good agreement, we have chosen to focus our study on inelastic dark matter captured by white dwarfs in dwarf spheroidal galaxies, and in the inner Milky Way, rather than in globular clusters which we have argued are unlikely to contain high densities of dark matter. The authors are supported by the US Department of Energy, including grant DE-FG02-95ER40896. DH is also supported by NASA grant NAG5-10842.%
|
\section{Introduction\label{intro}}
The integer quantized Hall effect (IQHE), observed at two
dimensional charge systems (2DCS) subject to strong perpendicular
magnetic fields $B$, is usually discussed within the single
particle picture, which relies on the fact that the system is
highly disordered~\cite{Kramer03:172,Schweitzer85:379}. These quantized
(spinnless) single particle energy levels are called the Landau
levels (LLs) and the discrete energy values are given by
$E_N=\hbar\omega_c(n+1/2)$, where $n$ is the Landau index and
$\omega_c=eB/m^*c$ is the cyclotron frequency of an electron with
an effective mass $m^*$ ($\approx 0.067m_e$, $m_e$ being the
bare electron mass at rest) and $c$ is the speed of light in
vacuum. In single particle models the disorder plays several
roles, such as Landau level broadening~\cite{Cai86:3967}, leading
to a finite longitudinal
conductivity~\cite{Ando74:959,Ando75:279}, spatial
localization~\cite{Nixon:90} \emph{etc}. Disorder can be created
by inhomogeneous distribution of dopant ions which essentially
generates the confinement potential~\cite{stopadisorder:96} for
the electrons. In the absence of disorder, the
density of states are Dirac delta-functions
$D(E)=\frac{1}{2 \pi
l^2}{\sum^{\infty}_{N=0}}{\delta(E-E_N)}$, where
$l=\sqrt{\hbar/eB}$ is the magnetic length, and the longitudinal
conductivity ($\sigma_{l}$) vanishes. For a homogeneous two dimensional electron system (2DES), by
the inclusion of disorder and due to collisions, LLs become
broadened. Therefore the longitudinal conductance becomes non-zero
in a finite energy (in fact magnetic field) interval. Long range
potential fluctuations generated by the disorder result in the so
called {\em classical localization}~\cite{Fogler94:1656},
\emph{i.e.} the guiding center of the cyclotron orbit moves along
closed equi--potentials~\cite{Efros88:1019}. In contrast to the
above mentioned bulk theories, the edge theories usually disregard
the effect of disorder to explain the (quantized) Hall resistance
$R_{\rm H}$ and accompanying (zero) longitudinal resistance
$R_{\rm L}$. However, the non-interacting edge theories still
require disorder to provide a reasonable description of the
transition between the plateaus. The Landauer-B\"uttiker approach
(known as the edge channel picture) \cite{Buettiker86:1761} and
its direct Coulomb interaction generalized version, \emph{i.e.}
the non-self-consistent Chklovskii
picture~\cite{Chklovskii92:4026}, also needs localization
assumptions in order to obtain quantized Hall (QH) plateaus of
finite width (see for a review \emph{e.g.} Datta's book
\cite{Datta} and Ref.~\cite{Efros88:1019} for the estimates
of plateau widths at the high disorder limit).
In contrast to above discussions very recent
experimental~\cite{Wilde06:disorder,Mares:09,josePHYSE,jose:prl}
and theoretical~\cite{dassarma:mobility,Macleod09:background}
results point the incomplete treatment of the disorder potential
and scattering mechanisms. Fairly recent theoretical
approaches~\cite{Guven03:115327}, the QH plateaus are obtained by
the inclusion of direct Coulomb interaction
self-consistently~\cite{siddiki2004} and the effect of the
disorder was handled via conductivity tensor
elements~\cite{Bilayersiddiki06:}, however, the source of the
disorder and its properties was left
unresolved~\cite{Siddiki04:condmat}. Whereas, the influence of
potential fluctuations on the QH plateaus were discussed
briefly~\cite{Gerhardts08:378,Siddiki:ijmp}.
This work provides a systematic investigation of the disorder
potential and its influence on the quantized Hall effect including
direct Coulomb interaction. The investigation is extended to realistic experimental
conditions in determining the widths of the quantized Hall
plateaus. We, essentially study the effect of disorder in two
distinct regimes, namely the short range and the long range. The
short range part is included to the density of states (DOS),
thereby influences the widths of the current carrying edge-states
and the entries of the conductivity tensor. Whereas, the long
range part is incorporated to the self-consistent calculations,
which determines the extend of the plateaus in turn. In
Sec~\ref{impurity} we introduce two types of single
impurity potentials, namely the Coulomb and the Gaussian, and
compare their range dependencies considering damping of the
dielectric material. In the next step we
discuss the \emph{screened} disorder potential within a pure electrostatic approach, by considering an
homogeneous two dimensional electron system (2DES) without an
external magnetic field and show that the long range part is well
screened, whereas the short range part is almost unaffected. Section~\ref{treed} is devoted to investigate
impurities numerically, where we solve the Poisson equation
self-consistently in three dimensions. The numerical and analytical calculations
are compared, considering the estimations of the disorder
potential range and its variation amplitude. We finalize our
discussion with Sec.~\ref{quanHP}, where we calculate the plateau
widths under experimental conditions for different sample widths
and mobilities.
\section{Impurity potential\label{impurity}}
The disorder potential experienced by the 2DES, resulting from the
impurities has quite complicated range dependencies. Since, the
potential generated by an impurity is (i) \emph{damped} by the
dielectric material in between the impurity and the plane where
the 2DES resides (ii) is screened by the homogeneous 2DES
depending on the density of states, which changes drastically with
and without magnetic field. It is common to theoreticians to
calculate the conductivities from single impurity potentials, such
as Lorentzian~\cite{Guven03:115327}, Gaussian~\cite{Ando82:437} or
any other analytical functions~\cite{Champel08:124302,TobiasK06:h}. However, the landscape of potential fluctuations
is also important to define the actual mobility of the sample at
hand, in particular in the presence of an external magnetic field.
\subsection{Pure Electrostatics}
We first discuss the different range dependencies
of the Coulomb and Gaussian donors, assuming open boundary conditions. Next, the effect of
the spacer thickness on the disorder potential is discussed, namely
the damping of the external (Coulomb) potential, and is compared
with the Thomas-Fermi screening. The different damping/screening
dependencies of the resulting potentials are discussed in terms of
range.
The Coulomb potential presents long range part, which leads to long range fluctuations due to
overlapping if several donors are considered. Whereas, the Gaussian potential decays exponentially on the length
scale comparable with the separation thickness. Since the Gaussian potential is relatively
short ranged, no overlapping of the
single donor potentials occur. Hence, the external potential
experienced by the electrons can be approximated to a homogeneous
potential fairly good. Thus one can conclude that approximating the
total disorder potential by Gaussians is not sufficient to
recover the long range part. Similar
arguments are also found in the
literature~\cite{Nixon:90,Efros88:1019,Siddiki:ijmp}. In order to
overcome the difference observed at the long range potential
fluctuations between the Coulomb and the Gaussian impurities, the
following procedure is applied: First we calculate the total disorder potential considering many impurities then we perform a two-dimensional
Fourier transformation of the Coulomb potential and make a back
transformation keeping the first few momentum $q$ components in
each direction, hence only the long range part of the potential is
left~\cite{Siddiki:ijmp}. Then we add the long range part of the Coulomb potential to
the potential created by donors, \emph{i.e.} to the confinement
potential. We take this as a motivation to simulate the short
range part of the impurity potential by Gaussian impurities, and
calculate the Landau level broadening and the conductivities,
described within the self-consistent Born
approximation (SCBA)~\cite{Ando82:437}.
Here we point to the effect of the spacer thickness on the impurity
potential experienced in the plane of 2DES. It is well known from
experimental and theoretical investigations that, if the distance
between the electrons and donors is large, the mobility is
relatively high and it is usually related with suppression of the
short range fluctuations of the disorder potential. These
results agree with the experimental observations of high mobility
samples and are easy to understand from the $z$ dependence of the
Fourier expansion of the Coulomb potential, \begin{equation}V^{
\vec{q}}(z)=\int d \vec{r} e^{-i \vec{q} \cdot \vec{r}}
\sum_{j}^{N} \frac{e^{2}/ \bar{\kappa}}{ \sqrt{( \vec{r}-
\vec{r_{j}})^{2}+z^{2}}}=\frac{2 \pi
e^{2}}{\bar{\kappa}q}e^{-|qz|}NS(\vec{q}), \label{fourier1}\end{equation}where $S(\vec{q})$ contains all the information about the in-plane
donor distribution and $N$ is the total number of the ionized
donors. We observe that if the spacer thickness is increased, the
amplitude of the potential decreases rapidly. We also see that the
short range potential fluctuations, which correspond to higher
order Fourier components, are suppressed more efficiently.
Next, we discuss electronic screening of the external potential
created by the donors discussed above. For a dielectric material
the relation between the external and the screened potentials are
given by, \begin{equation}V_{\rm scr}^q=V_{\rm ext}^q/\epsilon(q),
\label{linear-eps}\end{equation}where $\epsilon(q)$ is the dielectric
\emph{function} and is given by $\epsilon(q)=1+\frac{2\pi
e^2D_0}{\bar{\kappa}|q|}, $ with the constant 2D density of
states $D_0=m/(\pi \hbar^2)$ in the absence of an external $B$
field, and is known as the Thomas-Fermi (TF) function.
The simple linear relation above, together with the TF dielectric
function essentially describes the electronic screening of the
Coulomb potential given in Eq.~\ref{fourier1}, if there are
sufficient number of electrons~\cite{Efros88:1019} ($n_{\rm
el}>0.1\cdot10^{15}$ m$^{-2}$). Consider a case where the $q$
component approaches to zero, then the external (damped) potential
is well screened, hence the long range part of the disorder
potential. Whereas, the short range part remain unaffected,
\emph{i.e.} high $q$ Fourier components. Now we turn our attention
to the second type of impurities considered, the Gaussian ones. As
well known, the Fourier transform of a Gaussian is also of the
form of a Gaussian, therefore, similar arguments also hold for
this kind of impurity.
We should emphasize once more the clear
distinction between the effect of the spacer on the external
potential and the screening by the 2DES, \emph{i.e.} via
$\epsilon(q)$. The former depends on the Fourier transform of the
Coulomb potential and the important effect is the different decays
of the different Fourier components (see Eq.~\ref{fourier1}), so
that the short range part of the disorder potential is well
dampened, whereas the latter depends on the relevant DOS of the
2DES and the screening is more effective for the long range part.
\begin{figure}[!t]\center{
\includegraphics[width=0.55\linewidth]{fig1.pdf}
\vspace{-7cm}
\caption{Schematic representation of the crystal, which we investigate numerically.
The crystal is grown on a thick GaAs substrate, where the 2DES is
formed at the interface of the AlGaAs/GaAs hetero-junction.
The top AlGaAs layer is doped with Silicon 30 nm above the
interface. The crystal is spanned by a 3D matrix
($128\times128\times60$). \label{fig:5}}}
\end{figure}
We continue our investigation by solving the 3D Poisson equation
iteratively for randomly distributed single impurities, where
three descriptive parameters (\emph{i.e.} the number of
impurities, the amplitude of the impurity potential and the
separation thickness) are analyzed separately. Next, we discuss the
long range parts of the potential fluctuations investigating the
Coulomb interaction of the 2DES, numerically. The range is estimated from these investigations by performing Fourier
analysis and is related to the samples used in
experiments~\cite{josePHYSE,jose:prl} (Sec.~\ref{exper}).
\subsection{3D simulations\label{treed}}
In the previous section we took a rather simple way to study the
effect of interactions by assuming an homogeneous 2DES and
screening is handled by the TF dielectric function. Here, we
present our results obtained from a rather complicated numerical
method. We solve the Poisson equation in 3D starting from the
material properties of the wafer at hand, the typical material we
consider is sketched in Fig.~\ref{fig:5}. Namely, using the growth
parameters, we construct a 3D lattice where the potential and the
charge distributions are obtained iteratively assuming open
boundary conditions, \emph{i.e.} $V(x\rightarrow \pm\infty,
y\rightarrow\pm\infty, z\rightarrow\pm\infty)=0$. For such
boundary conditions, we chose a lattice size which is considerably
larger than the region that we are interested in. We preserve the
above conditions within a good numerical accuracy (absolute error
of $10^{-6}$). A forth order grid
approach~\cite{Weichselbaum03:056707} is used to reduce the
computational time, which is successfully used to describe similar
structures~\cite{Sefa08:prb}.
Figure~\ref{fig:5} presents the schematic drawing of the
hetero-structure which we are interested in. The donor layer is
$\delta-$ doped by a density of $3.3\times10^{16}$ m$^{-2}$
(ionized) Silicon atoms, $\sim 30$ nm above the 2DES, which
provide electrons both for the potential well at the interface and
the surface. It is worthwhile to note that most of the electrons
($\sim \%90$) escape to the surface to pin the Fermi energy to the
mid-gap of the GaAs. In any case, for such wafer parameters
there are sufficient number of electrons ($n_{\rm el}\gtrsim 3.0
\times10^{15}$ m$^{-2}$) at the quantum well to form a 2DES. To
investigate the effect of impurities we place positively charged
ions at the layer where donors reside. From Eq.~\ref{fourier1} we
estimate the amplitude of the potential of a single impurity to be
$\frac{e^2}{\kappa}\frac{V_{\rm imp}}{z_D}=0.033$ eV and assume
that \emph{some} percent of the ionized donors are generating the
disorder potential, that defines the long range fluctuations. In
our simulations we perform calculations for a unit cell with
areal size of $1.5~\mu$m$\times 1.5~\mu$m which contains $3.3
\times10^{16}$ donors per square meters, thus with 10 percent
disorder we should have $N_I\sim$ 3300 impurities.
\begin{figure}[!t]\centering
\begin{minipage}{1\linewidth}
\includegraphics[width=0.5\linewidth,angle=0]{fig2a.pdf}
\includegraphics[width=0.5\linewidth,angle=0]{fig2b.pdf}
\end{minipage}
\caption{(a) Electron density fluctuation considering 3300
impurities 30 nm above the electron gas. (b) The long-range part,
arrows are to guide the distance between two maxima. The
calculation is repeated for 50 random distributions, which lead to
a similar range.\label{fig:7}}
\end{figure}
Figure~\ref{fig:7}a depicts the actual density distribution,
when considering 3300 impurities, whereas Fig.~\ref{fig:7}b
presents only the long range part of the density fluctuation. The
arrows show the average distance between two maxima, which is
calculated approximately to be $550$ nm. To estimate an average
range of the disorder potential, we repeated calculations for such
randomly distributed impurities, where number of repetitions
scales with $\sqrt{N_I}$. Such a statistical investigation,
sufficiently ensembles the system to provide a reasonable
estimation of the long range fluctuations. We also tested for
larger number of random distributions, however, the estimation
deviated less than tens of nanometers. We show our main result of
this section in Fig.~\ref{fig:8}, where we plot the estimated long
range part of the disorder potential considering various number of
impurities $N_I$ and impurity potential amplitude $V_{\rm imp}$.
Our first observation is that the long range part of the total
potential becomes less when $N_I$ becomes large, not surprisingly.
However, the range increases nonlinearly while decreasing $N_I$,
obeying almost an inverse square law and tend to saturate at
highly disordered system. When fixing the distributions and $N_I$,
and changing the amplitude of the impurity potential we observe
that for large amplitudes the range can differ as large as 200 nm
at all impurity densities. We found that for impurity
concentration less than $\% 3$, the range of the potential is
larger than the unit cell we consider, \emph{i.e} $R>1.5 \mu$m. In
contrast to the long range part, the short range part is almost
unaffected by the impurity concentration, however, is affected by
the amplitude. Therefore, while defining the conductivities we
will focus our investigation on $V_{\rm imp}$. Another important
result is that the estimates of long range fluctuations does not
depend strongly on the spacer thickness, if one keeps the
amplitude of single impurity potential amplitude fixed,
Fig.~\ref{fig:8}b. All of the above numerical observations
coincide fairly good with our analytical investigations in the
previous section. However, the range dependency on the impurity
concentration cannot be estimated with the analytical formulas
given. We should also note that, similar or even complicated
numerical calculations are present in the
literature~\cite{Nixon:90,stopadisorder:96}. A indirect measure of
the screening effects on the potential can also be inferred by
capacitance measurements, supported by the above calculation
scheme in the presence of external field~\cite{Mares:09}.
\begin{figure}[!t]\centering
\begin{minipage}{1\linewidth}
\includegraphics[width=.5\linewidth,angle=0]{fig3a.pdf}
\includegraphics[width=.5\linewidth,angle=0]{fig3b.pdf}
\end{minipage}
\caption{Statistically estimated range of the density fluctuations
as a function of number of impurities, considering various
impurity strengths (a) and spacer thicknesses (b). The
calculations are done at zero temperature considering Coulomb
impurities. The long range potential fluctuations become larger
than the size of the unit cell if one considers less than $\%$5
disorder. \label{fig:8}}
\end{figure}
Next section is devoted to investigate the widths of the quantized
Hall plateaus utilizing our findings. We consider mainly two ``mobility'' regimes, where
the long range fluctuations is at the order of microns (high
mobility) and is at the order of few hundred nanometers, low
mobility. However, the amplitude of the total potential
fluctuations will be estimated not only depending on the number of
impurities but also depending on the spacer thickness, range and
amplitude of single impurity potential.
\section{Quantized Hall plateaus\label{quanHP}}
The main aim of this section is to provide a systematic
investigation of the quantized Hall plateau (QHP) widths within
the screening theory of the IQHE~\cite{siddiki2004}, therefore
here we summarize the essential findings of the mentioned theory.
In calculating the QHPs one needs to know local conductivities,
namely the longitudinal $\sigma_{\rm l}(x,y)$ and the transverse
$\sigma_{\rm H}(x,y)$. To determine these quantities it is
required to relate the electron density distribution $n_{\rm
el}(x,y)$ to the local conductivities explicitly. Here we utilize
the SCBA~\cite{Ando82:437}. However, the calculation of the
electron density and the potential distribution including direct
Coulomb interaction is not straightforward, one has to solve the
Schr\"odinger and the Poisson equations simultaneously. This is
done within the Thomas-Fermi approximation which provides the
following prescription to calculate the electron density \begin{equation}n_{\rm el}(x,y)=\int dE D(E)\frac{1}{e^{(E_F-V(x,y))/k_BT}+1},
\label{density}\end{equation}where $D(E)$ is the appropriate density of
states calculated within the SCBA, where $k_B$ is the Boltzmann
constant and $T$ temperature. The total potential is obtained from
\begin{equation}V(x,y)=\frac{2e^2}{\bar{\kappa}}\int dxdy K(x,y,x',y')n_{\rm
el}(x,y),\label{potential}\end{equation}and the Kernel $K(x,y,x',y')$ is the
solution of the Poisson equation satisfying the boundary
conditions to be discussed next.
In the following we assume a translation in variance in
$y$-direction and implement the boundary conditions $V(-d)=V(d)=0$
($2d$ being the sample width), proposed by Chklovskii
\emph{et.al.}~\cite{Chklovskii92:4026}, such a geometry allows us
to calculate the Kernel in a closed form. Hence,
Eqs.~(\ref{density}) and~(\ref{potential}) forms the self-consistency. For a given initial potential distribution, the electron
concentration can be calculated at finite temperature and magnetic
field, where the density of states $D(E)$ contains the information
about the quantizing magnetic field and the effect of short range
impurities. Here we implicitly assume that the electrons reside in
the interval $-b<x<b$ (where, $d_l=|d-b|/d$ is called the
depletion length), and is fixed by the Fermi energy, \emph{i.e.}
the number of electrons, hence donors. As a direct consequence of Landau
quantization and the locally varying electrostatic potential, the
electronic system is separated into two distinct regions, when
solving the above self-consistent equations iteratively: i) The
Fermi energy equals to (spin degenerate) Landau energy and due
to DOS the system illustrates a metallic behavior, the
compressible region, ii) The insulator like incompressible region,
where $E_F$ falls in between two consequent eigen-energies and no
states are available~\cite{Chklovskii92:4026,Siddiki03:125315}. It
is usual to define the filling factor $\nu$, to express the
electron density in terms of the applied $B$ field as, $\nu=2\pi
l^2 n_{\rm el}$. Since all the states below the Fermi energy are
occupied the filling factor of the incompressible regions
correspond to integer values (\emph{e.g.} $\nu=2,4,6...$), whereas
the compressible regions have non-integer values, due to partially
occupied higher most Landau level. The spatial distribution and
widths of these regions are determined by the confinement
potential~\cite{Chklovskii92:4026}, magnetic
field~\cite{Lier94:7757}, temperature~\cite{Oh97:13519} and
level broadening~\cite{Guven03:115327,siddiki2004}. For the purpose of the
present work we fix the confinement potential profile by
confining ourselves to the Chklovskii geometry and keeping the
donor concentration (and distribution) constant. Moreover we
perform our calculations at a default temperature given by
$k_BT/E_F^0=0.02$, where $E_F^0$ is the Fermi energy calculated
for the electron concentration at the center of the sample and is typically similar to 10 meV.
The next step is to calculate the global resistances, \emph{i.e.}
the longitudinal $R_{\rm L}$ and Hall $R_{\rm H}$ resistances,
starting from the local conductivity tensor elements. Such a
calculation is done within a relaxed local model that relates the
current densities $\textbf{j}(x,y)$ to the electric fields
$\textbf{E}(x,y)$, namely the local Ohm's law: \begin{equation}\textbf{j}(x,y)=\hat{\sigma}(x,y)\textbf{E}(x,y). \end{equation}The strict
locality of the conductivity model is lifted by an spatial
averaging process~\cite{siddiki2004} over the quantum mechanical
length scales and an averaged conductivity tensor
$\hat{\overline{\sigma}}(x,y)$ is used to obtain the global
resistances. It should be emphasized that, such an averaging
process also simulates the quantum mechanical effects on the
electrostatic quantities. To be explicit: if the widths of the
current carrying incompressible strips become narrower than the
extend of the wave functions, these strips become ``leaky'' which
can not decouple the two sides of the Hall bar and back-scattering
takes place. Therefore, to simulate the ``leakiness" of the
incompressible strips we perform coarse-graining over quantum
mechanical length scales.
\begin{figure}[!t]\center{
\includegraphics[width=.9\linewidth,angle=0]{fig4.pdf}
\caption{The Hall resistances versus magnetic field, calculated at
default temperature and considering a 10$~\mu$m sample for
different ranges of the single impurity potential. Inset depicts a
larger $B$ field interval, where $\nu=4$ plateau can also be
observed. \label{fig:9}}}
\end{figure}
Now let us relate the local conductivities with the local filling
factors. Since the compressible regions behave like a metal within
these regions there is finite scattering leading to finite
conductivity. In contrast, within the incompressible regions the
back-scattering is absent, hence, the longitudinal conductivity
(and simultaneously resistivity) vanishes. Therefore, all the
imposed current is confined to these regions. The Hall
conductivity, meanwhile is just proportional to the local electron
density. The explicit forms of the conductivity tensor elements
are presented elsewhere~\cite{siddiki2004}. Having the electron
density and local magneto-transport coefficients at hand, we
perform calculations to obtain the widths of the quantized Hall
plateaus utilizing the above described, microscopic model assisted
by the local Ohm's law at a fixed external current $I$. Further
details of the calculation scheme is reviewed in
Ref.~\cite{Gerhardts08:378}.
\subsection{Single impurity potentials: Level broadening and conductivities \label{singleimp}}
Since the very early days of the charge transport theory,
collisions played an important role. Such a scattering based definition of conduction also
applies for the system at hand, \emph{i.e.} a two-dimensional
electron gas subject to perpendicular magnetic field. Among many
other approaches~\cite{Gerhardts75:285,Guven03:115327,TobiasK06:h} the SCBA emerged as a
reasonable model to describe the DOS assuming Gaussian impurities,
considering short range scattering. A single impurity has two distinct
parameters that represents the properties of the resulting
potential, the range $R_{\rm g}$ (at the order of separation
thickness) and the amplitude of the potential (in relevant units),
$\widetilde{V}_{\rm imp}$. However, these two parameters are not
enough to define the widths of the Landau levels ($\Gamma$), another
important parameter is the number of the impurities, $N_I$. In the
previous section we have already investigated these three
parameters in scope of potential landscape, now we utilize our
findings to define the level widths and the conductivities. It is
more convenient to write the single impurity potential of the
form,
\begin{table}\center{
\begin{tabular}{|c|c|c|c|c|}
\hline
$d_l=$ 70 nm & $R_{\rm g}=$ 10 nm & 20 nm& 40 nm& 80 nm\\
\hline
\hline 2d= $2~\mu$m &0.120 & 0.120 & 0.100 & 0.050 \\
\hline $3~\mu$m& 0.135 & 0.125 & 0.090 & 0.035 \\
\hline $5~\mu$m&0.140 & 0.115 & 0.070 & 0.020 \\
\hline $8~\mu$m& 0.135 & 0.095 & 0.050 & 0.010 \\
\hline $10~\mu$m& 0.130 & 0.085 & 0.040 & 0.010 \\ \hline
\end{tabular}}
\end{table}
\begin{table}
\center{
\begin{tabular}{|c|c|c|c|c|}
\hline
$d_l=$ 150 nm & $R_{\rm g}=$ 10 nm & 20 nm& 40 nm& 80 nm\\
\hline
\hline 2d= $2~\mu$m &0.140 & 0.140 & 0.125 & 0.075 \\
\hline $3~\mu$m& 0.160 & 0.150 & 0.120 & 0.055 \\ \hline$5~\mu$m&
0.180 &
0.150 & 0.095 & 0.035 \\ \hline $8~\mu$m& 0.180 & 0.130 & 0.070 & 0.020 \\
\hline $10~\mu$m& 0.175 & 0.120 & 0.060 & 0.015 \\ \hline
\end{tabular}
\caption{The $\nu=2$ plateau widths obtained at default
temperature for two depletion lengths $d_l$ (left 75 nm, right 150
nm), while $\gamma_I=0.05$ is fixed (defined in
Eq.~\ref{eq:gammaI} and the related text below). The widths are
given in units of $\hbar\omega_c/E_F^0=\Omega_c/E_F^0$.}
\label{table:1}}
\end{table}
\begin{equation}V_{\rm g}(r)=\frac{\widetilde{V}_{\rm imp}}{\pi R_{\rm
g}^2}\exp{(-\frac{r^2}{R_{\rm g}^2})}. \end{equation}Together with the
impurity concentration, the relaxation time is defined as $\tau_0=\frac{\hbar^3}{N_I\widetilde{V}_{\rm imp}^2m^*}$ and in the
limit of delta impurities (\emph{i.e.} $R_{\rm g}\rightarrow0$)
the Landau level width $\Gamma$ takes the form $\Gamma=\sqrt{\frac{4N_I\widetilde{V}_{\rm imp}^2}{2 \pi l^2}}$. It is useful to
define the impurity strength parameter to investigate the effect of disorder by \begin{equation}\gamma_I^2=(\Gamma/\hbar \omega_c)^2= \frac{2N_I\widetilde{V}_{\rm
imp}^2m}{\pi \hbar^3 \omega_c}, \label{eq:gammaI}\end{equation} given in units of magnetic
energy $\hbar \omega_c=\frac{\hbar eB}{m}=\Omega_c$ and as a normalization parameter we
fix the magnetic energy at 10 T.
At this point we would like to make a remark on the concepts
short/long range impurities and short/long range potential
fluctuations, which is commonly mixed. By short range impurity
potential we mean that $R_{\rm g}\lesssim l$, however, by short
range potential fluctuation a length scale of the order of
$200-300$ nm is meant. The long range impurity potential
corresponds to $R_{\rm g}> l$ and long range potential
fluctuation is of the order of micrometers. Thus, when considering
short range impurities the potential fluctuations may be long
range, if $N_I$ is not large ($< \% 5$ of the donor
concentration). We have also observed that, the long-range
potential fluctuations are more efficiently screened by the 2DES
and their range can be at the order of 500 nm at most, when
assuming large impurity concentration, \emph{i.e.} $N_I>\% 10$.
\begin{figure}[!t]
\centering{
\includegraphics[width=.9\linewidth,angle=0]{fig5.pdf}
\caption{The calculated Hall resistances at default temperature
assuming a 5 $\mu$m sample considering three characteristic value
of broadening parameter. The lowest mobility ($\gamma_I$=0.3)
shows the narrowest plateau. \label{fig:10}}}
\end{figure}
In light of the above findings and formulation we now investigate
the widths of the quantized Hall plateaus. Figure ~\ref{fig:9}
presents the calculated Hall resistances at a fixed temperature
for typical single impurity ranges. We observe that, when
increasing $R_{\rm g}$ the plateau widths remain approximately the
same, with a small variation, which is in contrast to the
experimental findings, \emph{i.e.} if the system is low mobility
(small $R_{\rm g}\Rightarrow$ highly broadened DOS) the plateau
are larger. In fact changing $R_{\rm g}$ from 10 nm to 20 nm
should increase the zero $B$ field mobility almost an order of
magnitude, when fixing the other parameters (see \emph{e.g} table I of
Ref.~\cite{siddiki2004}). The contradicting behavior is due
to the fact that the levels become broader when increasing the
single impurity range, therefore the incompressible strips become
narrower, which results in a narrower plateau. However, the long
range potential fluctuations are completely neglected, therefore
the effect(s) of disorder on the quantized Hall plateaus cannot be
described in a complete manner. To investigate the effect of the
single impurity range we systematically calculated the plateau
widths; table~\ref{table:1} depicts the calculated widths of the
Hall plateaus considering different sample widths, depletion
lengths, filling factors and $R_{\rm g}$. One sees that the
plateau widths are affected by the increase of impurity range,
however, in a completely wrong direction,\emph{ i.e.} plateaus
become narrower when decreasing the mobility. As we show in the
next section, it is not sufficient to describe mobility only
considering the range of a single impurity. Moreover, we also show
that the other two parameters defining $B=0$ mobility are either
not important or behaves in the opposite direction when
calculating the resistances.
Next we investigate the effect of the remaining two parameters,
$\widetilde{V}_{\rm imp}$ and $N_I$. However, these two parameters
both effect the level width simultaneously, thereby the widths of
the incompressible strips. Hence, one cannot to distinguish their
influence on the QHPs separately. Typical Hall resistances are
shown in Fig.~\ref{fig:10} calculated at default temperature
considering different impurity parameters. Similar to the range
parameter, we observe that the plateau widths become narrower when
the mobility is low, which also points that our single particle
based level broadening calculations are not in the correct
direction. Such a behavior is easy to understand, when we decrease
the mobility either by increasing the impurity concentration or by
the amplitude of the impurity potential, the Landau levels become
broader due to collisions. This means that, both the energetic and
spatial gap between two consequent levels is reduced, hence the
resulting incompressible strips are also narrower and fragile even
at low temperatures. A detailed investigation on the
incompressible widths depending on impurity parameters are
reported in Ref.~\cite{Guven03:115327}. It is known that if
there exists an incompressible strip wider than the Fermi
wavelength the system is in the quantized Hall
regime~\cite{siddiki2004}, therefore, if the gap is reduced the
incompressible strips are smeared, thus the quantized Hall plateau
vanish. As a general remark on the single particle theories, we
should note that such a reduced gap is also a gross problem for
the non-interacting
models~\cite{Laughlin81,Buettiker88:317,Halperin82:2185}, however,
one can overcome this discrepancy by making localization
assumptions~\cite{Kramer03:172}. Namely, one assumes that even
within the broadened Landau levels there are states, which are
localized, therefore electrons cannot contribute to transport.
Hence, although the gap is small (levels are broad) these
localized states serves as a reasonable candidate to explain the
low mobility behavior. In the early days of IQHE it was a great
challenge to describe and observe these localized
states~\cite{Cai86:3967}. Recent
experiments~\cite{Ahlswede01:562,Yacoby00:3133,josePHYSE,Ashoori05:136804}
show clearly that, the localization assumptions are not relevant
in all the cases, \emph{i.e.} narrow and high mobility samples.
Moreover, the universal behavior of the localization length
dictated by these theories fail~\cite{criticalexpo:09}. An
explicit treatment of the activation energy~\cite{Serpilactivation:09} and critical exponents are left to an
other publication. \subsection{Size effects on plateau widths}
\begin{figure}[!t]\centering
\includegraphics[width=.8\linewidth,angle=0]{fig6a.pdf}
\includegraphics[width=.8\linewidth,angle=0]{fig6b.pdf}
\caption{a) The calculated Hall resistances at a large $B$ interval
at default temperature, setting $2d=5~\mu$m, $R_{\rm g}=20$ nm and
$\gamma_I=0.05$, while changing the depletion length. It is
clearly seen that depletion length is much more important than the
single impurity parameters in determining the plateau widths. (b) The direct comparison of the plateau widths considering
different sample sizes. The impurity parameters and depletion
lengths are kept constant. Calculations are done at
$k_BT/E_F^0=0.02$, whereas the donor density is $4\times10^{15}$
m$^{-2}$ for all sample sizes.
\label{fig:11}}
\end{figure}
Another important parameter in defining the plateau widths is the
depletion length $d_l$. The slope of the confinement potential
close to the edges essentially determines the widths of the
incompressible strips~\cite{Chklovskii92:4026}, which in turn determines the plateau
widths. In Fig.~\ref{fig:11} we show the $\nu$=2 plateau calculated
for two different depletion lengths, we see that for the larger
depletion the plateau is more extended. Since, the larger the
depletion is, the smoother the electron density is. Therefore,
resulting incompressible strips are wider, hence the plateau. Such an argument will fail if one considers a
highly disordered large sample, which we discuss in
Sec.~\ref{manyimp}. Next, we compare the plateau widths of
different sample sizes while keeping constant the disorder
parameters and depletion length. Figure ~\ref{fig:11} depicts the
sample size dependency of $\nu=2$ plateau width. It is seen that
the larger samples present wider plateaus, if the magnetic field
is normalized with the center Fermi energy, $E_F^0$. One can
understand this by similar arguments given above, \emph{i.e.} if
the sample is narrow the variation of the confinement potential is
stronger, therefore the incompressible strips become narrower,
hence, the plateaus. The discrepancy between the experimental
results and the screening theory of the IQHE is solved if one
considers not only the single impurity potentials but also the
overall \emph{disorder} potential landscape generated by the
impurities. In the next part of this section, we investigate the
effect of the long range potential fluctuations on the quantized
Hall plateaus and find that, when the mobility is reduced the
plateaus become wider and stabile, as it is observed in many
experiments, (see \emph{e.g.}
Refs.~\cite{Haug87:SdH,josePHYSE,jose:prl}).
\begin{figure}[!t]
\begin{minipage}{1\linewidth}\centering
\includegraphics[width=.70\linewidth,angle=0]{fig7.pdf}
\end{minipage}
\caption{Self-consistently obtained Hall resistances for a
modulated system considering a sample of 3 $\mu$m. The depletion
lengths and other single impurity parameters are kept fixed,
whereas the parity of the modulation period is set 5.\label{fig:13}}
\end{figure}
\subsection{Many many impurities: Potential fluctuations\label{manyimp}}
So far we have investigated the effect of single impurity
potentials on the overall potential landscape in Sec.~\ref{treed}
and on the widths of the plateaus in Sec.~\ref{singleimp}. We have
seen that, at high impurity concentration the overall potential
fluctuates over a length scale of couple of hundred nanometers,
whereas for low $N_I$ concentration such length scale can be as
large as micrometers. Now we include the effect of this long range
potential fluctuations \emph{into} our screening calculations via
modulation potential defined as $V_{\rm mod}(x)=V_0\cos{(\frac{
2 \pi x m_{p}}{2d}})$ where, the modulation period $m_{p}$, is
chosen such that the boundary conditions are preserved. At the
moment, we consider two modulation periods regardless of the
sample width and vary the amplitude of the modulation potential.
In the next section, however, we select these two parameters
from our estimations obtained in Sec.~\ref{impurity} and
Sec.~\ref{treed}.
Figure~\ref{fig:13} depicts the self-consistently calculated Hall
resistances, considering different modulation amplitudes $V_0$ for
a fixed sample width ($2d=3$ $\mu$m) and $m_{p}=5$. We observe
that, the plateaus become wider from the high $B$ field side, when
$V_0$ is increased, \emph{i.e} mobility is reduced. Such a
behavior is now consistent with the experimental findings. Since
the QHPs occur whenever an incompressible strip is formed
(somewhere) in the sample and the modulation forces the 2DES to
form an incompressible strip at a higher magnetic field, therefore
the plateau is also extended up to higher field compared with the
(approximately) non-modulated calculation, $V_0/E_F^0<0.1$.
Our investigation of the impurities lead us to conclude
that, one has to define mobility at high magnetic fields also
taking into account screening effects in general and furthermore
also the geometric properties of the sample such as the width and
depletion length. As an example if we consider an impurity
concentration of $\approx\% 1$ the long range part of the
potential fluctuation can be approximated to 900 nm. However, note
that the amplitude of this fluctuation varies between $\% 5-25$ of
the Fermi energy, considering different separation thicknesses, therefore the wafer
changes from low mobility to intermediate one. Another important
parameter is the number of modulations within the system: a sample
with an extend of $2~\mu$m and $V_{0}/E_F=0.1$ is a high mobility
sample with the same $m_p$ (only 2 maximum), however, sample with
a width of $10~\mu$m is low mobility (10 maximum). In the next
section we study the plateau widths of different mobility samples,
while keeping constant the extend and the amplitude of long range
potential fluctuations (\emph{i.e.} $V_0$ and $m_p$) and short
range impurity parameters ($\widetilde{V}_{\rm imp}$, $N_I$ and
$R_{\rm g}$) under experimental conditions.
\section{Discussion:Comparison with the experiments\label{exper}}
In this final section, we harvest our findings of the previous
sections to make quantitative estimations of the plateau widths,
considering narrow gate defined samples. Our aim is to show the
qualitative and quantitative differences between ``high'' and
``low'' mobility samples, by taking into account properties of the
single impurity potentials and the resulting disorder potential.
The experimental realizations of these samples are reported in the
literature~\cite{josePHYSE,jose:prl}. We estimated in
Sec.~\ref{treed} that, the range of the potential fluctuations is
$\lesssim 500$ nm for low mobility ($N_I>3300$) and is $\gtrsim
1~\mu$m at high mobility. Therefore, the modulation period is
chosen such that many oscillations correspond to low mobility, and
few oscillations correspond high mobility. As an specific example
let us consider a $10~\mu$m sample, for the low mobility we choose
$m_p=19-20$ and for the high mobility $m_p$ is taken as 9 or 10.
The amplitude of the disorder potential is damped to $\% 50$ of
the Fermi energy when considering the effect of spacer thickness,
however, including screening this amplitude is further reduced to
few percents. In light of this estimations the low mobility will
be presented by a modulation amplitude of $V_0/E_F^0=0.5$, whereas
high mobility corresponds to $V_0/E_F^0=0.05$. Therefore, we have
4 different combinations of the disorder potential parameters
yielding four different mobilities considering two sample widths,
as tabulated in table ~\ref{table:2}. The second important aspect
of the disorder is the single impurity parameters, for low
mobility set we choose $R_{\rm g}=20$ nm and $\gamma_I=0.3$,
whereas for high mobility $R_{\rm g}=10$ nm and $\gamma_I=0.05$ is
set. Remember that, the range of the single impurity is much less
important than $\gamma_I$ in determining the plateau width (see
sec.~\ref{singleimp}).
\begin{table}
\begin{tabular}{|c|c|c|c|}
\hline
mobility & $m_p$ (10 $\mu$m) & $m_p$ (2 $\mu$m) & $V_0/E_F^0$ \\
\hline
\hline
low & 19-20& 5-6 & 0.5 \\
\hline
intermediate 1 & 9-10 & 2-3 & 0.5 \\
\hline
intermediate 2 & 19-20 & 5-6& 0.05 \\
\hline
high & 9-10 & 2-3 & 0.05 \\
\hline
\end{tabular}
\centering \caption{A qualitative comparison of the mobility in
the presence of magnetic field also taking into account
self-consistent screening. Mobility also depends on the size of
the sample when screening is also considered.}\label{table:2}
\end{table}
Figure~\ref{fig:15} summarizes our results considering above
discussed mobility regimes for two different sample widths. In
Fig.~\ref{fig:15}a, we show the calculated Hall resistances for a
sample of 10 microns with the highest mobility (solid (black)
line) and intermediate 1 mobility (broken (red) line). The solid
line is the highest mobility since the range of the fluctuations
are at the order of 1 $\mu$m and the amplitude of the modulation
potential is five percent of the Fermi energy. The broken line
presents the intermediate mobility considering a modulation
amplitude of fifty percent. We observe that the lower mobility
wafer presents a larger quantized Hall plateau, which is now in
complete agreement with the experimental results. Moreover, our
calculation scheme is free of localization assumptions in contrast
to the known literature and we only considered a very limited
level broadening, \emph{i.e.} $\gamma_I=0.05$. In fact our results
also hold for Dirac-delta Landau levels, however, for the sake of
consistency we choose the broadening parameters according to the
selected disorder parameters. In Fig.~\ref{fig:15}c, we show two
curves for even lower mobilities, the solid line corresponds to
the intermediate 2 case, whereas the broken line is the lowest
mobility considered here. The potential fluctuation range
(\emph{i.e.} the modulation period) is chosen to present the low
mobility wafer. We again see that for the lowest mobility the
quantized Hall plateau is enlarged considerably from both edges of
the plateau. These results explicitly show that the quantized Hall
plateaus become broader if one strongly modulates the electronic
system by long range potential fluctuations, either by changing
the range \emph{or} the amplitude of the modulation. Similar
results are also obtained for a relatively narrower sample
$2d=3~\mu$m, Fig.~\ref{fig:15}b and ~\ref{fig:15}d, however, we
see that decreasing the range of the potential fluctuation is more
efficient in enlarging the quantized Hall plateaus when compared
to the effect of the amplitude of the modulation.
\begin{figure}[!t]
\centering{
\includegraphics[width=.9\linewidth,angle=0]{fig8.pdf}
\caption{Line plots of the Hall resistance as a function of
magnetic field considering two sample widths ($2d=10~\mu$m left
panels, $2d=3~\mu$m right panels) and impurity concentrations
($\sim\%3$ (a) and (b), $\sim \%20$ (c) and (d)). Here the single
impurity parameters are calculated from Eq.~\ref{eq:gammaI},
otherwise other parameters are the same. \label{fig:15}}}
\end{figure}
The last interesting investigation is on the parity of the
modulation period, \emph{i.e.} whether $m_p$ is odd or even.
Figure~\ref{fig:16} presents the different behavior when
considering even (a) or odd (b) periods. Here, all the disorder
parameters are kept fixed, other than the parity. We see that for
the even parity the plateau is shifted towards the high field
edge, both for $\nu=2$ and 4, whereas for the odd parity the
plateau is enlarged from both sides. This tendency is also
observed for the larger sample (not shown here). We attribute this
behavior again to the formation of the incompressible strips,
however, this time only to the one residing at the center of the
sample, \emph{i.e.} the bulk incompressible strip. The picture is as follows: If the maxima of the modulation
potential is at the center of the sample, the incompressible strip
is formed at a higher magnetic field value, whereas, the edge
incompressible strips become narrower at the lower field side.
Hence, due to the larger incompressible strip at the bulk of the
sample the plateau is shifted to the higher field, in contrast,
due to the narrower (compared to the unmodulated system) edge
strips the plateau is cut off at higher fields. Since, the edge
incompressible strip becomes narrower than the extend of the wave
function. For the odd parity, the edge incompressible strips
become wider, therefore, the plateau extends to the lower $B$
fields. The enhancement at the high field edge results from the
two maximum in the proximity of the center. For a better
visualization of the incompressible strip distribution we suggest
reader to look at Fig.2 of Ref.~\cite{Siddiki:ijmp} and
Fig.1 of Ref.~\cite{Siddiki02:Oxford}. Such a shift of the
quantized Hall plateaus is also reported in the
literature~\cite{Haug87:SdH} and is attributed to the asymmetrical
density of states due to the acceptors in the
system~\cite{Gerhardts87:asymmetry}. We claim that, the shift due
to the modulation parity change observed in our calculations
overlap with their findings. Note that in our calculations we only
consider symmetric DOS, however, replacing a maxima with a minima
at the confinement potential corresponds to the acceptor behavior
of the dopants. A systematic experimental investigation is
suggested to understand the underlying physical mechanism, where
the system is doped with small number of acceptors.
\begin{figure}[!t]
\centering{
\includegraphics[width=.5\linewidth,angle=0]{fig9a.pdf}
\includegraphics[width=.5\linewidth,angle=0]{fig9b.pdf}}
\caption{Even-odd parity dependency of the Hall plateaus at high
impurity concentration. (a) corresponds to a ``acceptor'' doped
wafer, whereas in (b) the ionized impurities are positively
charged. \label{fig:16}}
\end{figure}
\section{Conclusion}
In this work we tackled with the long standing and widely
discussed question of the effect of disorder on the quantized Hall plateaus. The distinguishing aspect of our approach relies on the
separate treatment of the long and short range of the disorder
potential. We show that assuming Gaussian impurities is not
sufficient to describe long range potential fluctuations, however,
is adequate to give a prescription in defining the density of
states broadening and conductivities. The discrepancy in handling
the long range potential fluctuations is cured by the inclusion of
a modulation potential to the self-consistent calculations. We
estimated the range of these fluctuations from our analytical and
numerical calculations considering the effect of dielectric spacer
and the screening of the 2DES. It is observed that spacer damps
the short range fluctuations effectively, whereas the direct
Coulomb interaction is dominant in screening the long range
fluctuations. Utilizing the estimations of the range and the
amplitude of potential fluctuations, we classified mobility in
four groups and calculated the Hall resistances within the
screening theory of the quantized Hall effect. We found that the
Hall plateaus are wider when decreasing the mobility, not
surprisingly. However, the most important point of our theory is
that, we do not consider any localization assumptions, still
obtain correct behavior of the plateau widths. We show that $B=0$
and/or short range impurity defined mobility is not adequate to
describe the actual mobility at high magnetic fields, moreover,
one has to include geometrical properties of the sample at hand.
A natural persecutor theoretical investigation of the present work
should deal with the activated behavior of the longitudinal
resistance within the screening theory. As it is well known, the
properties of the localized states, \emph{e.g}. the localization length,
is usually obtained from the activation experiments~\cite{Matthews05:497}. Moreover,
spin generalization of the screening theory~\cite{afifPHYSEspin} is necessary to
describe and investigate odd integer quantized plateaus also considering level broadening, namely disorder.
\ack
One of the authors (A.S.) would like to thank E. Ahlswede, S. C.
Lok and J. Weiss for the enlightening discussions on the disorder
from ``an experimentalist'' point of view. The Authors
acknowledges, the Feza-Grsey Institute for supporting the III.
Nano-electronic symposium, where this work has been conducted
partially and would like to acknowledge the Scientific and
Technical Research Council of Turkey (TUBITAK) for supporting
under grant no 109T083.
\section*{References}
\bibliographystyle{unsrt}
|
\section{Introduction}
\label{s_intro}
Finding and studying the most distant galaxies formed during the
epoch of reionisation, more recent than 1 Gyr after the Big Bang, is
one of the challenges of contemporary observational astrophysics.
Over the past few years considerable progress has been made
in this field, pushing the observable limits beyond redshift 6
with the use of ground-based facilities and satellites.
A variety of observational programs have tried to locate $z > 6$ galaxies
using different observational techniques, mostly involving either searches for
\ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ emission through narrow-band filters or searches using the Lyman break
technique -- also called the dropout technique.
These have been performed either in blank fields or in fields with galaxy clusters,
which act as strong gravitational lenses, targetting different depths and survey areas.
The objects found in this way are line emitters or Lyman break galaxies (LBGs).
Although \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ emitters are among the most distant galaxies
with spectroscopically confirmed redshifts \citep[see][]{Iye06,Ota08},
few have been found at $z \ga7$
\citep[see e.g.][]{Cuby07,Stark07,Willis08,Hibon09,Sobral09}.
Furthermore, because of their faintness the photometry available is
inadequate in terms of depth to allow studies of their stellar populations.
Surveys using strong gravitational lensing were among the first
to pave the way in the of study $z>6$ galaxies
\citep[see][]{Kneib04,Pello04,Egami05,Richard06,Richard08,Bradley08,Zheng09}.
Ultra-deep fields with the Hubble Space Telescope (HST) including
near-IR observations with NICMOS have uncovered a handful of $z \sim 7$
candidates in blank fields \citep{Bouwens04, Labbe06, Henry08}.
These pilot studies also showed that some of the $z \ga 7$ galaxies
could be detected at 3.6 and 4.5 $\mu$m\ with Spitzer,
thus probing the rest-frame optical emission from these objects
\citep{Egami05,Labbe06}.
Since then, surveys of z-dropout galaxies (targeting $z \sim 7$
objects) have been extended to cover larger areas, primarily with ground-based instruments
\citep{Mannucci07,Capak09,Castellano09,Hickey09,Ouchi09,Wilkins09},
but also with HST \citep{Henry07,Henry09,Gonzalez09}.
In most cases, however, only a few near-IR photometric bands are
available, providing so far information only on source counts
and luminosity functions, but precluding more detailed studies of the physical
properties of the sources.
Notable exceptions are the work of \citet{Capak09}, who present
three bright ($J \sim 23$) $z \ga 7$ galaxy candidates from the COSMOS 2 square
degree field, and the study of \citet{Gonzalez09} finding 11
fainter (J$_{\rm 110W}$ $\sim$ 26.--27.5) $z \sim 7$ galaxies in the two GOODS fields.
Both studies benefit from a coverage including optical, near-IR, and Spitzer
bands.
Observations taken recently with the newly installed WFC3 camera
on-board HST have just been released, resulting in publications from
four independent groups identifying faint (J$_{\rm 125}$ $\sim$ 27--29) $z \ga 7$
galaxies, based on the combination of the deepest available ACS/HST and
WFC3 data \citep{Oesch09,Bouwens09_z8,Bunker09,Mclure09,Yan09}.
While these objects are too faint to be detected at the current
limits of the deepest Spitzer images, a stack of 14 z-dropout galaxies
from \citet{Oesch09} shows tentative (5.4 and 2.6 $\sigma$) detections
at 3.6 and 4.5 $\mu$m, respectively \citep{Labbe10}.
Given these detected $z \sim 7$ galaxies (or candidates) with
available multi-band photometry, it is of interest to
determine their physical properties such as stellar ages, reddening,
stellar masses, star-formation rates, and related properties such as
their formation redshift, specific star-formation rate, and others.
Several studies have addressed these questions using
different modeling tools \citep[see][]{Bouwens09_betaz7,Capak09,Gonzalez09,Labbe10}.
However, some consider only special types of star-formation
histories (constant star-formation rate), or zero dust extinction, and except for
\citet{Capak09} none of them
accounts for the effects of nebular emission (lines and continua)
present in star-forming galaxies. Neglecting the latter may
in particular lead to systematically older stellar ages,
to lower dust extinction, and differences in stellar masses, as shown
by \citet{SdB09} for $z \sim 6$ galaxies.
Furthermore, the uncertainties in the derived physical parameters
are not always determined or addressed.
Last, but not least, no ``uniform'' study of the entire data sets
of $z \sim 7$ galaxies has yet been undertaken using the
same methodology and modeling tools.
For all these reasons, we present a critical analysis of the physical
properties of the majority of $z\sim$ 6--8 galaxies that have been
discovered recently.
Nebular emission can significantly alter
the physical parameters of distant star-forming galaxies derived
from broad-band photometry. The main reason for this is that the
emission lines, which are invariably present in the
H~{\sc ii}\ regions accompanying massive star-formation, strengthen with
redshift, because their observed equivalent width scales with
$(1+z)$. Since the main emission lines are in the optical
(rest-frame) domain and few are in the UV, their presence can mimick
a Balmer break in absorption, a signature usually interpreted as an age
indicator for stellar populations \citep{Kauffmann03,Wiklind08}.
This effect of emission lines, and to a lesser
extent also nebular continuum emission, can lead to degeneracies in
broad-band SED fits of high-$z$ galaxies as e.g., shown by
\citet{Zackrisson08} and \citet{SdB09}.
The presence of both nebular lines and continua and their contribution
to broad-band photometry is well known in nearby star-forming galaxies,
such as very metal-poor objects (e.g., I Zw 18, SBS 0335-052, and others),
blue compact dwarf galaxies and related objects
\citep[cf.][]{Izotov97,Papaderos02,Pustilnik04,Papaderos06}.
The strongest evidence of a significant contribution of the nebular continuum
in some nearby star-forming galaxies is the observational finding of a Balmer jump
in emission \cite[see][]{Guseva07}.
For these reasons, it is important to include nebular emission
in SED fits of distant starbursts and to examine their effect
on the derived physical properties.
In the present paper, we analyse samples of $z\sim$ 6--8 galaxies discovered
recently. The data are compiled from the literature,
including the brightest objects from the sample of \citet{Capak09},
the ``intermediate'' sample of \citet{Gonzalez09}, and
the faintest z-dropouts recently found with the WFC3 camera.
Applying our up-to-date spectral energy distribution
(SED) fitting tool, we search in particular for possible trends
in the physical parameters of $z \sim 7$ galaxies over a range of
$\sim$ 6 magnitudes, i.e., a range of $\sim$ 250 in flux. First results
from our analysis are presented here. A more detailed and extensive
study of the properties of z-dropout galaxies and comparisons with
objects at lower redshift will be published elsewhere.
In Sect.\ \ref{s_obs}, we summarise the galaxy sample and the SED
fitting method. In Sect.\ \ref{s_res}, we present our results for the three
subsamples. The overall results of the whole $z\sim 7$ LBG sample
and implications are discussed in Sect.\ \ref{s_discuss}, where we also
compare our results to those for LBGs at lower redshift.
Our main conclusions are discussed and summarised in Sect.\ \ref{s_conc}.
We assume a flat $\Lambda$CDM cosmology with $H_0=70$ km s$^{-1}$\
Mpc$^{-1}$, $\Omega_M=0.3$, and $\Omega_{\rm vac}=0.7$.
All magnitudes are given in the AB system.
\section{Observational data and modelling tools}
\label{s_obs}
\subsection{$z \approx 7$ galaxy samples}
To determine the physical properties of $z \approx 7$ galaxies and their
uncertainties, we chose the following three samples:
\begin{itemize}
\item Two of the three bright ($J \sim 23$) z dropout galaxies from the COSMOS survey, discovered
by \citet{Capak09}. We refer to these as the ``bright sample''.
\item The 11 z$_{\rm 850LP}$\ dropout objects identified by \citet{Gonzalez09} from the HST ACS
and NICMOS data in the GOODS and HUDF fields, plus their mean SED.
These objects typically have J$_{\rm 110W}$ $\sim$ 26.--27.5, and are referred to as the
``intermediate sample''.
\item The ``faint sample'', including 15 of the 16 z$_{\rm 850LP}$\ dropout candidates
found by \citet{Oesch09} in the HUDF using the newly installed WFC3 camera of HST,
and the 15 additional objects identified as $z \sim$ 6--9 candidates by \citet{Mclure09}.
The photometry is taken from \citet{Mclure09}.
They typically span a range from J$_{\rm 125}$ $\sim$ 27 to 29.
We also include the stacked SED obtained by \citet{Labbe10} for 14 objects
from the \citet{Oesch09} sample, which shows tentative (5.4 and 2.6 $\sigma$)
detections in the 3.6 and 4.5 $\mu$m\ bands of Spitzer.
\end{itemize}
The following photometric data/filters was used for the samples:
{(1)} $i^+$, $z^+$ from SuprimeCam on SUBARU, $J$, $H$, $K$ from WIRCAM on the CFHT,
and channels 1-4 of IRAC/Spitzer for the \citet{Capak09} sample.
Since object 2 is detected in the $i+$ band and at 24 $\mu$m, and its SED
indicates a low redshift \citep[$z\sim 1.6$,][]{Capak09}, we exclude it from
our analysis.
{(2)} B$_{\rm 435}$, V$_{\rm 606}$, i$_{\rm 776}$, z$_{\rm 850LP}$\ filters of ACS/HST, J$_{\rm 110W}$\ and H$_{\rm 160W}$\
of NICMOS/HST, $Ks$, and channels 1-2 of IRAC/Spitzer for the \citet{Gonzalez09} objects.
We adopted the properties of the $Ks$ filter of ISAAC/VLT for all objects.
{(3)} B$_{\rm 435}$, V$_{\rm 606}$, i$_{\rm 776}$, z$_{\rm 850LP}$\ filters of ACS/HST, Y$_{\rm 105}$, J$_{\rm 125}$, H$_{\rm 160}$\
filters of WFC3/HST, and channels 1-2 from IRAC/Spitzer for the faint sample.
The original photometry from the respective papers was adopted.
Except for the 3 objects in the bright sample for which one spectral line
was found for each of them, no spectroscopic redshifts are available
for these objects. We therefore treat the redshift as a free parameter
for all objects.
\subsection{SED fitting tool}
To analyse the broad-band photometry, we use a modified version of the {\em Hyperz}\
photometric redshift code of \citet{hyperz} described in
\citet{SdB09}.
The main improvement with respect to both earlier versions and other
SED fitting codes is the treatment of nebular emission (lines
and continua), which can have a significant impact on the broad-band
photometry of high redshift galaxies and hence their
derived properties \citep[see][]{SdB09}.
We use a large set of spectral templates
\citep[primarily the GALAXEV synthesis models of ][]{BC03},
covering different metallicities and a
wide range of star formation (SF) histories (bursts, exponentially
decreasing, or constant SF), and we add the effects of nebular
emission. Models with a more sophisticated description of
stellar populations, chemical evolution, dust evolution,
and different geometries \citep[see e.g.,][]{Schurer09}
are not used, given the small number of observational constraints.
We adopt a Salpeter IMF from 0.1 to 100 \ifmmode M_{\odot} \else M$_{\odot}$\fi,
and we accurately consider the returned ISM mass from stars.
Nebular emission from continuum processes and lines is added to the predicted spectra
from the GALAXEV models, as described in \citet{SdB09}, proportional
to the Lyman continuum photon production.
The relative line intensities of He and metals are taken from \cite{Anders03},
including galaxies grouped into three metallicity intervals covering
$\sim$ 1/50 \ifmmode Z_{\odot} \else Z$_{\odot}$\fi\ to \ifmmode Z_{\odot} \else Z$_{\odot}$\fi.
Hydrogen lines from the Lyman to the Brackett series are included
with relative intensities given by case B.
Our treatment therefore covers the main emission lines of H, He, C, N, O, and S from the
UV (\ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi) to 2 $\mu$m\ (rest-frame), necessary for fitting the SED of
galaxies at $z>4$ up to 10 $\mu$m\ (IRAC channel 4).
The free parameters of our SED fits are:
the redshift $z$,
the metallicity $Z$ (of stars and gas),
the SF history described by the timescale
$\tau$ (where the SF rate is SFR $\propto \exp^{-t/\tau}$),
the age $t$ defined since the onset of star-formation,
the extinction $A_V$ described here by the Calzetti law \citep{Calzetti00},
and whether or not nebular emission is included.
In some cases, we exclude the \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ line from the synthetic
spectra, since this line may be attenuated by radiation transfer processes
inside the galaxy or by the intervening intergalactic medium.
Here we consider
$z \in [0,12]$ in steps of 0.1, three metallicities $Z/\ifmmode Z_{\odot} \else Z$_{\odot}$\fi=$1, 1/5, 1/20,
$\tau=$ 5, 7, 10, 30, 50, 70, 100, 300, 500, 700, and 1000 Myr in addition to
bursts and SFR=constant,
ages up to the Hubble time, and $A_V=$ 0--2 (or 4) mag in steps of 0.2.
In general, the combination of all parameters leads to $\sim 3 \times 10^6$
models for each object.
Non-detections are included in the SED fit with {\em Hyperz}\ by setting the
flux in the corresponding filter to zero, and the error to the $1 \sigma$
upper limit.
For all the above combinations we compute the $\chi^2$ and the scaling
factor of the template, which provides information about the SFR and $M_\star$,
from the fit to the observed SED.
Minimisation of \ifmmode \chi^2 \else $\chi^2$\fi\ over the entire parameter space yields the best-fit
parameters.
To illustrate the uncertainties in the resulting fit parameters,
we examine the distribution of \ifmmode \chi^2 \else $\chi^2$\fi\ across the entire parameter
space. To determine confidence intervals from the
\ifmmode \chi^2 \else $\chi^2$\fi\ distribution, the degree of freedom must be known to
determine the $\Delta \ifmmode \chi^2 \else $\chi^2$\fi$ values corresponding to different
confidence levels, or Monte Carlo simulations must be carried out.
In any case, the photometric uncertainties, typically taken
from SExtractor, would also need to be examined critically, since
these may be underestimated, and since errors in the relative
photometric calibration between different telescopes/instruments,
which affect SED fits, are usually not taken into account.
We chose to plot the 1D \ifmmode \chi^2 \else $\chi^2$\fi\ distribution for the parameter
of interest, marginalised over all other parameters, so that the reader
is able to appreciate these distributions.
Illustrative confidence intervals are determined by assuming
$\Delta \ifmmode \chi^2 \else $\chi^2$\fi \approx 1$, the value for one degree of freedom.
This should provide a lower limit to the true uncertainties.
More quantitative estimates of the uncertainties will be given in
a subsequent publication, which will include the analysis of a larger sample
of LBGs at different redshift.
\section{Results}
\label{s_res}
\subsection{Photometric redshifts: overview of the full sample}
\label{s_zphot}
The photometric redshifts \ifmmode z_{\rm phot} \else $z_{\rm phot}$\fi\ of the objects from the three subsamples
were discussed by \citet{Capak09,Gonzalez09}, and \citet{Mclure09}. Since these
authors use different spectral templates and methods, it is useful
to examine briefly the redshifts we derive from our SED fits,
and their dependence om nebular emission.
Figure \ref{fig_zphot} shows the best-fit model values for \ifmmode z_{\rm phot} \else $z_{\rm phot}$\fi\ using
either standard templates (i.e., neglecting nebular emission), or including
nebular emission (lines and continua), and the latter but neglecting
the contribution from \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi.
Clearly, the contribution of \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ can lead to higher photometric
redshifts, since it can compensate for the drop of the flux
shortward of \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi, and hence lead to drop-out at higher \ifmmode z_{\rm phot} \else $z_{\rm phot}$\fi.
With the prescription used for \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ in our models (i.e.\ maximum
emission according to Case B recombination) this typically leads to $\Delta \ifmmode z_{\rm phot} \else $z_{\rm phot}$\fi \la 1$.
In some cases, e.g., for 4 objects from the \citet{Gonzalez09} sample,
the shift is larger. The reason for this large shift is the
available filter set, which include z$_{\rm 850LP}$\ and J$_{\rm 110W}$\ for this sample,
whereas Y$_{\rm 105}$, a filter that is intermediate between z$_{\rm 850LP}$\ and J$_{\rm 125}$, is available
for the WFC3/HST (faint) subsample.
By including the nebular continuum and all spectral lines {\em except} \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\
(blue symbols) one recovers essentially the same photometric redshifts
as with standard templates. This is expected, since the Lyman break
--- the main feature determining \ifmmode z_{\rm phot} \else $z_{\rm phot}$\fi\ --- can only be strongly affected
by \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi.
We compared our photometric redshifts against objects
with known spectroscopic redshifts, where possible. For a sample of $B$, $V$, and
$i$-dropouts from the GOODS fields, we find good agreement for the
majority of objects using the GOODS-MUSIC photometry \citep{Santini09}.
For this sample, spanning objects with $z_{\rm spec} \sim$ 4--6,
our results are essentially the same with/without nebular emission,
and with/without \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi.
Since \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ emission may be weaker than predicted by the models,
because of the multiple scattering in the presence of dust \citep{Verh08}
and/or because of the intervening IGM, we subsequently consider
models including all nebular lines except \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi.
It must, however, be noted that for objects with strong \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ emission
the true redshift may be higher than \ifmmode z_{\rm phot} \else $z_{\rm phot}$\fi\ obtained from photometric codes
neglecting this line.
\begin{figure}[tb]
\includegraphics[width=8.8cm]{13946fg1.eps}
\caption{Comparison of best-fit model photometric redshifts for the three samples using SEDs
without (x-axis) and with nebular emission (y-axis). Circles, squares, and triangles
indicate the bright, intermediate, and faint samples, respectively. Black symbols show the
comparison with all nebular lines, i.e.\ including also \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi, blue symbols
with \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ suppressed. The error bars shown here denote the 68\% confidence interval
derived by {\em Hyperz}\ from the redshift probability distribution derived assuming
$P(z) \propto \exp(-\ifmmode \chi^2 \else $\chi^2$\fi(z))$, where $\ifmmode \chi^2 \else $\chi^2$\fi(z)$ is the minimum chi-square value
over all other parameters.
Discussion is given in text.
}
\label{fig_zphot}
\end{figure}
\begin{figure}[tb]
\includegraphics[width=8.8cm]{13946fg2.eps}
\caption{Observed (blue points) and fitted SEDs (solid lines) of object 3
from \citet{Capak09} using our spectral templates with nebular
emission. \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ emission has been supressed here.
Two best-fit models are shown comparing the standard (Calzetti) attenuation law
for $A_V=1.2$ with the Galactic extinction law of \citet{Seat79}. Using the latter
with $A_V=1.6$ reproduces nicely the observed dip in the H-band.
The errorbars of the observed wavelength indicate the surface of the normalised
filter transmission curve. Upper limits in flux indicate $1 \sigma$ limits.
Red crosses show the synthesised flux in the filters.}
\label{fig_capak_3}
\end{figure}
\subsection{Bright sample}
\label{s_bright}
Our SED fits for these objects yield results (redshift probability distributions
and physical parameters) broadly in agreement with \citet{Capak09}, which is
unsurprising since these authors also include nebular lines in their
analysis using the Le Phare code.
Since we include SEDs spanning metallicities from \ifmmode Z_{\odot} \else Z$_{\odot}$\fi\ to 1/50 \ifmmode Z_{\odot} \else Z$_{\odot}$\fi\
(in contrast to \ifmmode Z_{\odot} \else Z$_{\odot}$\fi\ only), we obtain a wider range of acceptable
fit parameters.
For example, we find evidence of significant reddening in object 1,
with $A_V \sim$ 1--2.6, in agreement with the best-fit value of $A_V=1.2$
given by \citet{Capak09}. The corresponding \ifmmode \chi^2 \else $\chi^2$\fi\ distribution
is shown in Fig.\ \ref{fig_capak_av}.
The situation is similar for object 3, although for
more moderate extinction ($A_V \sim$ 0.6--1.6).
The corresponding range of ages, SFRs, and stellar masses for both
objects, obtained with and without nebular emission, are illustrated in
Figs.\ \ref{fig_capak_age} to \ref{fig_capak_sfr}.
Approximately (within $\Delta \ifmmode \chi^2 \else $\chi^2$\fi \approx$ 1-2.3)
object number 1 (3) has thee best-fit model parameters
$t \sim$ 0--30 (10-200) Myr,
$\log(M_\star) \sim$ 10.8--12. (10.6--11.6) \ifmmode M_{\odot} \else M$_{\odot}$\fi,
and SFR $\sim$ 100--10$^6$ (10--10$^4$) \ifmmode M_{\odot} {\rm yr}^{-1} \else M$_{\odot}$ yr$^{-1}$\fi.
We note that a very large range in SFR is obtained from the SED fits
since both the SF history and the extinction are kept free.
The high SFR tail is related to solutions with high extinction
and very young populations, where the UV output per unit SFR remains
below the equilibrium value reached after typically $\ga$ 100 Myr.
For comparison, using the standard SFR(UV) calibration of \citet{Kenn98}
and assuming $z=8$, one obtains SFR$_{\rm obs} \sim$ 240 \ifmmode M_{\odot} {\rm yr}^{-1} \else M$_{\odot}$ yr$^{-1}$\fi\
without extinction correction for object 1, and SFR$_{\rm corr} \sim$ 6500 \ifmmode M_{\odot} {\rm yr}^{-1} \else M$_{\odot}$ yr$^{-1}$\fi\
adopting $A_V=1.2$ and Calzetti's attenuation law.
As noted by \citet{Capak09}, object 3 is not detected in H, possibly indicating
a dip in the flux between J and K (see Fig.\ \ref{fig_capak_3}). If real, this
dip could be explained by the 2175 \AA\ dust absorption feature, as also mentioned
by \citet{Capak09}. Using the Galactic extinction law from \citet{Seat79}
provides excellent fits with $A_V \sim 1.6$, as shown in Fig.\ \ref{fig_capak_3}.
The possible indication for a 2175 \AA\ dust absorption feature at such a high redshift
is in contrast to evidence so far, suggesting the absence of this feature
\citep[see e.g.,][]{Maiolino04}.
\begin{figure}[tb]
\includegraphics[width=8.8cm]{13946fg3.eps}
\caption{Redshift probability distribution $P(z) \propto \exp(-\ifmmode \chi^2 \else $\chi^2$\fi(z))$
for all \citet{Gonzalez09} objects as derived from {\em Hyperz}\ including our standard
templates. Coloured lines indicate the mean SED (green), and selected objects
discussed in the text (GNS-zD2, GNS-zD3, GNS-zD4 in red, blue, magenta).
}
\label{fig_pz_gonz}
\end{figure}
\begin{figure}[tb]
\includegraphics[width=8.8cm]{13946fg4.eps}
\caption{
Observed (blue points) and fitted SEDs (solid lines) of object 8
(GNS-zD4) from \citet{Gonzalez09} using standard Bruzual \& Charlot solar metallicity
models. The best-fit ($\chi^2=0.3$) is at $z=7.14$
with a stellar population of $\sim 130$ Myr (black line), the secondary solution
($chi^2=2.1$) is a 4.5 Gyr old population at $z=1.29$ (magenta).
The errorbars of the observed wavelength indicate the surface of the normalised
filter transmission curve. Upper limits in flux indicate $1 \sigma$ limits.
Red crosses show the synthesised flux in the filters.}
\label{fig_sed_gonz8}
\end{figure}
\begin{figure}[tb]
\includegraphics[width=8.8cm]{13946fg5.eps}
\caption{
Observed (blue points) and best-fit SEDs (solid lines) of object 1
(UDF-640-1417) from \citet{Gonzalez09} using standard Bruzual \& Charlot solar metallicity
models.
The magenta line shows the best-fit for templates without nebular emission, and
assuming that SFR=constant and $A_v=0$ following \citet{Gonzalez09}.
The age of the population
is found to be $\sim$ 500 Myr.
The black line show the best-fit allowing for nebular emission and arbitrary
SF histories and extinction, yielding a much younger age plus some extinction
($t=6$ Myr, $A_V=1.2$).
}
\label{fig_sed_gonz1}
\end{figure}
\subsection{Intermediate sample}
\label{s_intermediate}
The redshift probability distribution of the objects from the \citet{Gonzalez09}
sample is shown in Fig.\ \ref{fig_pz_gonz}. All objects have a best-fit
photometric redshift at high $z$ ($z \ga$ 6--6.5). However, 6 of the 11 objects
(ID 4, 5, 6, 7, 8, 10) also have an acceptable fit at low redshift ($z
\sim$ 1--2) with a probability comparable to the high-$z$ solution;
the most unreliable objects being ID 6, 7, and 8.
The 3 brightest objects in J$_{\rm 110W}$\ (ID 1, 9, 11) favour clearly high-$z$ solutions,
since they provide the largest ``leverage'' on the Lyman-break between
J$_{\rm 110W}$\ and the optical data.
Figure \ref{fig_sed_gonz8} shows an example of an object with
both acceptable low- or high-redshift solutions of similar quality.
The observed SED is reproduced well by a low extinction,
young, starburst at high-$z$ ($z=7.14$ here) or by a 4.5 Gyr old
stellar populations with $A_V=0.4$ at $z=1.29$.
Reducing the probability of a low-$z$ interloper would obviously
be possible with deeper optical photometry, which would place tighter constraints
on the Lyman break.
Deep K-band data (not available for this object) or other
constraints on the shape of the SED between 2 and 3.6 $\mu$m, may allow us
to distinguish between the two solutions shown here, and provide stronger
constraints on the possible Balmer break -- hence the age -- of this
object.
Object 1 (UDF-640-1417) from \citet{Gonzalez09} is such an example,
benefiting from deeper optical imaging and $Ks$ data, as shown in
Fig.\ \ref{fig_sed_gonz1}. The former leads to a well-defined and clearly
most probable high-$z$ solution at $\ifmmode z_{\rm phot} \else $z_{\rm phot}$\fi \sim$ 6.7--6.9.
However, here the observed spectral shape between the rest-frame UV and
optical range (probed by JHK and 3.6-4.5 $\mu$m, respectively)
may imply a degeneracy between age and extinction.
While \citet{Gonzalez09} fit this SED with a dust-free
population of several 100 Myr age (cf.\ our magenta line),
we obtain a tighter fit with models including nebular emission
for a young population plus dust reddening
(typically $t \la$ 10 Myr, $A_V \sim$ 0.8--1.2).
The distribution of \ifmmode \chi^2 \else $\chi^2$\fi\ for $A_V$ and other parameters are shown
in the Appendix (Figs.\ \ref{fig_gonz_chi2_av} to \ref{fig_gonz_chi2_sfr}).
In total, we find three objects
(UDF-640-1417, GNS-zD5, HDFN-3654-1216)
with best-fit solutions for $A_V \approx$ 0.6--1.2 and relatively young age ($t \la 10$ Myr).
Incidentally these are the three brightest objects in J$_{\rm 110W}$, which may
suggest a trend of extinction with magnitude (cf.\ below).
However, the significance of non-zero extinction is not very high,
in particular for GNS-zD5 and HDFN-3654-1216, where the 3.6 $\mu$m\
flux is affected by a bright neighbouring source \citep[cf.][]{Gonzalez09}.
Considering the entire parameter space for the whole sample
(cf.\ Figs.\ \ref{fig_gonz_chi2_av} to \ref{fig_gonz_chi2_sfr}), we
find that age and dust extinction of most objects are not well constrained,
and could reach from few Myr up to the age of the universe at that redshift,
and from $A_V \sim$ 0 to $\la 1.6$ mag for some objects.
In particular, the data does not allow us to conclude that these
galaxies show no sign of dust extinction.
Furthermore, their age and hence formation redshift remains poorly constrained.
The same is also true for the mean SED from \citet{Gonzalez09}, which
yields results compatible with those of the individual objects, as expected
(see thick line in the plots).
For comparison, \citet{Gonzalez09} find fairly old best-fit stellar-mass-weighted ages
$t_w \sim 200-400$ Myr\footnote{With their definition, one has $t_w = 0.5 \times t$
for SFR=const assumed by these authors, where $t$ is our definition of the stellar age.}
typically.
Both the assumption of SFR=const and $A_v=0$ lead to the highest age, since both
effects minimise the ratio of the rest-frame visible/UV light, the main age constraint.
Allowing for wide range of SF histories, variable extinction, and for nebular emission
yields, on average, a broader age range (between a few Myr and up to the maximum age),
room for
extinction up to $A_V \la 1.$, stellar masses from $10^{8.5}$ to $10^{10}$ \ifmmode M_{\odot} \else M$_{\odot}$\fi, and
SFR $\sim$ 2--100 \ifmmode M_{\odot} {\rm yr}^{-1} \else M$_{\odot}$ yr$^{-1}$\fi. In other words, the properties of the galaxies
from this sample of intermediate brightness (J$_{\rm 110W}$ $\sim$ 26--28), are clearly
more uncertain than indicated by \citet{Gonzalez09}, who consider only a restricted
range of the parameter space.
\subsection{Faint sample}
\label{s_faint}
Using the first UDF observations taken with the newly installed WFC3 camera
onboard HST, four studies have identified $\sim$ 11 to 20
$z\sim7$ galaxy candidates (or z$_{\rm 850LP}$\ drop-outs) \citep{Oesch09,Bunker09,Mclure09,Yan09}.
What can be said about their physical properties?
\subsubsection{Photometric redshifts}
\citet{Mclure09} had previously examined the photometric redshifts and uncertainties
for their sample, which also covers the majority of $z \sim 7$ galaxies found
by the other groups \citep{Oesch09,Bunker09,Yan09}.
Unsurprisingly, our results using a modified version of the {\em Hyperz}\ code
also used by \citet{Mclure09} and a slightly more extended template library,
confirm their findings. In particular, for the $z \sim 7$ sample of \citet{Oesch09}
we find that their objects consistently show photometric redshifts with well-defined
probability distributions peaking between $z \sim$ 6.3 and 7.6.
For fainter z-dropouts and Y-dropouts, the photometric redshift becomes
far more uncertain, and a significant fraction of the objects could
also be low-$z$ galaxies. As already pointed out by \citet{Mclure09} and \citet{Capak09},
the depth of the optical imaging becomes the limiting factor for objects
that faint in the near-IR.
\subsubsection{UV slope}
One group pointed out that the fainter of these objects
had very blue UV-continuum slopes, $\beta$, indicative of
``non-standard'' properties of these galaxies \citep{Bouwens09_betaz7}.
Their data, shown as red squares in Fig.\ \ref{fig_beta}, exhibits
a trend of decreasing $\beta$ (as estimated from their (J$_{\rm 125}$-H$_{\rm 160}$) colour) towards fainter
magnitudes.
From the very steep slopes (i.e., low values of $\beta \sim -3$) reached
in faint objects, \citet{Bouwens09_betaz7} claim that extremely low metallicities and
large Lyman continuum escape fractions seem to be
required to understand these objects, since ``standard'' evolutionary synthesis models
predict minimum values of $\beta \sim -2.5$ for young stellar populations.
As the data and the errorbars from different groups plotted
in Fig.\ \ref{fig_beta} show, we cannot reach similar conclusions,
given the uncertainties in the colour measurement used to determine $\beta$.
For the bulk of the sources, the $\beta$-slope is compatible within $1 \sigma$ with
normal values of $\beta \sim -2.5$ or flatter slopes.
Furthermore, it is unclear whether the (J$_{\rm 125}$-H$_{\rm 160}$) colour exhibits any
systematic trend towards fainter magnitudes.
The observations do not exclude different properties such as
extremely low metallicities and large Lyman continuum escape fractions
for some of the objects at $z \sim 7$.
However, the low significance of these deviations do not justify
making assumptions that differ significantly from those commonly adopted for
the analysis of lower redshift objects.
\begin{figure}[tb]
\includegraphics[width=8.8cm]{13946fg6.eps}
\caption{UV slope $\beta$ between $\approx$ 1550 and 1940 \protect\AA\ computed
from $\beta = 4.29($J$_{\rm 125}$-H$_{\rm 160}$$)-2.0$ for the $z \sim 7$ galaxy candidates
from different WFC3/UDF samples.
The photometry from various sources has been used: \citet{Oesch09} (red squares),
\citet{Bunker09} (yellow triangles),
\citet{Yan09} (green circles), \citet{Mclure09} (blue squares, only objects
in common with Oesch et al.).
}
\label{fig_beta}
\end{figure}
With the WFC3 filters used in this survey, it is possible
to generate unusually blue (J$_{\rm 125}$-H$_{\rm 160}$) colours in certain
circumstances from spectral templates including nebular emission.
Such a case is illustrated in Fig.\ \ref{fig_lyaboost}, showing
a fit to object 2502 from the \citet{Mclure09} sample.
Shown here is a model of a very young stellar population
with solar metallicity including lines and nebular continuum
emission redshifted to either $z=6.96$ (dashed, magenta line) or
$z=7.97$ (solid, black line). In the latter case, the strong
intrinsic \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ emission (with $W(\ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi)^{\rm rest} \approx 200$ \AA)
boosts the flux in both J$_{\rm 125}$\ and Y$_{\rm 105}$\ since
these filters overlap by $\approx 0.1$ $\mu$m.
This provides a very blue (J$_{\rm 125}$-H$_{\rm 160}$) colour, and enough
flux in Y$_{\rm 105}$\ to ensure that this object does not appear as a Y$_{\rm 105}$-drop
even at $z \sim 8$. If at $z \sim 7$, as implied by SED fits excluding
the \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ line (magenta line) or from simple colour-criteria
designed to select $z \sim 7$ galaxies \citep[e.g][]{Oesch09},
the (J$_{\rm 125}$-H$_{\rm 160}$) colour is not affected by \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi.
However, this is not necessarily the case for all z-dropout galaxies
since strong \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ emission, if present, may mimic a lower
redshift (cf.\ Sect.\ \ref{s_zphot}).
The likelyhood of this situation remains difficult to establish,
especially since \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ may be differentially affected by dust,
and scattered by the IGM.
\begin{figure}[tb]
\includegraphics[width=8.8cm]{13946fg7.eps}
\caption{Best-fit SEDs of object 2502 (photometry in blue symbols) of
\citet{Mclure09} for solar metallicity models including nebular emission.
The black solid line shows the model including all lines (\ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ in particular)
with a best-fit \ifmmode z_{\rm phot} \else $z_{\rm phot}$\fi=7.97. Red crosses indicate the flux in the filters
for this model.
The dashed magenta line shows the best-fit model
excluding \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi, found at \ifmmode z_{\rm phot} \else $z_{\rm phot}$\fi=6.96.
This object has an observed $($J$_{\rm 125}$-H$_{\rm 160}$$)=-0.51$ colour, corresponding to $\beta=-4.2$
Note the strong effect \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ may have on (J$_{\rm 125}$-H$_{\rm 160}$) and \ifmmode z_{\rm phot} \else $z_{\rm phot}$\fi,
although this object is a z$_{\rm 850LP}$-dropout, thought to be
at $z \sim 7$.}
\label{fig_lyaboost}
\end{figure}
\subsubsection{Age and reddening}
For the 15 objects in common between the Oesch and McLure samples, we find young
stellar populations ($t \la 10$ Myr) as best-fits, and zero extinction, except for objects
688, 835, and 1092, with $A_V \sim$ 0.2--0.6.
However, as for the objects from the intermediate sample, the distribution of \ifmmode \chi^2 \else $\chi^2$\fi\ is
very flat (cf.\ Figs.\ \ref{fig_mclure_chi2_av} to \ref{fig_mclure_chi2_sfr}), allowing a
wide range of extinctions ($A_V \sim$ 0--1.2), ages of $t \sim$ 0 to several 100 Myr,
stellar masses from $10^7$ to few times $10^9$ \ifmmode M_{\odot} \else M$_{\odot}$\fi, and SFRs from 0.1 (or less) to
$\la 200$ \ifmmode M_{\odot} {\rm yr}^{-1} \else M$_{\odot}$ yr$^{-1}$\fi, for most objects in the faint sample.
The wide age range, is possible, e.g., since the upper limits at 3.6 and 4.5 $\mu$m\
do not provide a strong enough constraint on the optical to UV flux of these faint
objects. Given the rapid evolution with time in the mass/light ratio involved here
(mostly the UV--optical
domain), the uncertainty in the ages translates into a large spread in
stellar masses, as show by Fig.\ \ref{fig_mclure_chi2_mass}. The wide range of
acceptable SFR values is due to both age and SF history (parametrised
here by the e-folding timescale $\tau$) being kept free, in contrast e.g., to
commonly used SFR(UV) calibrations assuming SFR=constant and ages $t \ga$ 100 Myr.
\begin{figure}[tb]
\includegraphics[width=8.8cm]{13946fg8.eps}
\caption{Best-fit SEDs to the SED of the stack of 14 z-dropout galaxies
from \citet{Oesch09} measured by \citet{Labbe10}.
Photometry is shown by blue symbols; including the 1 $\sigma$ limit for Ks.
Red crosses indicate the synthetic model fluxes in the filters.
The magenta line shows the best-fit using solar metallicity templates
without nebular emission, yielding a maximum age ($t \sim 700 Myr$), $A_V=0$,
and other parameters also similar to \citet{Labbe10}.
The black line shows the best-fit model including nebular emission (and \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\
suppressed), with an age of $t=4$ Myr, $A_V=0.2$, and a stellar mass
of $M_\star \sim 5.\times 10^7$ \ifmmode M_{\odot} \else M$_{\odot}$\fi, more than 1 dex lower than the mass
corresponding to the old fit.}
\label{fig_labbe_stack}
\end{figure}
\subsubsection{Physical properties from the stacked SED}
The individual z-dropout candidates of \citet{Oesch09} and \citet{Mclure09} are
undetected in the deep, available 3.6 and 4.5 $\mu$m\ Spitzer images, but
\citet{Labbe10} stacked the images of 14 of the 16 z-dropout galaxies from \citet{Oesch09},
obtaining 5.4 and 2.6 $\sigma$ detections in these bands.
Fitting the SED of this stack, we find that the physical parameters
are more tightly constrained, as the thick line in Figs.\ \ref{fig_mclure_chi2_av} to
\ref{fig_mclure_chi2_sfr} show.
Overall our best-fit values (with or without nebular
emission) are very similar to those obtained by \citet{Labbe10},
bearing the different definitions of stellar ages $t$ and $t_W$ in
mind. Furthermore, the values of the physical parameters derived for
the majority of the individual objects is compatible with the values
determined from the stack.
An uncertainty remains, however, in the age and consequently also in the stellar
mass determination. We first obtain secondary solutions with $\Delta \ifmmode \chi^2 \else $\chi^2$\fi \sim$ 1--2
with young ages ($\sim$ 4--5 Myr) and a small extinction ($A_v \la$ 0.2--0.4)
as we consider all metallicities, both with our without nebular
emission. Furthermore, if we suppress the \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ line we improve the fit, leading
to a best-fit at young ages ($\sim$ 2--7 Myr).
The corresponding \ifmmode \chi^2 \else $\chi^2$\fi\ distributions illustrating these results are shown
in Figs.\ \ref{fig_stack_av} to \ref{fig_stack_sfr}, and
SED fits from these models are shown in Fig.\ \ref{fig_labbe_stack}.
This figure clearly illustrates how an apparent Balmer break can
be explained by an old population (here $t \sim$ 700 Myr) or by nebular
emission from a younger population, as already shown by \citet{SdB09}.
A suppression of the \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ line is justified since \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\
may be attenuated by radiation transfer processes inside the galaxy or by the
intervening intergalactic medium.
In any case, neglecting nebular emission is inconsistent for spectral
templates with recent ($\la 10$ Myr) or ongoing star formation.
Finally, in comparison with the bright and intermediate samples
we may also question why the faintest $z\sim 7$ objects
should have the oldest stellar populations, whereas the ages
of brighter objects are compatible with a broad range of ages,
including young ones.
The uncertainty in the age also translates into an uncertainty in stellar mass.
Whereas the estimated average mass is $M_\star \sim (1-2)\times 10^9$ \ifmmode M_{\odot} \else M$_{\odot}$\fi\
for the old population \citep[cf.][]{Labbe10}, it is more than a factor of
10 lower for young ages (see Fig.\ \ref{fig_stack_mass}), since nebular
emission contributes partly to the rest-frame optical domain.
The SFR is, however, hardly affected by this uncertainty (see Fig.\ \ref{fig_stack_sfr}),
since it is more sensitive to the rest-frame UV light present in
both young and old star-forming populations.
In consequence, the specific SFR (SFR$/M_\star$) could be significantly higher
than advocated by \citet{Labbe10}.
Before performing spectroscopy for these objects -- a currently
impossible task -- to examine if the 3.6 $\mu$m\ filter is truly
affected by emission lines as predicted by the model, the present data
does not allow us to completely rule out one or the other solution.
\section{Discussion}
\label{s_discuss}
We discuss the effects of varying the model assumptions on
the physical parameters derived.
We then examine the main derived properties and possible correlations
among them for the ensemble of galaxies studied here, and compare
any correlations found to those of lower redshift galaxies.
Finally, we discuss some implications of our results.
\begin{figure}[tb]
\includegraphics[width=8.8cm]{13946fg9.eps}
\caption{Comparison of best-fit values for the age $t$ (upper left), extinction $A_V$ (upper right),
stellar mass $M_\star$ (lower left), and star-formation rate SFR (lower right)
obtained with different model assumptions. The x-axis is the value obtained from
our ``reference model'' (including nebular emission but no \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi) except for the
SFR panel; the other values
are plotted using different colours:
models with $\tau \ge 10 Myr$ (yellow),
models with \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ (blue),
models without nebular emission (red), and
models without nebular emission and SFR$=const$ (green).
For the SFR comparison, we use the SFR(UV) value for the x-axis, and black
symbols for the reference model.
Filled circles, square, triangles show the bright, intermediate, and faint
samples. The dashed line is the one-to-one relation.
See discussion in text.}
\label{fig_compare}
\end{figure}
\begin{figure}[tb]
\includegraphics[width=8.8cm]{13946fg10.eps}
\caption{Best-fit value for the age (upper left), extinction (upper right),
stellar mass (lower left), and star-formation rate (lower right) as a function of
the observed $J_{AB}$ magnitude. All best-fit values are taken from the
models with $\tau \ge 10 Myr$ (model 1).
Filled circles, square, triangles show the bright, intermediate, and faint
samples. Large open circles show the fit values obtained by
\citet{Capak09}, \citet{Gonzalez09}, and \citet{Labbe10} for the 2 bright objects
and from the mean/stacked SED, respectively.
The dashed line in the SFR--$J_{AB}$ panel shows the standard SFR(UV) line
assuming $z=7$ and no extinction for all objects. See discussion in text.}
\label{fig_overview}
\end{figure}
\subsection{Effects of varying model assumptions}
Varying the model assumptions in broad-band SED fits affects the inferred
physical parameters of distant galaxies in ways that
have been discussed in several studies, e.g.,
by \citet{Yabe09} in quite some detail \citep[cf.\ also][ and others]{Papovich01,Sawicki07,Gonzalez09}.
The most relevant assumptions are the star-formation histories, metallicity, the inclusion
of dust extinction, and the adopted extinction law. Furthermore, the inclusion
of nebular emission and the assumptions made to do so also affect the results
as shown here and in \citep[cf.\ also][ and the latter for a discussion of the
effect of \ifmmode {\rm H}\alpha \else H$\alpha$\fi]{SdB09,Yabe09}.
The impact of different model assumptions on the best-fit parameters of our sample is shown
in Fig.\ \ref{fig_compare}, where we plot the values from our ``reference model'' (including
nebular emission, \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ suppressed, all SF histories, all ages , all extinction values, and all
metallicities) on the x-axis, and the same from comparison models on the y-axis.
The models we consider here for comparison are:
{\em 1)} models with $\tau \ge 10$ Myr (yellow),
{\em 2)} models with \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ (blue),
{\em 3)} models without nebular emission (red), and
{\em 4)} same as 3), but for SFR=constant only (green).
SEDs without nebular emission (3) may be taken as an extreme case of
models with a very large escape fraction ($f_{\rm esc} \approx 1$) of ionising
radiation from the Lyman continuum, whereas in the other cases
we implicitly assume no escape ($f_{\rm esc} \approx 0$), maximising
thus nebular emission if young massive stars are present.
One may theoretically expect an evolution of $f_{\rm esc}$ with redshift,
galaxy mass, and other properties, although large differences remain even
between simulations \citep[see e.g.,][]{Gnedin08,Wise09,Razoumov09}.
Intermediate values of $f_{\rm esc}$ may be included in future models.
Circles, squares, and triangles in Fig.\ \ref{fig_compare} represent objects from
the bright, intermediate, and faint samples, respectively.
In Figs.\ \ref{fig_compare} to \ref{fig_mass_sfr}, we now
include all objects of \citet{Mclure09} in the faint sample, i.e.\ 15 objects
in addition to the 15 of \citet{Oesch09}, to maximise the sample size and
because the results with the additional sample do not show noticeable differences.
We do not show the effect of fixing metallicity, since it is small compared
to the other effects discussed here.
We now discuss the dependence of the physical parameters on these models assumptions
one-by-one.
\subsubsection{Stellar mass}
As noted and discussed by \citet{Yabe09} and others, the
stellar mass $M_\star$ is the least sensitive parameter to the model assumptions,
especially when measurements
for the rest frame optical domain are available. This is mainly because
it is derived from the absolute scaling of the overall SED, and that the
mass-to-light ratio in the optical does not change much with age and star-formation
history.
\subsubsection{Age}
In contrast, age is the most sensitive quantity \citep[cf.][]{Yabe09}.
From Fig.\ \ref{fig_compare} it is clear that models without nebular emission (red and green)
yield older ages in most cases, as already discussed above \citep[cf.][]{SdB09}.
By assuming constant SFR, we also tend to infer older ages, as is well known, but not in all cases (see green
symbols).
The inclusion (suppression) of \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ can also lead to younger (older) ages, as shown by
the blue symbols. This is the case for some objects because of the
increase of \ifmmode z_{\rm phot} \else $z_{\rm phot}$\fi\ (cf.\ Sect.\ \ref{s_zphot}), which
in turn demands a younger age (steeper UV spectrum) to reproduce the same observed
slope in the UV. In other words, to some extent there is also a degeneracy between age and redshift
\footnote{In addition, there is the well-known age--extinction degeneracy.}
when \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ is taken into account.
Naively one may have expected the opposite, namely that the inclusion of \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ would
lead to older age estimates, since for a fixed redshift, the contribution of \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\
produces a bluer UV spectrum \citep[cf.][]{Finkelstein09}. This shows the importance
of consistently fitting the physical parameters and redshift using the same spectral templates.
\subsubsection{Extinction}
Although a spread is obtained in $A_V$ for the different model assumptions,
the best-fit model values of $A_V$ correlate reasonably well around our ``reference value''
for most objects. The greatest differences arise when the SF history is varied
-- here imposing $\tau \ge 10$ Myr leads to lower $A_V$ and higher ages (because of the
age--extinction degeneracy), as shown by the yellow symbols -- and, for the same reason,
when nebular lines are omitted, as shown by some outliers with red symbols.
The largest differences in stellar mass are found for some objects where
nebular emission leads to a strong age reduction and hence a lower M/L ratio.
\subsubsection{SFR}
The star-formation rate SFR deserves special comment.
The value of the current SFR
(SFR$(t)$) obtained from SED fits depends strongly on both SFH and age.
Formally, SFR=0 for instantaneous burst models, which are also considered here
and in other publications. Furthermore since the UV luminosity emitted
per unit SFR varies by $\sim 1$ order of magnitude within $\sim$ 100 Myr even
for constant star formation, the current SFR resulting from SED fits can
differ\footnote{In general, one has SFR$(t) > $SFR(UV), since stars over a broad
age range contribute to the UV output, and since $L_{\rm UV}/$SFR is lower
at young ages.}
from the usual SFR(UV) calibrations by 1 dex or more depending on both SF history and age.
For this reason, we compare in Fig.\ \ref{fig_compare} the best-fit SFR values to
SFR(UV), the value derived from the $J$ magnitude using the \citet{Kenn98} calibration,
and assuming no extinction.
The following differences can be seen: SFR $>$ SFR(UV) is obtained for many objects, where
$A_V>0$ and $t \ll$ 100 Myr.
Lower SFR$(t) <$SFR(UV) values are obtained for some objects with relative short
timescales $\tau$, i.e., rapidly declining SF histories.
The comparisons shown here all include a marginalisation over the three
metallicities $Z$ considered in our models. As already mentioned, fixing $Z$
leads to small differences.
Given our limited knowledge of high-$z$ galaxies, other variations in the input physics,
e.g., for the star-formation history or the reddening law, could be considered.
Quantifying this is beyond the scope of this paper. The effect of
different reddening laws on the physical parameters, however, can be understood
quite simply, and is e.g., illustrated in \citet{Yabe09}.
Rising SF histories, which are not considered here, have, e.g., been advocated
by \citet{Finlator07}. We note, that these SF histories would in general
correspond to a higher extinction than decreasing ones, to reproduce the observed
SEDs of LBGs \citet[cf.][]{Finlator07}. This would strengthen our arguments
in favour of dust at $z \sim 7$.
\subsection{Correlations between physical properties in $z \sim 7$ galaxies}
Although the uncertainties in the physical parameters are relatively large for most
objects, it is helpful to search for correlations between them, and
with observed quantities.
Figure \ref{fig_overview} shows the best-fit values for age, $A_V$, stellar mass, and
SFR of all objects from the 3 samples, plotted as a function of the $J$-band magnitude,
which traces approximately the rest-frame UV at 1500 \AA.
The values plotted here are taken from our model 1, i.e., the reference
model plus the constraint that $\tau \ge 10$ Myr, in particular to assure
that SFR is properly defined.
We also compare the best-fit values obtained by \citet{Gonzalez09} and \citet{Labbe10}
for the mean/stacked SED for the intermediate and faint samples, and the available
fit results from \citet{Capak09} for the bright objects. The origin of the
differences with these authors have already been discussed above.
As is clear for the different samples, stellar ages and
extinction cover a wide range of values with no clear or strong trend.
It is possible that a tendency of increasing $A_V$ for brighter objects exists, as
would be expected from other studies at lower redshift (cf.\ below).
This trend becomes clearer when $A_V$ is plotted as a function of stellar mass,
as shown in Fig.\ \ref{fig_mass_av}.
We also find a tentative trend of the extinction $A_V$ with galaxy mass $M_\star$,
as shown in Fig.\ \ref{fig_mass_av}.
Both the stellar mass and SFR show clear trends with the $J$-band magnitude, albeit
with a significant scatter. We note
that the scatter in the mass-magnitude relation for the $z \sim 7$ objects
does not significantly decrease when 3.6 or 4.5 $\mu$m\ photometry is used.
Deviations of the best-fit SFR from the simple ``standard'' calibration, also indicated
in Fig.\ \ref{fig_overview} for $A_V=0$, are caused by non-zero extinction, age effects,
and exponentially decreasing SF histories, or combinations thereof, as already discussed above.
Figure \ref{fig_mass_sfr} shows the corresponding correlation between stellar
mass and SFR, suggesting a fairly well-defined mass--SFR relation at $z \sim 7$.
Our best-fit values yield on average higher specific star-formation rates
(SSFR$=$SFR$/M_\star$) than the values derived by \citet{Gonzalez09} and \citet{Labbe10}
shown by large open circles, and tend to indicate a relatively large
spread in SSFR.
\begin{figure}[tb]
\includegraphics[width=8.8cm]{13946fg11.eps}
\caption{Black circles, green squares, blue triangles show best-fit values of
$A_V$ versus $M_\star$ derived from our reference model for ojects from the bright,
intermediate, and faint samples respectively.
Red filled symbols show the same but from the reference
model imposing the restriction of $\tau \ge 10$ Myr (model 1).
The dotted line, given by $A_V = \log(M_\star /10^8)^{0.5}$, is shown to
guide the eye.
Small crosses show the fit results from \citet{Yabe09} at $z \sim 5$.
}
\label{fig_mass_av}
\end{figure}
\begin{figure}[tb]
\includegraphics[width=8.8cm]{13946fg12.eps}
\caption{Stellar mass--SFR relation for the $z \sim 7$ galaxies.
Open symbols (circles, squares, triangles) show the ``standard'' SFR(UV) value
(not corrected for extinction) versus mass derived from our reference
model for objects from the bright, intermediate, and faint samples respectively.
Red filled symbols show the best-fit model SFR and $M_\star$ values
when assuming $\tau \ge 10$ Myr (model 1).
Filled hexagons denote the mean values derived by \citet{Gonzalez09} and \citet{Labbe10}.
Small crosses show the fit results from \citet{Yabe09} at $z \sim 5$.
The dashed line shows the relation found by \citet{Daddi07} for $z \sim 2$ SF galaxies.
The dotted (dash-dotted) lines show the locus for SFR=const from $z=\infty$ (10) to 7
corresponding to SSFR=1.3 (3.6) Gyr$^{-1}$.
}
\label{fig_mass_sfr}
\end{figure}
\subsection{Comparison with lower redshift galaxies}
Although an exhaustive comparison with studies of the physical
properties of LBGs at lower redshift is clearly beyond the scope
of the present paper, it is worthwhile comparing briefly
our results with those obtained at $z \sim 5$, and at lower redshift.
To do so we refer to the detailed SED fitting analysis and discussions
of \citet[][ and references therein]{Yabe09}, who have compared the physical parameters of
$z=2$, 3, 5, and 6 LBGs from different samples.
The analysis of both Yabe et al.\ and \citet{Verma07}
of $z \sim 5$ galaxies find
clear signs of significant extinction. The former study finds $E(B-V)$ values ranging
from 0 to $\sim$ 0.5, with a median of $E(B-V)-0.22$, corresponding to
$A_V=0.89$ i.e.\ a factor 9 attenuation of the UV flux!
Our typical $A_V$ values are lower and the uncertainties are large;
but in many cases the best-fit model extinction is $A_V>0$ for the $z \sim 7$ LBGs.
For comparison, \citet{Bouwens09_beta,Bouwens09_betaz7} advocate
an attenuation of the UV flux by a factor 1.35--1.6 (corresponding
to $A_V \sim$ 0.12--0.19 for the Calzetti law) at $z \approx 7$, and
an attenuation by 2.7 ($A_V \sim 0.4$) at $z \approx 5$.
The extinction obtained by \citet{Yabe09} and from our analysis is thus typically
higher by a factor 3 than in the work of Bouwens and collaborators,
who estimate the dust attenuation from the UV slope.
We find a trend of increasing extinction with galaxy mass (Fig.\ \ref{fig_mass_av}).
At $z \sim$ 0--2, it is known that dust extinction increases with
bolometric luminosity (i.e., also with SFR), which in turn increases with stellar
mass \citep[e.g.,][]{Buat05,Buat08,Burgarella07,Daddi07,Reddy06,Reddy08}.
Luminosity-dependent dust corrections have been proposed, e.g., by
\citet{Reddy09} and \citet{Bouwens09_beta}, based on observed variations of the UV slope
with $M_{\rm UV}$. From these results, it is not surprising to find a
similar relation at $z \sim 7$, here expressed as extinction versus
stellar mass. The origin of a mass--extinction relation is most likely
related to that of the mass--metallicity relation
\citep[cf.][]{Tremonti04,Erb06,Finlator07,Maiolino08}.
The data of \citet{Yabe09} at $z \sim 5$, also plotted in
Fig.\ \ref{fig_mass_av}, may show a less clear trend with mass and some offset.
A systematic differential analysis of LBG samples at different redshifts,
will be necessary to shed more light on these issues.
The best-fit model values of the other physical parameters (age, mass, and SFR)
span a similar range as found by \citet{Yabe09} for $z \sim 5$ LBGs.
In contrast to the results of \citet{Gonzalez09} and \citet{Labbe10}, our relatively
young ages resemble those found by \citet{Verma07} and \citet{Yabe09} for $z \sim 5$,
which are younger than $z \sim$ 2--3 LBGs \citep[cf.][]{Shap01,Shap05,Sawicki07,
Yabe09}. However, other studies \citep[e.g.,][]{Stark07_massdensity,Eyles07,Stark09}, find relatively
old ages at $z \sim 6$, and \citet{Yabe09} confirm some of them with
their method. Whether LBGs show a clear trend of decreasing age with
increasing redshift, as one may naively expect, thus remains to be clarified.
For the IRAC-detected objects at $z\sim 7$ (the bright and intermediate
samples), the relation between stellar mass
and absolute optical magnitude (derived from the 4.5 $\mu$m\ flux)
is very similar to that of the $z \sim 5$ objects. The lower masses of the
faint $z \sim 7$ subsample -- undetected by IRAC -- provide a natural
extension towards fainter objects.
A similar behaviour is also found for the SFR, when comparing our
$z \sim 7$ results to those of \citet{Yabe09}.
The $M_\star$--SFR relation found at $z \sim 7$ resembles that of $z \approx 5$ LBGs
(see Fig.\ \ref{fig_mass_sfr}).
We note that most objects lie above the $M_\star$--SFR relation found
at $z\approx 2$ \citep[cf.][]{Daddi07,Sawicki07}.
However, both the behaviour and the scatter found for the $z \sim 7$ objects resembles
that of $z \approx 5$ LBGs, also shown in this plot.
Given the large uncertainties for the $z\sim 7$ objects, and the different
methods used in these analysis, it is possible that the same $M_\star$--SFR
relation (with a similar scatter) is in place from $z \sim 7$ to 2
and that the specific SFR declines to lower redshift
\citep[cf.][]{Daddi07,Elbaz07,Noeske07,Gonzalez09}.
A detailed study of the physical parameters of LBGs over a wide
redshift domain with our modeling tools, including more in-depth comparisons
of the results from different methods/groups, will be presented elsewhere.
\subsection{Implications}
An important result of this study is the possible discovery of dust extinction
in LBGs at $z \sim 7$ from detailed SED modeling, which contrasts with claims of
basically dust-free objects from studies of their UV slope or from too
restrictive SED modeling \citep[cf.][]{Bouwens09_betaz7,Gonzalez09,Labbe10}.
If this were correct, the UV attenuation -- up to $A_V \sim 1.2$ (factor 20)
for the brightest objects -- would imply a higher SFR
density at high redshift than previously inferred.
However, to quantify the average dust correction remains difficult, especially
given the large uncertainties for the individual objects and the possible
dependence of $A_V$ on the UV luminosity and/or stellar mass.
Even if our results imply that a non-negligible dust correction
is required at $z \sim 7$, it is possible that the SFR density decreases with
redshift from $z \ga$ 3--5 on, at least if the median attenuation of
$A_V \sim 0.9$ found by \citet{Yabe09} at $z \approx 5$ is representative.
Our analysis, covering a wider domain in parameters space than previous studies
and also allowing for the effects of nebular emission, shows that a wide range
of stellar ages is acceptable for most $z \sim 7$ LBGs.
The Balmer break observed for some objects does not always imply old ages,
or correspondingly high formation redshifts as already pointed out
by \citet{SdB09}. It can often equally well
be explained by nebular emission, younger populations at non-solar
metallicity, extinction, or a combination thereof.
Younger ages and variations in the star-formation history can lead to solutions
with lower masses and higher SFR for many of the $z \sim 7$ LBGs than
estimated by other groups \citep{Gonzalez09,Labbe10}.
The specific star-formation rate (SSFR$=$SFR$/M_\star$) of
these high-$z$ galaxies may also be higher than thought.
In short, our SED analysis including in particular the effects
of nebular emission and considering a wide parameter space
(SF histories, extinction etc., shows that the physical
parameters of $z \sim 7$ galaxies may differ significantly
from those advocated by other groups, with possibly many
implications for our picture of galaxy formation and evolution
in the early universe.
However, we emphasize that the physical properties of
these objects are affected by large uncertainties (cf.\ Sect.\
\ref{s_res} and Appendix), given their faintness and the available data.
Furthermore, comparisons with results for lower redshifts may
be problematic because of the use of different methods.
Obviously, additional and higher quality data, and a detailed differential study
of the physical parameters of LBGs over a wide redshift domain using state-of-the-art
tools will be very helpful in providing a clearer and more accurate picture
of the properties and evolution of high-$z$ galaxies in the distant universe,
and their link to lower redshift galaxy populations.
\section{Conclusions}
\label{s_conc}
We have presented a homogeneous and detailed analysis of broad-band
photometry for three samples of $z \sim$6--8 galaxies discovered
by the COSMOS survey and with HST
\citep[see][]{Capak09,Gonzalez09,Oesch09,Mclure09, Bunker09}. Their $J$-band
magnitude span a range from $\sim$ 23 to 29, the bulk of them having
$J_{AB} \sim$ 26--29. The broad-band SEDs have been fitted using our
modified version of the {\em Hyperz}\ photometric redshift code
described in \citet{SdB09}, which accounts for the effects of nebular
emission.
In contrast to earlier studies that assumed, e.g., constant
star formation and/or no extinction, we have explored a wide
range of the parameter space without using priors.
The free parameters (and range) of our SED fits are:
the redshift $z$ ($[0,12]$),
the metallicity $Z$ ($Z/\ifmmode Z_{\odot} \else Z$_{\odot}$\fi=$1, 1/5, 1/20),
the SF history described by the e-folding timescale $\tau$
($[0,1000]$ Myr, $\infty$),
the age $t$ defined since the onset of star-formation ($\le t_H$),
the extinction $A_V$ ($[0,2]$ in general) described by the Calzetti law \citep{Calzetti00},
and whether or not nebular emission is included
\footnote{In some cases, we exclude the \ifmmode {\rm Ly}\alpha \else Ly$\alpha$\fi\ line from the synthetic
spectra, since this line may be attenuated by radiation transfer processes
inside the galaxy or by the intervening intergalactic medium.}.
Our main results can be summarised as follows:
\begin{itemize}
\item Overall, we find that the physical parameters of most
galaxies studied here cover a wide range of acceptable values, e.g., within
$\Delta \ifmmode \chi^2 \else $\chi^2$\fi \la 1$ from the best-fit modeling. This finding is independent of
whether nebular emission is included or not.
\item Stellar ages, in particular, are not tightly constrained, even for objects
detected with Spitzer, i.e., with photometry both blue- and redward of
the Balmer break. When nebular emission is taken into account, we find
that the majority of the objects (and the stacked SEDs as well) are
most accurately reproduced by
ages $t < 100$, which are younger than derived in other studies of the same objects
\citep{Gonzalez09,Labbe10}. The younger ages are due to the contribution of
nebular lines to the broad-band rest-frame optical filters, which mimic
to some extent a Balmer break, as already shown by \citet{SdB09}.
\item Examination of the UV slope and SEDs of faint z-dropout
galaxies found with WFC3/HST shows no need for ``unusual''
stellar populations, extreme metallicities, or other physical processes,
advocated previously by \citet{Bouwens09_betaz7}, when the uncertainties are
taken into account \citep[see also][]{Finkelstein09}.
\item Albeit with large uncertainties, our fit results show indications
of dust attenuation in some of the $z \approx$ 6--8 galaxies, with
best-fit model values of $A_V$ up to $\sim$ 1, even among relatively faint
objects ($J_{AB} \sim$ 26--27; cf.\ Fig.\ \ref{fig_overview}).
\item We find a possible trend of increasing dust attenuation with
the stellar mass of the galaxy (Fig.\ \ref{fig_mass_av}) and a
relatively large scatter in specific star-formation rates, SFR/$M_\star$.
\item Our results, including the evidence of dust in $z \approx$ 6--8
galaxies, are consistent with the results and trends from other SED studies
of LBG samples at $z \sim 5$ \citep[see e.g.][]{Verma07,Yabe09}.
\end{itemize}
We will present elsewhere a systematic study of the evolution of the
physical parameters of LBGs at different redshifts, adopting a
homogeneous method and including a detailed error analysis.
\acknowledgements
We would like to thank Matthew Hayes, David Valls-Gabaud,
Rychard Bouwens, and Masami Ouchi for interesting discussions,
and Roser Pell\'o for regular discussions and support with {\em Hyperz}.
This work was supported by the Swiss National Science Foundation.
\newpage
|
\section{\label{intro}Inroduction}
Many properties of high temperature cuprate superconductors,
such as the transition temperature $T_c$, and the presence of pseudogap,
have a strong doping dependence. What happens when two regions with
different doping are put into contact? Since pairing in
cuprate superconductors has a very local
character\cite{Gomes:2007p301,Howald:2001p505,Pan:2001p393,McElroy:2005p326},
one expects the coherence length to be short, and any proximity effects to
occur only on microscopic length scales. Nonetheless, some experiments
have demonstrated that high-$T_c$ Josephson junctions can be made with thicknesses
many times the coherence length\cite{Bozovic:2004p506}.
Other experiments on bilayers have shown transition temperatures higher than that of
either layer in isolation\cite{Yuli:2008p507,Berg:2008p508,Gozar:2008p509}.
Microscopic theories to explain these phenomena include Josephson
coupling between superconducting islands\cite{Kresin:2006p830},
and enhancement of the phase stiffness in underdoped regions
by proximity to overdoped regions\cite{Berg:2008p508,PhysRevLett.101.156401,PhysRevB.79.174509}.
Such scenarios are inherently interesting due to the possibility that
interface superconductivity can occur at temperatures above the maximum
possible in bulk samples. To date no experiments have been performed to
investigate proximity effects locally at microscopic length scales.
Here we report a proximity effect in the cuprate superconductor $\textrm{Bi}_2\textrm{Sr}_2\textrm{CaCu}_2\textrm{O}_{8+\delta}$\ using
scanning tunneling microscopy (STM) and the intrinsic nanoscale spatial
variation of the sample. We show that indeed patches of the sample with
weaker superconductivity can be enhanced if surrounded by patches with
stronger superconductivity, demonstrating that the collapse of superconductivity
is caused by more than local thermal pair breaking.
Our experiments were performed in a compact variable temperature scanning tunneling microscope (STM), with the unique ability to acquire spectroscopy on nearly identical grids of points over a wide range of temperatures.
This allows us to perform spatial correlations using data from multiple temperatures. The pairing gap is measured from the local differential conductance ($dI/dV$), which is proportional to the local density of states.
Figure \ref{fig:spec} shows a representative $dI/dV$ spectrum for the high temperature superconductor $\textrm{Bi}_2\textrm{Sr}_2\textrm{CaCu}_2\textrm{O}_{8+\delta}$ (Bi-2212). At 30~K there is a visible gap in the spectrum, by 70~K it has diminished considerably,
and by 78~K it is gone completely. The gap has been previously described by a $d$-wave
extension to the BCS theory with the addition of a lifetime broadening term\cite{Alldredge:2008p568,Pasupathy:2008p298}.
\begin{figure}
\includegraphics[width=75mm]{fig1}
\caption{\label{fig:spec}Representative $dI/dV$ spectrum at the same point at different
temperatures. The setpoint was at -200~mV. The evolution of the spectrum shows a
gap that closes near 74~K ($T_c$ is 65~K). The inset shows the gap as extracted from the peak location as a function of temperature.}
\end{figure}
For high temperature cuprate superconductors, the superconducting gap
varies from smallest to largest by a factor of two on nanometer scales\cite{Gomes:2007p301,Howald:2001p505,Pan:2001p393,McElroy:2005p326}. By identifying the same atoms
in the topography at multiple temperatures we are able to measure dI/dV spectra
for the same point at different temperatures.
Recently, we have used this technique to visualize the development of superconductivity
on the atomic scale in Bi-2212.
We have found that the temperature dependence of the spectroscopic gap ($\Delta(r,T)$)
is inhomogenous, and that the gap vanishes at a local temperature $T_p(r)$,
which is positively correlated in space with the gap size\cite{Gomes:2007p301}.
Furthermore, the spectroscopic gap below $T_c$ is spatially correlated to the density of states at the Fermi level well above $T_c$ where the gap is absent\cite{Pasupathy:2008p298}.
For this work we simplify the analysis of a large number of points by using a simple,
operational definition for the gap and the pairing temperature.
We follow the definition used previously\cite{Howald:2001p505,Gomes:2007p301},
defining the gap value as the point of maximum differential conductance for
positive bias. This simple procedure agrees with more elaborate fitting schemes
to within 10\% in most cases. We also need a simple method to describe
the temperature evolution of the gap. For this, we define the pairing temperature $T_p$,
which is the highest temperature at which a maximum in the $dI/dV$ spectrum can
be observed within experimental resolution at positive bias.
In underdoped samples, multiple energy features are present in STM spectroscopy\cite{Gomes:2007p301}. While high energy features survive to temperatures well above $T_c$, lower energy features show the closing of a nodal gap near $T_c$\cite{Boyer:2007p633,Pushp:2009p297}.
However, for strongly overdoped samples pseudogap behavior is absent, there is only one spectroscopic feature in tunneling data, and the median value of $T_p$ is near the resistive transition temperature $T_c$. Furthermore, an inhomogeneous superconducting transition is supported by Nernst measurements\cite{Wang:2006p634} and
$\mu$SR\cite{Sonier:2008p637}.
Therefore, $\Delta$ may be taken as the superconducting gap, and $T_p$ as the temperature
for the onset of pairing. We have done detailed measurements of the
local $T_p$ and the local $\Delta$ for a grid of points on an overdoped sample of Bi-2212 with
$T_c = 65\textrm{ K}$, for a range of temperatures from 50~K to 76~K.
These measurements provided evidence that nanoscale regions distinguished
by inhomogeneity can influence each other via the proximity effect.
We show that the fractional variation in $T_p$ is less than that for $\Delta$.
The pairing order parameter in a disordered sample
has been analyzed theoretically at low temperature\cite{fang:017007,balatsky:373} and near the
transition\cite{andersen:060501}, but a direct comparison
of spatial variation between $T_p$ and $\Delta$ was not made.
Our measurements show that the length scale for correlations in $T_p$ is longer by 20\%
than the length scale for the gap. Furthermore, we find that the local gap
can disappear at distinct temperatures for different spatial positions
on the sample, even if these positions have the same magnitude of gap at
low temperature, and this anomalous temperature evolution of the local gap
is influenced by the size of gaps in the surrounding 1 nm region.
\begin{figure*}
\includegraphics[width=150mm]{fig2}
\caption{\label{fig:manymaps}Evolution of the gap distribution with temperature.
\textbf{a}, map of the local pairing temperature ($T_p$). The pairing temperature
is determined from the sequence of images in \textbf{b}-\textbf{o} by finding the
temperature above which there is no maximum in the conductance on the
positive bias side. \textbf{b}-\textbf{o}, a 25 x 28 nm ``gap map'' taken at several different temperatures
showing the distribution of gaps in real space. At each temperature, the
gap map has been aligned to allow point-by-point comparison between maps.}
\end{figure*}
We have the spectral evolution for more than 10,000 points in a 25 x 28 nm
grid\footnote{These data are taken in the same area of the same sample as used
in Ref.~\onlinecite{Pasupathy:2008p298}} at fourteen different temperatures from 50~K to 76~K.
For each temperature,
we can determine the gap at each point, and plot a series of ``gap maps'',
or 2D false color images showing the spatial distribution of gaps,
shown in figure \ref{fig:manymaps}b-o. Using simultaneously recorded topographic information,
we can align gap maps taken at different temperatures, despite their being taken
over slightly different areas with varying amounts of thermal drift and piezo creep.
This is accomplished by interpolating the two dimensional data in order to maximize the agreement between temperatures. To avoid biasing the results all alignment is based exclusively on topographic data.
Points with median-sized gaps surrounded by large gapped regions (consider points with gaps near 20~mV within the box in the upper right of the panels of figure \ref{fig:manymaps}) survive to higher temperatures
than regions with similar absolute gap sizes elsewhere (consider the oval shaped region in the upper left), which supports our main claim.
From the data of figure \ref{fig:manymaps}b-o, we can determine $T_p$ for each point,
restricted to the temperature range for which we have data,
so that we can test our claims statistically. Figure \ref{fig:manymaps}a
shows the spatial distribution of $T_p$. This is well correlated to the low temperature gap value
from figure \ref{fig:manymaps}b, as we expect if the temperature evolution is determined locally.
Figure \ref{fig:hist}a shows a histogram of all measured gaps and their
corresponding $T_p$ values, confirming that the two quantities are well
correlated (81\%). If we use the median value of the gap (20~mV) and the median value
of $T_p$ (64~K), we obtain $2\Delta/k_BT_p = 7.4$ (the green line in figure \ref{fig:hist}a),
consistent with previous work\cite{Gomes:2007p301}.
By choosing bins of gap size and averaging the $T_p$ values in each bin,
we can establish a more detailed relation between $\Delta$ and $T_p$
(the red line in figure \ref{fig:hist}a). The slope of the distribution is significantly
shallower than the linear relation $2\Delta/k_BT_p = 7.4$.
That is, the relative variation of $\Delta$ is greater than that for $T_p$. Indeed, this is quite a significant effect, as it means that, for example, 17~mV gaps will survive, on average, to temperatures 6~K above where one would expect based on a linear relation.
We take this as evidence that $T_p$ is determined not only by the local gap,
but also by the gap in the surrounding region.
Therefore, the variation in $T_p$ will be reduced,
as this quantity is spatially averaged over some window.
\begin{figure}
\includegraphics[width=75mm]{fig3}
\caption{\label{fig:hist}Relation between $\Delta$ and $T_p$. \textbf{a}, two dimensional histogram of $\Delta$ and $T_p$ values, showing the strong correlation. The green line represents the linear relation $2\Delta/k_BT_p = 7.4$, which is determined from the median values of $\Delta$ and $T_p$. The red line represents the average value of $T_p$ as a function of $\Delta$. \textbf{b}, The deviation in $T_p$ from the average, that is $\delta T_p = T_p - \left<T_p \right>$, where $\left<T_p \right>$ is the average $T_p$ for gaps with the same size.}
\end{figure}
We have found direct evidence that the temperature evolution can be
more accurately predicted given knowledge of the gap in the vicinity of a point,
compared with knowledge of the gap at a single point only.
One can clearly see from figure \ref{fig:hist}a that knowledge of the gap
can only determine $T_p$ in a statistical way, that is, different points of equal gap size have measurably different values of $T_p$. If the pairing temperature were to depend only
locally on the gap, the distribution of $T_p$ values with the same gap
would only be due to random fluctuations. Therefore, when we subtract
the mean $T_p$ value (based on gap size) from each measured $T_p$ value,
the result should be uncorrelated noise. However, figure \ref{fig:hist}b shows a
significant spatial pattern associated with this quantity, which we call $\delta T_p$.
\begin{figure}
\includegraphics[width=75mm]{fig4}
\caption{\label{fig:corr}Correlation length scales. The cross correlation between $\delta T_p$ and $\Delta$ as a function of separation, and the autocorrelations of $\Delta$ and $T_p$, radially averaged. The cross correlation shows that areas with large $T_p$ values for their gap size are on average surrounded by larger gap regions. The autocorrelations show that $T_p$ has stronger correlation on length scales above 1~nm. The results are not significantly different if averaged along the $a$ or $b$ axis instead of radially.}
\end{figure}
We can understand the meaning of $\delta T_p$ and relate it to the
gap map using the cross correlation as a function of separation distance.
We define the cross correlation as
\begin{eqnarray}
\label{eqn:cross_corr}\rho(\mathbf{x}) = \frac{1}{N\sigma_{\delta T_p}\sigma_\Delta}\sum_{\mathbf{y}} \delta T_p(\mathbf{y})\Delta(\mathbf{y}-\mathbf{x}),
\end{eqnarray}
where $\sigma_{\delta T_p}$ and $\sigma_\Delta$ are the standard deviations of $\delta T_p$ and $\Delta$ respectively, and $\mathbf{y}$ runs over the $N$ available data points. Although the supermodulation on the $b$ axis does have a small correlation
with the gap, in practice $\rho$ is isotropic within the noise, so we plot a circular average in figure \ref{fig:corr}.
For zero separation, the cross correlations mathematically must be zero,
because we have explicitly eliminated the local relation between $\Delta$ and $T_p$
when we subtract the mean to produce $\delta T_p$. The peak at positive separation
indicates that $T_p$ at a particular point is significantly determined by gaps 1~nm away,
which we attribute to the proximity effect. In addition, there is a long tail of positive correlation,
indicating that $T_p$ is also dependent on gaps several nanometers away.
Figure \ref{fig:corr} also shows the autocorrelation of $\Delta$ and $T_p$ averaged
along both axes and plotted as a function of separation, showing that
the pairing temperature has a longer length scale (1.1~nm at half-maximum)
than the gap (0.9~nm at half-maximum). These correlations occur on length scales longer enough than the 0.25~nm pixel spacing to rule out ``smoothing'' due to the interpolation used in aligning between different temperatures. The strength of $T_p$ correlations
is also higher than that of gap correlations over the tail of the distribution (1 -- 6~nm).
Although the exact shape of this tail may reflect details of the specific area chosen for study,
the fact that the $T_p$ correlation is consistently higher and the $\delta T_p$ -- $\Delta$ cross correlation also extends to this distance indicates that the process by which
the gap collapses is more sensitive to the surrounding several nanometers than the
process that sets the gap energy scale.
We have performed the first detailed measurements of the temperature
evolution of the superconducting gap for a large number of points in
overdoped $\textrm{Bi}_2\textrm{Sr}_2\textrm{CaCu}_2\textrm{O}_{8+\delta}$. Due to dopants and crystalline disorder, the tendency toward pairing is natrually inhomogeneous. Correspondingly, the superconducting gap
is also inhomogeneous, albeit smoothed out by interactions between regions. The temperature evolution is also inhomogeneous, but we have found that the compared to the gap the magnitude of its variation is more and the length scale of its variation is shorter, that is, the temperature evolution is even more smoothed out. Also, regions of the sample with a small gap magnitude can have this gap survive to
higher temperatures if surrounded by larger gapped regions. We speculate
that the closing of the gap may have to do not just with thermal pair-breaking,
but also with phase fluctuations, so that regions bordered by larger gapped
regions have the phase pinned up to a higher temperature than similar regions
bordered by small gapped regions.
|
\section{Introduction}
Ramsey theory refers to a large body of deep results in mathematics
whose underlying philosophy is captured succinctly by the statement that
``Every large system contains a large well organized subsystem.''
Since the publication of the seminal paper of Ramsey \cite{R30}
in 1930, this subject has grown into one of the most active areas of research within combinatorics,
overlapping variously with number theory, geometry, analysis, logic and computer science.
Given a graph $G$, the {\it Ramsey number} $r(G)$ is defined to be the smallest
natural number $N$ such that, in any two-coloring of the edges of the complete graph
$K_N$ on $N$ vertices, there exists a monochromatic copy of $G$.
Existence of $r(G)$ for all graphs follows from Ramsey's theorem
and determining or estimating these numbers is one of the
central problems in combinatorics (see, e.g., the book \cite{GRS90} for details).
Probably the most famous question in the field is that of estimating the Ramsey number
$r(K_n)$ of the complete graph on $n$ vertices. A classical result of Erd\H{o}s and
Szekeres~\cite{ES35}, which is a quantitative version of Ramsey's
theorem, implies that $r(K_n) \leq 2^{2n}$ for every positive
integer $n$. Erd\H{o}s~\cite{E47} showed using probabilistic
arguments that $r(K_n)> 2^{n/2}$ for $n> 2$. Over the last sixty years, there have been several
improvements on these bounds (see, e.g., \cite{C08}). However,
despite efforts by various researchers, the constant factors in
the above exponents remain the same. Unsurprisingly then, the field has stretched in different directions and the focus
has turned towards the study of the numbers $r(G)$ for general graphs.
One such direction that has become fundamental in its own right is that of
estimating Ramsey numbers for various types of sparse graphs.
In 1975, Burr and Erd\H{o}s \cite{BE75} posed the problem of showing that every
graph $G$ with $n$ vertices and maximum degree $\Delta$ satisfied $r(G) \leq
c(\Delta) n$, where the constant $c(\Delta)$ depends only on $\Delta$. That this is indeed
the case was shown by Chv\'atal, R\"{o}dl, Szemer\'edi and Trotter
\cite{CRST83} in one of the earliest applications of Szemer\'edi's celebrated
regularity lemma \cite{Sz76}. Remarkably, this means that for graphs of fixed
maximum degree the Ramsey number only has a linear dependence on the number of
vertices. However, the use of the regularity lemma only gives a tower-type bound
on $c(\Delta)$, showing that $c(\Delta)$ is at most an exponential tower of 2s with a
height that is itself exponential in $\Delta$.
A remarkable new approach to this problem, which avoids the use of any regularity lemma, was developed by
Graham, R\"{o}dl and Ruci\'nski \cite{GrRoRu}. Their proof shows that
that the function $c(\Delta)$ can be taken to be $2^{c \Delta \log^2 \Delta}$ for some
absolute constant $c$. For bipartite graphs, Graham, R\"{o}dl and Ruci\'nski \cite{GRR01} did even better, showing
that if $G$ is a bipartite graph with $n$ vertices and maximum degree $\Delta$
then $r(G) \leq 2^{c \Delta \log \Delta} n$. They also constructed bipartite graph with $n$ vertices, maximum degree $\Delta$ and
Ramsey number at least $2^{c' \Delta} n$, providing a lower bound on $c(\Delta)$.
Together with Conlon and Fox \cite{C07, FS07, CFS}, we recently further improved these results. Removing the $\log
\Delta$ factor in the exponents, we proved that for general graphs $c(\Delta) \leq 2^{c \Delta \log \Delta}$.
In the bipartite case we achieved an essentially best possible estimate, showing that $r(G) \leq 2^{c \Delta} n$.
Another (somewhat related) problem on Ramsey numbers of general graphs was posed in 1973 by Erd\H{o}s and Graham. Among all graphs with $m$ edges, they wanted to find
the graph $G$ with maximum Ramsey number. Since, the results we mention so far clearly show that sparse graphs have slowly growing Ramsey numbers, one would
probably like to make such a $G$ as dense as possible. Indeed, Erd\H{o}s and Graham \cite{EG} conjectured that among all the graphs with $m={n \choose 2}$ edges (and no
isolated vertices), the complete graph on $n$ vertices has the largest Ramsey number. This conjecture is very difficult and so far there has been no progress on this
problem. Motivated by this lack of progress, in the early 80s Erd\H{o}s \cite{E1} (see also \cite{CG98}) asked whether one could at least show that the Ramsey number of
any graph with $m$ edges is not substantially larger than that of the complete graph with the same size. Since the number of vertices in a complete graph with $m$ edges is a constant
multiple of $\sqrt{m}$, Erd\H{o}s conjectured that $r(G) \leq 2^{c\sqrt{m}}$ for every graph $G$ with $m$ edges and no isolated vertices.
Together with Alon and Krivelevich \cite{AlKrSu} we showed that for all graphs with $m$ edges $r(G) \leq 2^{c\sqrt{m}\log m}$
and also proved this conjecture in the special case when $G$ is bipartite. In this paper we
establish Erd\H{o}s' conjecture in full generality.
\begin{theorem}\label{main}
If $G$ is a graph on $m$ edges without isolated vertices, then $r(G) \leq 2^{250\sqrt{m}}$.
\end{theorem}
This theorem is clearly best possible up to a constant factor in the exponent, since the result of
Erd\H{o}s (mentioned above) shows that a complete graph with $m$ edges has Ramsey number at least $2^{\sqrt{m/2}}$.
The rest of this short paper is organized as follows. In the next section we present several extensions of
the well know results which will be our main tools in establishing Theorem \ref{main}.
The proof of this theorem appears in Section 3. The last section of the paper contains
some concluding remarks and open questions. Throughout the paper, we systematically omit floor and ceiling signs whenever they are not crucial for the sake of clarity of presentation. All
logarithms are in the base $2$. We also do not make any serious attempt to optimize absolute constants in our statements and proofs.
\section{Monochromatic pairs and other tools}
In this section we develop the machinery which we use to establish Theorem \ref{main}.
We need the following important definition.
\begin{definition}
In an edge-coloring of $K_N$, we call an ordered pair $(X,Y)$ of disjoint subsets of vertices monochromatic if all edges in
$X \cup Y$ incident to a vertex in $X$ have the same color.
\end{definition}
Our proof has several ingredients, including extensions of two well known results to monochromatic pairs.
The first uses the original argument of Erd\H{o}s and
Szekeres~\cite{ES35} to show how to find such a pair in every $2$-edge-coloring of a complete graph.
\begin{lemma}\label{EESzek}
For all $k$ and $\ell$, every $2$-edge-coloring of $K_N$, contains a monochromatic pair $(X,Y)$
with $$|Y| \geq {k+\ell \choose k}^{-1}N-k-\ell$$ which is red and has $|X|=k$ or is blue with $|X|=\ell$.
\end{lemma}
\begin{proof}
The proof is by induction on $k+\ell$. The base case when $\min(k,\ell)=0$ is trivial. Let $v$ be an arbitrary vertex of $K_N$.
Then $v$ has either red degree at least $\frac{k}{k+\ell}(N-1)$ or blue degree at least $\frac{\ell}{k+\ell}(N-1)$.
If $v$ has red degree at least $\frac{k}{k+\ell}(N-1)$, then by induction its set of red neighbors contains a pair
$(X,Y)$ with $$|Y| \geq {k-1+\ell \choose k-1}^{-1}\frac{k}{k+\ell}(N-1) -(k-1)-\ell \geq
{k+\ell \choose k}^{-1}N -k-\ell$$ that is monochromatic blue with $|X|=\ell$ (and then we are done)
or monochromatic red with $|X|=k-1$. In the latter case, we can add $v$ to $X$ to obtain a monochromatic red pair
$(X',Y)$ with $X'=X\cup\{v\}$ and $|X'|=k$. A very similar argument, which we omit, can be used to finish the proof in the case when
$v$ has blue degree at least $\frac{\ell}{k+\ell}(N-1)$.
\end{proof}
Although we still do not know how to improve substantially the upper bound for $r(K_n)$,
Erd\H{o}s and Szemer\'edi \cite{ErSzem} showed that this is possible in the case when one color class in the $2$-edge-coloring of $K_N$ is
very sparse or very dense. The edge density of a graph $G$ is the fraction of pairs of distinct vertices of $G$ that are
edges. Our next lemma extends the result of Erd\H{o}s and Szemer\'edi to monochromatic pairs.
\begin{lemma}\label{EESzem}
Let $0<\epsilon\leq 1/7$ and let $t$ and $N$ be positive integers satisfying $t \geq \epsilon^{-1}$ and $N \geq t\epsilon^{-14\epsilon t}$. Then
any red-blue edge-coloring of $K_N$ in which red has edge density $\epsilon$
contains a monochromatic pair $(X,Y)$ with $|X| \geq t$ and $|Y| \geq \epsilon^{14\epsilon t}N$.
\end{lemma}
\begin{proof}
As long as there is a vertex whose red degree is still at least $\epsilon N$ delete it. Since the number of red edges is at most $\epsilon N^2/2$, we have deleted at most $N/2$ vertices.
Let $S$ denote the set of remaining vertices, so $|S| \geq N/2$ and every vertex in $S$ has red degree at most $\epsilon N$.
If $S$ does not contain a blue clique of size $2t$, let $B$ be a maximum blue clique in $S$. Otherwise, let $B$ be a blue clique in $S$ of size $2t$.
Delete all vertices of $S \setminus B$ which have at least $3\epsilon |B|$ red neighbors in $B$, and let $S'$ denote the set of remaining vertices. Since every vertex in $B$ has
red degree at most $\epsilon N$, there are at most $\epsilon N|B|$ red edges from $B$ to $S$, and hence the number of deleted vertices from $S \setminus B$ is at
most $\frac{\epsilon N|B|}{3\epsilon |B|}= N/3$. Using that $7^{14}t \leq N$, we have $|S'| \geq |S \setminus B|-N/3 \geq N/2-2t-N/3 \geq N/7$. For each
subset $R \subset B$ of size $3\epsilon |B|$, let $S_R$ denote the set of vertices in $S'$ whose set of red neighbors in $B$ is contained in $R$.
Note that $S'$ is the union of the sets $S_R$, as each vertex in $S'$ has at most $3\epsilon|B|$ red neighbors in $B$.
Since there are ${|B|\choose 3\epsilon |B|}$ such sets, by the pigeonhole principle we have that there is $R$ for which
$$|S_R| \geq {|B|\choose 3\epsilon |B|}^{-1} |S'| \geq \left(\frac{e}{3\epsilon}\right)^{-3\epsilon|B|} N/7\geq \epsilon^{3\epsilon|B|}N/7\,,$$
where we used the well known fact that ${a \choose b} \leq (ea/b)^b$.
If $|B|=2t$, then let $X=B \setminus R$ and $Y=S_R$. Note that, by definition of $S_R$, all the edges between $B \setminus R$ and $S_R$ are blue. This gives us the monochromatic blue pair $(X,Y)$ with
$|X| \geq (1-3\epsilon)|B| \geq |B|/2=t$ and
$|Y| \geq \epsilon^{6\epsilon t}N/7 \geq \epsilon^{14\epsilon t}N$, so we are done. Hence, suppose that $|B|<2t$.
In this case, note that, there is
no blue clique of size $|R|+1$ in $S_R$. Indeed, such a blue clique $Q$ in $S_R$ together with $(B \setminus R)$ would form a blue clique in $S$ of size larger than $|B|$,
contradicting the maximality of $B$. Apply Lemma \ref{EESzek} with $k=t$ and $\ell=7 \epsilon t \geq 3\epsilon |B|+1=|R|+1$ to the coloring restricted to $S_R$.
Since $S_R$ has no blue clique of size $\ell$, it contains a monochromatic red pair
$(X,Y)$ with $|X|=t$ and
\begin{eqnarray*}
|Y| &\geq& {t+\ell \choose \ell}^{-1}|S_R|-t-\ell \geq
\left(\frac{\ell}{e(t+\ell)}\right)^{\ell}|S_R|-2t \geq
(1.2\epsilon)^{7\epsilon t}|S_R|-2t\\
&\geq& \frac{1.2^7}{7}\epsilon^{7\epsilon t} \epsilon^{3\epsilon|B|}N-2t \geq
\frac{1}{2}\epsilon^{13\epsilon t}N-2\epsilon^{14\epsilon t}N \geq \left(\frac{1}{2}-2\epsilon \right) \epsilon^{13\epsilon t}N \geq \epsilon^{14\epsilon t}N,
\end{eqnarray*}
where we used that $\epsilon \leq 1/7,~ \frac{\ell}{e(t+\ell)}=\frac{7\epsilon}{e(1+7\epsilon)} \geq 1.2\epsilon,~ 1.2^7 \geq 3.5 ,~ t \leq \epsilon^{-1},~ 2t \leq 2\epsilon^{14\epsilon t}N \leq
2\epsilon \cdot \epsilon^{13\epsilon t}N$.
This completes the proof of the lemma.
\end{proof}
Finally, we need some tools developed by
Graham, R\"odl and Ruci\'nski \cite{GrRoRu} to study the Ramsey numbers of sparse graphs
(see also \cite{FS} for simpler proofs and generalizations). We start with some
notation. Let $H$ be a graph with vertex set $V$ and let $U$ be a
subset of $V$. Then we denote by $H[U]$ the subgraph of $H$ induced by $U$ and by
$e(U)$ its number of edges.
The {\em edge density} $d(U)$ of $U$ is defined by
$$d(U)=\frac{e(U)}{{|U| \choose 2}}.$$
Similarly, if $X$ and $Y$ are two disjoint subsets of $V$, then
$e(X,Y)$ is the number of edges of $H$ adjacent to exactly one vertex from
$X$ and one from $Y$ and the density of the pair $(X,Y)$ is defined by
$$d(X,Y)=\frac{e(X,Y)}{|X||Y|}.$$
We say that $H$ is {\em $(\rho, \epsilon)$-sparse} if
there is a pair of disjoint subsets $X, Y \subset V$ with $|X|=|Y| \geq \rho|V|$ and
$d(X,Y)\leq \epsilon$. The following lemma was proved in \cite{GrRoRu} (see Lemma 2). It shows that
if the density between every two sufficiently small disjoint subsets of $H$ is at least $\epsilon$, then $H$ contains
every bounded degree graph $G$ of order proportional to $V(H)$.
\begin{lemma}
\label{l23}
Let $\Delta$ and $n$ be two integers, $\epsilon \leq 1/2$ and $\rho=\epsilon^{\Delta}/(\Delta+1)$.
Let also $G$ be a graph on $n$ vertices with maximum degree at most $\Delta$. If $H$ is a graph of order at least
$(\Delta+1)\epsilon^{-\Delta}n$ which contains no copy of $G$ then $H$
is $(\rho, \epsilon)$-sparse.
\end{lemma}
Note that if the graph $H$ contains no copy of graph $G$ then after finding one sparse pair $(X,Y)$ one can apply the last lemma again to the subgraphs of $H$ induced by sets $X$ and $Y$.
By doing this recursively, it was proved in \cite{GrRoRu} and \cite{FS} that $H$ contains a sparse subset. We use the following statement from \cite{FS}
(see Corollary 3.4).
\begin{lemma}
\label{l24}
Let $0 \leq \epsilon, \rho \leq 1$, $h=\log (2/\epsilon)$ and let $H=(V,E)$ be a graph such that for every subset $U$ of $H$ of size
at least $(\rho/2)^{h-1}|V|$ the induced subgraph $H[U]$ is $(\rho, \epsilon/8)$-sparse. Then $H$ contains a subset $S, |S| \geq 2\rho^h|V|$
with edge density $d(S) \leq \epsilon$.
\end{lemma}
Combining these two lemmas we obtain the last ingredient, which we need for the proof of our main result.
\begin{corollary}
\label{GRR}
Let $G$ be a graph with $n$ vertices, maximum degree $\Delta$ and let $\epsilon \leq 1/8$.
If $H$ has $N \geq \epsilon^{4\Delta\log \epsilon}n$ vertices and does not contain a copy of $G$, then it has a
subset $S$ of size $|S| \geq \epsilon^{-4\Delta \log \epsilon}N$ with edge density $d(S) \leq \epsilon$.
\end{corollary}
\begin{proof}
Let $\rho=(\epsilon/8)^{\Delta}/(\Delta+1)$, $h=\log (2/\epsilon)$ and let $U$ be a subset of $H$ of size
$(\rho/2)^{h-1}N$. Using that $\epsilon \leq 1/8$ it is rather straightforward to check that $n \leq (\rho/2)^h N$ and therefore
$|U| \geq \rho^{-1} n =(\Delta+1) (\epsilon/8)^{-\Delta}n$. Moreover the induced subgraph $H[U]$ contains no copy of $G$,
and therefore satisfies the conditions of Lemma \ref{l23} (with $\epsilon/8$ instead of $\epsilon$). Therefore $H[U]$ is
$(\rho, \epsilon/8)$-sparse. Thus we can apply Lemma \ref{l24} to $H$ and find a subset $S, |S| \geq 2\rho^h N $
with density $d(S) \leq \epsilon$. Since $2\rho^h \geq \epsilon^{-4\Delta \log \epsilon}$, this completes the proof.
\end{proof}
\section{Proof of the main result}
We start by describing the idea of the proof. Suppose we have a red-blue edge-coloring of $K_N$ without any monochromatic copy of a certain graph $G$ with
$m$ edges. Using Lemma \ref{EESzek}, we find a monochromatic pair $(X,Y)$ (say in red), where the size of $X$ is of order $\sqrt{m}$.
Split the graph $G$ into two parts $A$ and $G'=G\setminus A$, where $A$ is a set of $|X|$ vertices of the largest degree in $G$. Then
the graph $G'$ will have maximum degree bounded by $2m/|X|=O(\sqrt{m})$. Embed $A$ into $X$ and try to find a red copy of $G'$ in $Y$. If $Y$ has no red copy of $G'$,
use Corollary \ref{GRR} to conclude that it has a relatively large subset $S$ with few red edges. Now we can apply Lemma \ref{EESzem} to $S$ to
find a new monochromatic pair $(X',Y')$ in blue with the
following
property.
The size of $X'$ will be considerably larger than the size of $X$. On the other hand, the size of
$Y'$ will not decrease substantially compared with the
size of $Y$. The proof will follow by repeated application of this procedure, since at some point the size of the
monochromatic clique $X'$ will be larger than the number of vertices in $G$. The following key lemma gives a precise formulation of the amplification step.
\begin{lemma}\label{pairslemma}
Let $G$ be a graph with $m$ edges and without isolated vertices and suppose $27 \leq \alpha \leq \frac{1}{8}\log^3 m$.
If a red-blue edge-coloring of a complete graph on $N$ vertices has no monochromatic copy
of $G$ and contains a monochromatic pair $(X,Y)$ with $|X| \geq \alpha \sqrt{m}$ and $|Y| \geq 2^{125\alpha^{-1/3}\sqrt{m}}$,
then it also contains a monochromatic pair $(X',Y')$ with $|X'| \geq 2^{2\alpha^{1/3}}\sqrt{m}$
and $|Y'| \geq 2^{-120\alpha^{-1/3}\sqrt{m}}|Y|$.
\end{lemma}
\begin{proof}
Without loss of generality, assume that the color of the monochromatic pair $(X,Y)$ is red.
Let $G'$ be the induced subgraph of $G$ formed by deleting the $|X|$ vertices of $G$ of
largest degree. As $G$ has $m$ edges, it has $n \leq 2m$ vertices and the maximum degree of $G'$ satisfies $\Delta(G') \leq \frac{2m}{|X|}=\frac{2}{\alpha}\sqrt{m}$.
The coloring restricted to $Y$ contains no monochromatic red copy of $G'$ as, otherwise, together with $X$,
we would get a monochromatic copy of $G$. Let $\epsilon=2^{-3\alpha^{1/3}}$ and let $t=2^{2\alpha^{1/3}}\sqrt{m}$.
Since $2^{\alpha^{1/3}}\leq \sqrt{m}$ we have that $t \geq \epsilon^{-1}$.
Also note that, since $27 \leq \alpha \leq \frac{1}{8}\log^3 m$, we have that $42\alpha^{1/3}2^{-\alpha^{1/3}}\leq 48\alpha^{-1/3}$ and
$2^{5\alpha^{-1/3}\sqrt{m}} \geq 2^{10\sqrt{m}/\log m} \geq m^{3/2} \geq 2^{2\alpha^{1/3}}\sqrt{m}=t$.
Applying Corollary \ref{GRR} to the red graph restricted to $Y$, we find a subset $S \subset Y$ with
\begin{eqnarray*}
|S| &\geq& \epsilon^{-4\Delta(G') \log \epsilon}|Y| \geq
\Big(2^{-3\alpha^{1/3}}\Big)^{-4 \cdot (2\alpha^{-1}\sqrt{m}) \cdot(-3 \alpha^{1/3})}|Y| = 2^{-72\alpha^{-1/3}\sqrt{m}}|Y|\\
&\geq&
2^{53\alpha^{-1/3}\sqrt{m}}>m^3 \geq n
\end{eqnarray*}
such that the red density in $S$ is at most $\epsilon$.
Then, the size of $S$ satisfies
$$|S| \geq 2^{53\alpha^{-1/3}\sqrt{m}} \geq 2^{5\alpha^{-1/3}\sqrt{m}}\cdot 2^{48\alpha^{-1/3}\sqrt{m}}
\geq t\, 2^{42\alpha^{1/3}2^{-\alpha^{1/3}}\sqrt{m}}=t \epsilon^{-14\epsilon t},$$
and we can apply Lemma \ref{EESzem} to $S$. By this lemma, $S$ contains a monochromatic pair
$(X',Y')$ with $|X'|=t$ and $|Y'| \geq \epsilon^{14\epsilon t}|S|$. To complete the proof, recall that $|S| \geq 2^{-72\alpha^{-1/3}\sqrt{m}}|Y|$, and therefore
$$|Y'| \geq \epsilon^{14\epsilon t}|S| \geq 2^{-48\alpha^{-1/3}\sqrt{m}}|S| \geq 2^{-120\alpha^{-1/3}\sqrt{m}}|Y|.$$
\end{proof}
\vspace{0.1cm}
\noindent {\bf Proof of Theorem \ref{main}.}
Let $G$ be a graph with $m$ edges and without isolated vertices. Note that $G$ has at most $2m$ vertices.
Suppose for contradiction that there is a red-blue edge-coloring of $K_N$ with $N=2^{250\sqrt{m}}$ which contains no
monochromatic copy of $G$. Since, as was mentioned in the introduction, $r(K_{2m}) \leq 2^{4m}$ we have that
$m \geq 60^2$. Applying Lemma \ref{EESzek} with $k=\ell=27\sqrt{m}$, we find a monochromatic pair $(X_1,Y_1)$ with $|X_1| \geq 27\sqrt{m}$ and
$$|Y_1| \geq {k+\ell \choose k}^{-1}N-k-\ell \geq 4^{-27\sqrt{m}}N= 2^{196\sqrt{m}}.$$
Define $\alpha_1=27$ and $\alpha_{i+1}=2^{2\alpha_i^{1/3}}$. An easy computation shows that $\alpha_{i+1}\geq (4/3)^3 \alpha_i$ for all $i$ and therefore
$\alpha_i^{-1/3} \leq 3^{-1}(3/4)^{i-1}$. In particular this implies that
$$\sum_{j =1}^i \alpha_j^{-1/3} \leq \frac{1}{3}\sum_{j=0}^{i} (3/4)^{-j}=
\frac{1}{3}\sum_{j \geq 0} (3/4)^{-j}- \frac{1}{3}\sum_{j \geq i+1} (3/4)^{-j} \leq 4/3-4\alpha_{i+1}^{-1/3}.$$
Since the red-blue edge-coloring has no monochromatic copy of $G$, we can repeatedly apply Lemma \ref{pairslemma}.
After $i$ iterations, we have a monochromatic pair $(X_{i+1},Y_{i+1})$ with $|X_{i+1}| \geq \alpha_{i+1}\sqrt{m}$ and
\begin{eqnarray*}
|Y_{i+1}| &\geq& 2^{-120\alpha_i^{-1/3}\sqrt{m}}|Y_i| \geq 2^{-120\sqrt{m} \sum_{j= 1}^i \alpha_j^{-1/3}}|Y_1| \geq
2^{196\sqrt{m}} 2^{-120\sqrt{m}(4/3-4\alpha_{i+1}^{-1/3})}\\
&\geq& 2^{(36+480\alpha_{i+1}^{-1/3})\sqrt{m}}.
\end{eqnarray*}
Hence we can continue iterations until the first index $i$ such that $\alpha_{i} \geq \frac{1}{8}\log^3 m$. Then
for $\alpha = (\log m/2)^3 \geq 5^3$ we have a monochromatic pair $X=X_i, |X| \geq \alpha \sqrt{m}$ and $Y=Y_i, |Y| \geq 2^{36\sqrt{m}} \geq
2^{125\alpha^{-1/3}\sqrt{m}}$. Then applying
Lemma \ref{pairslemma} one more time we obtain a monochromatic pair $(X',Y')$ with $|X'| \geq 2^{2\alpha^{1/3}}\sqrt{m}=m^{3/2} \geq 2m$
and we can embed $G$ (which has at most $2m$ vertices) into the monochromatic clique $X'$, a contradiction.
\hfill $\Box$
\section{Concluding remarks}
A graph is {\it $d$-degenerate} if every induced subgraph of it has a vertex
of degree at most $d$. Notice that graphs with maximum degree $d$
are $d$-degenerate. This notion nicely captures the concept of
sparse graphs as every $t$-vertex subgraph of a $d$-degenerate graph
has at most $td$ edges. (Indeed, remove from the subgraph a vertex
of minimum degree, and repeat this process in the remaining
subgraph.) Burr and Erd\H{o}s \cite{BE75} conjectured that, for each
positive integer $d$, there is a constant $c(d)$ such that $r(H)
\leq c(d)n$ for every $d$-degenerate graph $H$ on $n$ vertices. This
well-known and difficult conjecture is a substantial generalization
of the results on Ramsey numbers of bounded-degree
graphs (mentioned in introduction) and progress on this problem was made only recently.
Improving an earlier polynomial bound of \cite{KoRo1}, we obtained, together with Kostochka,
the first nearly linear bound on the Ramsey numbers of $d$-degenerate
graphs. In \cite{KoSu} we proved that such graphs satisfy $r(H)
\leq c(d)n^{1+\epsilon}$ for any fixed $\epsilon>0$. The best current estimate, showing that
$r(H) \leq 2^{c(d) \sqrt{\log n}} n$, is due to Fox and Sudakov \cite{FS09}.
In spite of this progress, the Burr-Erd\H{o}s conjecture is still open even for the very special case of $2$-degenerate
graphs. However, it seems plausible that $r(H) \leq 2^{c d}n$ (for some constant $c$) holds in general
for every $d$-degenerate graph $H$ on $n$ vertices. Such an estimate would be a far reaching generalization of the
results about Ramsey numbers of bounded degree graphs and also of Theorem \ref{main}. Indeed, it is easy to check that
every graph with $m$ edges is $\sqrt{2m}$-degenerate.
Finally, we would like to stress that the proofs given in this paper are highly
specific to the 2-color case.
The $k$-color Ramsey number $r_k(G)$ is the least positive integer $N$ such
that every $k$-coloring of the edges of a complete graph
$K_N$, contains a monochromatic copy of the graph $G$ in one of the colors.
It would be interesting to understand, for $k \geq 3$, the order of magnitude of the $k$-color Ramsey number of a
graph with $m$ edges.
Also for bounded degree graphs the best results that are known in the $k$-color
case are much worse than for $2$ colors. For example (see \cite{FS07}), for a
graph $G$ on $n$ vertices with maximum degree $\Delta$,
we only know the bound $r_k(G) \leq 2^{c(k) \Delta^2} n$. Improving it to $r_k(G) \leq 2^{c(k) \Delta^{1+o(1)}} n$ would be of considerable
interest.
\vspace{0.3cm}
\noindent
{\bf Acknowledgment.} I would especially like to thank Jacob Fox for valuable comments and
for writing a draft of the proof of the main theorem. I also had many stimulating discussions with David Conlon and want to thank
him for reading very carefully an early version of this manuscript.
|
\section{Introduction.}
A curve endowed with a frame, called {\it a framed curve},
in a space-form plays important roles in topology, geometry and singularity theory.
For example, as is well-known, the self-linking number in
$3$-space is defined via framing (\cite{Pohl}).
The fundamental theory of curves is formulated via osculation framing. Surface boundaries have adapted framings, etc.
Two kinds of frames are considered in this paper; adapted frames and osculating frames
from the viewpoint of duality. Then we classify the singularities of envelopes associated to framed curves.
The singularities of envelopes in $E^3$ was studied in \cite{Ishikawa6} to apply to
the flat extension problem of a surface with boundary.
The problem on extensions by tangentially degenerate surfaces motivates to study the envelopes
associated to framings on curves in a space form.
In this paper we consider framed curves in
$X = E^{n+1}$, Euclidean space, $S^{n+1}$, the sphere or $H^{n+1}$, the hyperbolic
space of dimension $n+1$, and understand commonly in terms of projective geometry.
Actually we work with the models
$$
S^{n+1} = \{ x \in \mathbf{R}^{n+2} \mid x^2 = 1 \}, \quad
H^{n+1} = \{ x \in \mathbf{R}^{1, n+1} \mid x^2 = - 1, \ x_0 > 0
\},
$$
where $\mathbf{R}^{1, n+1} = \mathbf{R}^{n+2}_1 = \{ (x_0, x_1, \dots, x_{n+1}) \}$ is the Minkowski space of
index $(1, n+1)$ (See for instance \cite{IPS}\cite{Harvey}). The inner product in
$\mathbf{R}^{1, n+1}$ is defined by $x\cdot y = - x_0y_0 + \sum_{i=1}^{n+1} x_iy_i$.
Moreover we identify Euclidean space $E^{n+1}$ with $\{ x \in \mathbf{R}^{n+2} \mid x_0 = 1\} \subset \mathbf{R}^{n+2}$
if necessary.
Let $\gamma : I \to X$ be a $C^\infty$ immersion from an interval or a circle $I$.
In general, we mean by a {\it framing} of the immersed curve
$\gamma$, an oriented orthonormal frame
$(e_1, e_2, \dots, e_{n+1})$ along $\gamma$.
We always
pose the condition that
$e_{n+1}$ is orthogonal to the velocity vector $\gamma'$.
Then the unit normal vectors $e_{n+1}$
provide a $1$-parameter family of tangent hyperplanes to $\gamma$
and its envelope $E(\gamma)$.
In particular, in three dimensional case ($n = 2$),
if a framed curve $\gamma$ is given, then we have a $1$-parameter family of planes
and its envelope surface $E(\gamma)$ in three space.
For a $1$-parameter family of framed curves $\gamma_\lambda$, we have the $1$-parameter family of
envelopes $E(\gamma_\lambda)$. Then we will show
\begin{Th}
\label{envelope}
Let $\gamma_\lambda$ be a generic $1$-parameter family of framed
curves in $E^3, S^3$ or $H^3$.
Then the local singularity in the associated envelope $E(\gamma_\lambda)$
is given by one of following $5$-classes:
{\rm (I)} the cuspidal edge, {\rm (II)} the swallowtail,
{\rm (III)} the cuspidal breaks, {\rm (IV)} the cuspidal butterfly,
and {\rm (V)} the full-folded-umbrella.
\stepcounter{Cor
In particular, the list of singularities (diffeomorphism classes), is the same for all of three geometries.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=5truecm, height=3truecm, clip, bb=0 520 500 850]{swallowtail10.pdf}
\hspace{2truecm}
\includegraphics[width=5truecm, height=3truecm, clip, bb=100 440 520 800]{beak10.pdf}
\\
the swallowtail
and
the cuspidal beaks
\end{center}
\begin{center}
\includegraphics[width=5truecm, height=3truecm, clip, bb=0 560 440 800]{butterfly10.pdf}
\hspace{2truecm}
\includegraphics[width=5truecm, height=3truecm, clip, bb=10 400 510 800]{FFU10.pdf}
\\
the cuspidal butterfly
and
the full-folded-unbrella
\end{center}
\end{figure}%
The cuspidal edges and the swallowtails appear generically and stably.
The swallowtails can appear in isolated positions on the envelope in any moment $\lambda$.
The cuspidal breaks, the cuspidal butterflies, or the full-folded-umbrellas
appear in isolated positions momentarily at isolated value $\lambda$.
Along the parameter $\lambda$, both the cuspidal breaks and butterflies bifurcate within wavefronts (\cite{Arnold1}). However we see later they have different character in our theory.
The full-folded-umbrella is not a wavefront (image of a non-singular Legendre submanifold), but,
a frontal surface (image of a singular Legendre variety).
The singularities of envelopes are closely related to singularities of tangent developables of curves.
Tangent developables are flat in $E^3$. However they are not flat but "extrisically flat" or tangentially degenerate
in $S^3$ and $H^3$ (cf. \cite{AG}\cite{KRUSUY}).
In this paper the notion of {\it types} $(a_1, a_2, a_3)$ for a curve-germ is introduced and
the cuspidal edge, (resp. the swallowtail, the cuspidal break, the cuspidal butterfly) is obtained
as the tangent developable of a curve of type $(1, 2, 3)$ (resp. $(2, 3, 4)$, $(1, 3, 4)$, $(3, 4, 5)$).
The cuspidal break is called also {\it Mond surface} \cite{Ishikawa4}. (See also \cite{Mond1}\cite{Mond2}).
We have adopted the notations in \cite{IS}.
We remark that the cuspidal butterflies bifurcate within tangent developables, however,
the cuspidal breaks (Mond surfaces) do not. In fact we observe the Mond surface is stable
for the deformations of curves with {\it osculating frames}.
The full-folded-umbrella contains the tangent developable of a curve of type $(1, 2, 4)$.
Each singularity mentioned above is given by the generating family
$$
F(t, x_1, x_2, x_3) =
\frac{t^{a_3}}{a_3!} + x_1\frac{t^{a_3 - a_1}}{(a_3 - a_1)!} + x_2\frac{t^{a_3 - a_2}}{(a_3 - a_2)!} + x_3 = 0
$$
of (totally geodesic) planes, where the normal form of the envelope is
given by $\{ (x_1, x_2, x_3) \mid F = \frac{\partial F}{\partial t} = 0 {\mbox{\rm { for some
}}} t \}$ (\cite{Ishikawa1}).
In this paper two kinds of frames are involved:
one is an {\it adapted frame} of $\gamma$ which satisfies just the condition $e_1 = \gamma'$,
the unit velocity vector field, or the differential by the arc-length parameter.
Then $e_{n+1}$ is orthogonal to $\gamma'$. For the classification problem of envelops just that
$e_{n+1}$ is orthogonal to $\gamma'$ is essential.
Another is the Frenet-Serre frame of $\gamma$ along ordinary points
where the derivatives
$
\gamma'(t), \gamma''(t), \dots, \gamma^{(n)}(t),
$
are linearly independent. Then our main idea is to introduce two kinds of distributions, or differential systems,
on flag manifolds and regard framed curves as integral curves to those distributions.
Bifurcations of wavefronts based on Legendre singularity theory are established by
Arnold-Zakalyukin's theory (\cite{Arnold1}\cite{Arnold2}\cite{Zakalyukin1}\cite{Zakalyukin2}).
The application of singularity theory to differential geometry has been developed by many authors
(see for instance \cite{BG}).
The singularity theory based geometry of submanifolds in hyperbolic space $H^{n+1}$ is
initiated by Izumiya et al. (\cite{IPS}\cite{IPS2}\cite{IPT}).
The Legendre duality developed in
\cite{Izumiya}\cite{CI} enables us to unify the theory of framed curves in any space form as describes in this paper.
In \S \ref{Legendre duality.},
we recall Legendre duality (see \cite{Bruce}\cite{IMo}\cite{CI})
within the level we need in this paper.
We understand the duality in the framework of moving frames and flags in \S \ref{Moving frames and flags.}.
After touching with non-oriented flags in \S \ref{Reduced Legendre duality.}, we introduced
two distributions in \S \ref{Pseudo-contact and canonical distributions.}. This is very essential to study the bifurcation problem of envelopes in this paper.
In \S \ref{Osculating flags on curves of finite type.}, the notion of type of curves is introduced
and that of osculating flags are considered. Two kinds of framed curves are regarded as
integral curves to two kinds of distributions. Then
we prove codimension formulae for framed curves in \S \ref{Integral curves and codimension formula.},
which implies the classification results of singularities of envelopes including Theorem\ref{envelope},
in \S \ref{Singularities of envelopes and their bifurcations.}.
\section{Legendre duality.}
\label{Legendre duality.}
Though in this paper we mainly treat curves in Riemannian spaces $X = E^{n+1}, S^{n+1}, H^{n+1}$,
regarding the duality, naturally we work in other spaces as well.
In particular we are led to consider {\it de Sitter space}
$$
S^{1, n} = \{ x \in \mathbf{R}^{1, n+1} \mid x^2 = 1\},
$$
which is a semi-Riemannian manifold,
since any vector
of an frame $(e_1, \dots, e_{n+1})$ along
a curve in $H^{n+1}$ belongs to $S^{1, n}$.
We regard $\widehat{\gamma} = e_{n+1}$ a curve in $Y = S^{n+1}$
(resp. in $Y = S^{1, n}$)
if $X = S^{n+1}$ (resp. $X = H^{n+1}$).
In Euclidean case, we set
$\widehat{\gamma} = (- \gamma\cdot e_{n+1}, e_{n+1})$
and regard it as a curve in $Y = \mathbf{R}\times S^{n+1}$.
We call $\widehat{\gamma}$ the {\it frame dual} to $\gamma$.
Then, in any case, the \lq\lq type" of the curve $\widehat{\gamma}$ in $Y$
describes the singularities of the envelope
$E(\gamma)$ in $X$.
We denote by $Z = {\widetilde{{\mbox {\rm Gr}}}}(n, TX)$ the manifold of oriented tangent hyperplanes of $X$
and by $\pi_1 : Z \to X$ the projection which maps a hyperplane $\Pi \subset T_xX$ to $x \in X$.
A framed curve $\gamma : I \to X$ with the framing $(e_1, \dots, e_{n+1})$
lifts to a curve $\widetilde{\gamma} : I \to Z$ which is
defined by $\widetilde{\gamma}(t) = \langle e_1(t), \dots, e_n(t) \rangle_{\mathbf{R}}$.
In each of three cases, $Z$ is identified with the unit tangent bundle $T_1X$ via the metric,
actually with
$T_1E^{n+1} = E^{n+1}\times S^n$,
$$
\begin{array}{rcl}
T_1S^{n+1} & = & \{ (x, y) \in S^{n+1}\times S^{n+1} \mid x\cdot y = 0 \}, \quad {\mbox{\rm and}}, \\
T_1H^{n+1} & = & \{ (x, y) \in H^{n+1}\times S^{1, n} \mid x\cdot y = 0 \}.
\end{array}
$$
Then, under the above identification,
the lifting $\widetilde{\gamma} : I \to Z$ is given by $\widetilde{\gamma}(t) = e_{n+1}(t)$ (\cite{IM2}).
Consider the contact structure on $Z$:
the one-form $\theta = v\cdot dx$ restricted to $Z = E^{n+1}\times S^n$,
$\theta = y\cdot dx$ restricted to $Z = T_1S^{n+1}$ or $T_1H^{n+1}$,
is a contact form on $Z$.
In elliptic or hyperbolic case, let $\pi_2 : Z \to Y$ be the projection defined by $\pi_2(x, y) = y$.
In Euclidean case, let $\pi_2 : Z \to Y$ be the projection
defined by $\pi_2(x, y) = (-x\cdot y, \ y) (x \in E^{n+1}, y \in S^n)$.
Then we see both $\pi_1 : Z \to X$ and $\pi_2 : Z \to Y$ are Legendre fibrations.
Suppose the framing of $\gamma : I \to X$ satisfies the condition $e_1 = \gamma'$.
Then $e_{n+1}$ is normal to $\gamma'$.
And then we see that the lifting
$\widetilde{\gamma} : I \to Z$ of $\gamma : I \to X$ turns to be
{\it integral} in the sense that ${\widetilde{\gamma}}^* \theta = 0$.
The lifting $\widetilde{\gamma} : I \to Z$ of a framed immersion
$\gamma : I \to X$ defines
a \lq\lq sub-front" $\widehat{\gamma} =
\pi_2\circ\widetilde{\gamma} : I \to Y$ possibly with singularities,
in the sense that the integral lifting $\widetilde{\gamma}$ with respect to $\pi_2$
is attached to the just parametrised curve $\widehat{\gamma}$ in $Y$.
Note that, in the case $X = H^{n+1}$, $Z = T_1 H^{n+1}$ is identified with
$T_{-1} S^{1, n}$, the manifold of tangent vectors $v \in T_y S^{1, n}$ with $v^2 = -1$
to the semi-Riemannian manifold $S^{1, n}$ (\cite{IM2}).
As the model of duality, we do have the projective duality (\cite{O.P.Scherbak1}\cite{IMo}); we set
$$
Z \ = \ {\mathcal I}_{n+2} \ := \ \{ ([x], [y]) \in P^{n+1}\times P^{n+1*} \mid x\cdot y = 0 \}.
$$
Here $P^{n+1*}$ is the dual projective space and
$\cdot$ means the natural paring. The contact structure on ${\mathcal I}_{n+2}$
is defined by $dx\cdot y = x\cdot dy = 0$ (\cite{IMo}).
The projections $\pi_1 : {\mathcal I}_{n+2} \to X = P^{n+1}, \pi_2 : {\mathcal I}_{n+2} \to Y = P^{n+1*}$
are both Legendre fibrations.
The following fact is basic to unify our treatment:
\begin{Prop}
\label{flatness}
{\rm (\cite{IM}\cite{IM2})}
All Legendre double fibrations $X \longleftarrow Z \longrightarrow Y$ constructed above
are locally isomorphic to each other. In particular each of them is locally isomorphic to
the double fibration of the projective duality $P^{n+1} \longleftarrow {\mathcal I}_{n+2} \longrightarrow P^{n+1*}$.
\stepcounter{Th
The proof of Proposition \ref{flatness} is proved naturally via the underlying
flag structure that we are going to explain.
\section{Moving frames and flags.}
\label{Moving frames and flags.}
For a framed curve $\gamma : I \to X$ with a frame
$(e_1, \dots, e_{n+1})$, naturally there is associated a
\lq\lq moving frame" $\widetilde{\gamma} : I \to G$, for each case,
in a Lie subgroup of ${\mbox {\rm GL}}_+(n+2, \mathbf{R})$, regular matrices with positive determinant.
If $X = E^{n+1}$, then we set $e_0(t) = \gamma(t) \in E^{n+1}$,
and we have the moving frame
$\widetilde{\gamma} = (e_0, e_1, \dots, e_{n+1}) : I \to G = {\mbox{\rm{Euc}}}(E^{n+1})
\subset {\mbox {\rm GL}}_+(n+2, \mathbf{R})$ in
the group of orientation preserving Euclidean motion on $E^{n+1}$.
If $X = S^{n+1}$, then we set $e_0(t) = \gamma(t) \in S^{n+1}$,
and we have the moving frame
$\widetilde{\gamma} = (e_0, e_1, \dots, e_{n+1}) : I \to G = SO(n+2) \subset {\mbox {\rm GL}}_+(n+2, \mathbf{R})$.
If $X = H^{n+1}$, then we set $e_0(t) = \gamma(t) \in H^{n+1}$,
and we have the moving frame
$\widetilde{\gamma} = (e_0, e_1, \dots, e_{n+1}) : I \to G = SO(1,n+1) \subset {\mbox {\rm GL}}_+(n+2, \mathbf{R})$.
In any of three cases, the frame manifold $G$ is identified with an open subset of
the oriented flag manifold $\widetilde{\mathcal F}_{n+2}$
consisting of oriented complete flags
$$
V_1 \subset V_2 \subset \cdots \subset V_{n+1} \subset \mathbf{R}^{n+2}
$$
in $\mathbf{R}^{n+2}$. For each $g = (e_0, e_1, \dots, e_{n+1}) \in {\mbox {\rm GL}}_+(n+2, \mathbf{R})$,
we set the oriented subspace
$$
V_i = \langle e_0, e_1, \dots, e_{i-1}\rangle_{\mathbf{R}} \subset \mathbf{R}^{n+2}, \ (1 \leq i \leq n+1).
$$
This induces an open embedding $G \to \widetilde{\mathcal F}_{n+2}$.
Note that $\widetilde{\mathcal F}_{n+2}$ is the quotient
space of ${\mbox {\rm GL}}_+(n+2, \mathbf{R})$ by upper triangular matrices, and
$\dim G = \dim \widetilde{\mathcal F}_{n+2} = \frac{(n+1)(n+2)}{2}$.
Moreover note that the inner product restricted to each $V_i$ is non-degenerate.
Therefore $G$ is embedding in non-degenarate flags $\widetilde{\mathcal F}_{n+2}^{0} \subset \widetilde{\mathcal F}_{n+2}$ consisting of flags $(V_1, \dots, V_{n+1})$ where
the inner product restricted to each $V_i$ is non-degenerate. Remark that
$\widetilde{\mathcal F}_{n+2}^{0}$ is open dense in $\widetilde{\mathcal F}_{n+2}$.
However note that in \cite{Izumiya}\cite{CI}\cite{IS}, more general framings are considered to treat also
{\it the light cone} in Minkowski space.
Thus, for a framed curve $\gamma : I \to X$ in $X = E^{n+1}, S^{n+1}, H^{n+1}$,
with the frame $(e_1, \dots, e_{n+1})$,
we have the \lq\lq framed curve" $\widetilde{\gamma} : I \to \widetilde{\mathcal F}_{n+2}^{0}$
by setting
$$
V_i(t) = \langle e_0(t), e_1(t), \dots, e_{i-1}(t) \rangle_{\mathbf{R}} \subset \mathbf{R}^{n+2}, \ (1 \leq i \leq n+1).
$$
Then $\widetilde{\gamma}$ is a lifting of
$\gamma$ for the projection $\pi_1 : \widetilde{\mathcal F}_{n+2}^{0} \to \widetilde{{\mbox {\rm Gr}}}(1, \mathbf{R}^{n+2})$
to Grassmannian of oriented lines in $\mathbf{R}^{n+2}$. Note that
there is the natural open embedding $X \subset \widetilde{{\mbox {\rm Gr}}}(1, \mathbf{R}^{n+2})$ in each of three cases.
\section{Reduced Legendre duality.}
\label{Reduced Legendre duality.}
The vector space $\mathbf{R}^{n+2}$ have $\mathbf{Z}/2\mathbf{Z}$-action defined by
$x \mapsto - x$. To describe the duality we are going to use properly,
it is natural to take quotient and set
$$
\begin{array}{cc}
AG(n, n+1) \ := \ \{ (r, y) \in \mathbf{R}\times S^n \}/(\mathbf{Z}/2\mathbf{Z}), \quad
P^{n+1} \ := \ \{ x \in \mathbf{R}^{n+2} \mid x^2 = 1\}/(\mathbf{Z}/2\mathbf{Z}),
\vspace{0.2truecm}
\\
H^{n+1} \ := \ \{ x \in \mathbf{R}^{1, n+1} \mid x^2 = -1\}/(\mathbf{Z}/2\mathbf{Z}),
{\mbox{\rm
\ and, \ }}
P^{1, n} \ := \ \{ x \in \mathbf{R}^{1, n+1} \mid x^2 = 1\}/(\mathbf{Z}/2\mathbf{Z}).
\end{array}
$$
We call $P^{n+1}$ the {\it elliptic space} and $P^{1, n}$ the {\it reduced de Sitter space}.
Remark that $AG(n, n+1)$ is identified with the set of affine non-oriented hyperplanes in $E^{n+1}$.
We regard $P^{n+1}$ (resp. $P^{1, n}$) as the double-quotient of the sphere $S^{n+1}$ (resp.
$S^{1, n}$) with the induced metric.
Set
$X = E^{n+1}, P^{n+1}, H^{n+1}$ in Euclidean, elliptic, hyperbolic case,
respectively. Then we set $Y = AG(n, n+1), P^{n+1}, P^{1, n}$ respectively.
We consider the incidence manifold in each geometry:
$$
Z \ := \ \{ ([x], [y]) \in X\times Y \mid x\cdot y = 0 \},
$$
for elliptic and hyperbolic cases,
and
$$
Z \ := \ \{ (x, [r, y]) \in X\times Y \mid x\cdot y + r = 0 \},
$$
for Euclidean case.
In each case, $Z$ is regarded naturally as an open subset of $PT^*X$
and is endowed with the standard contact structure.
Then the double fibrations
$\pi_1 : Z \to X$ and $\pi_2 : Z \to Y$ are Legendre. Moreover all
Legendre double fibrations $X \longleftarrow Z \longrightarrow Y$ are
locally isomorphic
to the projective duality $P^{n+1} \longleftarrow {\mathcal I}_{n+2} \longrightarrow P^{n+1*}$ as in
Proposition \ref{flatness}.
Let ${\mathcal F}_{n+2}$ be the manifold of non-oriented complete flags
$$
V_1 \subset V_2 \subset \cdots \subset V_{n+1} \subset \mathbf{R}^{n+2},
$$
consisting of vector subspaces $V_i$ of dimension $i$ in $\mathbf{R}^{n+2}$.
The forgetful mapping $\pi :
\widetilde{\mathcal F}_{n+2} \to {\mathcal F}_{n+2}$ forms a covering of order
$2^{n+1}$.
For a framed curve in a reduced space, the lifting is a curve in a non-oriented flag manifold.
\section{Pseudo-contact and canonical distributions.}
\label{Pseudo-contact and canonical distributions.}
We will consider two classes of curves in the frame manifold $\widetilde{\mathcal F}_{n+2}$,
by introducing two kinds of distributions
$\widetilde{\mathcal C} \subset \widetilde{\mathcal D} \subset T\widetilde{\mathcal F}_{n+2}$.
Denote by $\pi_i : \widetilde{\mathcal F}_{n+2} \to \widetilde{{\mbox {\rm Gr}}}(i, \mathbf{R}^{n+2})$ the projection to
Grassmannian of oriented $i$-planes in $\mathbf{R}^{n+2}$ defined by
$
\pi_i(V_1, \dots, V_i, \dots, V_{n+1}) = V_i.
$
Then we define,
for $v \in T\widetilde{\mathcal F}_{n+2}$,
$v \in \widetilde{\mathcal D}_{(V_1, \dots, V_{n+1})}$ if
${\pi_1}_*(v) \in T\widetilde{{\mbox {\rm Gr}}}(1, V_{n+1}) (\subset T\widetilde{{\mbox {\rm Gr}}}(1, \mathbf{R}^{n+2}))$, while
$v \in \widetilde{\mathcal C}_{(V_1, \dots, V_{n+1})}$ if
${\pi_i}_*(v) \in T\widetilde{{\mbox {\rm Gr}}}(i, V_{i+1}) (\subset T\widetilde{{\mbox {\rm Gr}}}(i, \mathbf{R}^{n+2})), (1 \leq i \leq n)$.
We call the distribution $\widetilde{\mathcal D}$ {\it pseudo-contact distribution} and
$\widetilde{\mathcal C}$ {\it canonical distribution}.
Note that
the rank of $\widetilde{\mathcal C}$ (resp.
$\widetilde{\mathcal D}$) is $n+1$ (resp. $\frac{(n+1)(n+2)}{2} - 1$) in
$T\widetilde{\mathcal F}_{n+2}$.
Both $\widetilde{C}$ and $\widetilde{D}$ are bracket generating; in fact, $n$-th bracket $\widetilde{\mathcal C}^n$
of $\widetilde{\mathcal C}$ coincides with $\widetilde{\mathcal D}$.
Denote by $\widetilde{\mathcal I}_{n+2}$ the flag manifold
consisting of flag $V_1 \subset V_{n+1} \subset \mathbf{R}^{n+2}$ with an oriented line $V_1$ and
an oriented hyperplanes $V_{n+1}$.
Consider the canonical projection
$
\pi_{1, n+1} : \widetilde{\mathcal F}_{n+2} \longrightarrow \widetilde{\mathcal I}_{n+2}
$
defined by $\pi_{1, n+1}(V_1, V_2, \dots, V_{n+1}) = (V_1, V_{n+1})$.
Then we have $\widetilde{\mathcal D} = (\pi_{1, n+1})_*^{-1}(D)$,
the pull-back of the contact structure $D$ on
$\widetilde{\mathcal I}_{n+2}$:
for $v \in T\widetilde{\mathcal I}_{n+2}$, $v \in \widetilde{D}_{(V_1, V_{n+1})}$ if $(\pi_1)_*(v)
\in T\widetilde{{\mbox {\rm Gr}}}(1, V_{n+1})$. The contact structure $D$ on $\widetilde{\mathcal I}_{n+2}$
is the pull-back of the contact structure on ${\mathcal I}_{n+2}$ introduced in \S \ref{Legendre duality.}.
Similar constructions go as well for non-oriented case.
Define two distributions (vector sub-bundles) ${\mathcal C} \subset {\mathcal D}
\subset T{\mathcal F}_{n+2}$
on the non-oriented flag manifold ${\mathcal F}_{n+2}$
as follows: For $v \in T{\mathcal F}_{n+2}$,
$v \in {\mathcal D}_{(V_1, \dots, V_{n+1})}$ if
${\pi_1}_*(v) \in T{{\mbox {\rm Gr}}}(1, V_{n+1}) (\subset T{{\mbox {\rm Gr}}}(1, \mathbf{R}^{n+2}))$, while
$v \in {\mathcal C}_{(V_1, \dots, V_{n+1})}$ if
${\pi_i}_*(v) \in T{{\mbox {\rm Gr}}}(i, V_{i+1}) (\subset T{{\mbox {\rm Gr}}}(i, \mathbf{R}^{n+2})), (1 \leq i \leq n)$.
We call also the distribution ${\mathcal D}$ {\it pseudo-contact distribution} and
${\mathcal C}$ {\it canonical distribution}.
Clearly the forgetful covering $\pi : \widetilde{\mathcal F}_{n+2} \to {\mathcal F}_{n+2}$
induces a local isomorphism of $\widetilde{\mathcal D}$ and ${\mathcal D}$ (resp.
$\widetilde{\mathcal C}$ and ${\mathcal C}$). The pseudo-contact structure ${\mathcal D}$
is the pull-back of the contact structure on ${\mathcal I}_{n+2}$
via the canonical projection $\pi_{1, n+1} : {\mathcal F}_{n+2} \to {\mathcal I}_{n+2}$.
Now we describe the local structure of the canonical distribution
${\mathcal C} \subset T{\mathcal F}_{n+2}$.
Since ${\mathcal C}$ is ${\mbox {\rm GL}}(n+2, \mathbf{R})$-invariant, we describe ${\mathcal C}$
in a neighbourhood of the standard flag $E \in {\mathcal F}_{n+2}$
which corresponds to the unit matrix.
The flag manifold has local coordinates $x_i^{\ j}, (0 \leq j < i \leq n+1)$ near $E$ as components of lower
triangular matrices.
Then ${\mathcal C}$ is defined by the system of $1$-forms
$$
dx_i^{\ j} - x_i^{\ j+1} dx_{j+1}^{\ j} = 0, \qquad (0 \leq j, \ j+1 < i).
$$
Therefore a ${\mathcal C}$-integral curve $\Gamma(t) = (x_i^{\ j}(t))_{0 \leq j < i \leq n+1}$
through the standard flag $E \in {\mathcal F}_{n+2}$
is determined just by $x_j^{j-1}(t), (1 \leq j \leq n+1)$.
\begin{Rem}
{\rm
The complete flag manifold ${\mathcal F}_{n+2} = {\mathcal F}(\mathbf{R}^{n+2})$ (resp.
$\widetilde{\mathcal F}_{n+2} = \widetilde{\mathcal F}(\mathbf{R}^{n+2})$) possesses the duality
between
${\mathcal F}^*_{n+2} = {\mathcal F}(\mathbf{R}^{n+2*})$ (resp.
$\widetilde{\mathcal F}^*_{n+2} = \widetilde{\mathcal F}(\mathbf{R}^{n+2*})$)
by
$$
(V_1, V_2, \dots, V_n, V_{n+1}) \mapsto
(V_{n+1}^{\vee}, V_n^{\vee}, \dots, V_2^{\vee}, V_1^{\vee})
$$
where $V^{\vee} \subset \mathbf{R}^{n+2*}$ is the annihilator for $V \subset \mathbf{R}^{n+2}$.
Then, for each metric on $\mathbf{R}^{n+2}$, the dual space $\mathbf{R}^{n+2*}$ is identified with
$\mathbf{R}^{n+2}$. Thus we have the canonical involution on ${\mathcal F}(\mathbf{R}^{n+2})$ (resp.
$\widetilde{\mathcal F}(\mathbf{R}^{n+2}), {\mathcal F}(\mathbf{R}^{1, n+1}), \widetilde{\mathcal F}(\mathbf{R}^{1, n+1})$).
Similarly we have the canonical involution on
${\mathcal I}(\mathbf{R}^{n+2})$ (resp.
$\widetilde{\mathcal I}(\mathbf{R}^{n+2}), {\mathcal I}(\mathbf{R}^{1, n+1}), \widetilde{\mathcal I}(\mathbf{R}^{1, n+1})$).
This justifies our theory.
}
\enr
\section{Osculating flags on curves of finite type.}
\label{Osculating flags on curves of finite type.}
In general we treat a curve {\it of finite type} and define an analogue of Frenet-Serre frame
even when the curve is not an immersion.
Here, since $X \subset \mathbf{R}^{n+2} \setminus \{ 0\}$, we regard $\gamma$ as a curve in
$\mathbf{R}^{n+2} \setminus \{ 0\}$. The metric on $\mathbf{R}^{n+2}$ does not concern here.
Let $\gamma : I \to \mathbf{R}^{n+2} \setminus \{ 0\}$ be a $C^\infty$ curve.
The curve $\gamma$ is called of finite type at $t = t_0 \in I$
if the $(n+2)\times \infty$-matrix
$$
\tilde{A}(t) =
\left(
\gamma(t), \gamma'(t), \gamma''(t), \cdots, \gamma^{(r)}(t), \cdots
\right),
$$
is of rank $n+2$ for $t = t_0$.
We set $(n+2)\times(r+1)$-matrix
$$
\tilde{A}_r(t) =
\left(
\gamma(t), \gamma'(t), \gamma''(t), \cdots, \gamma^{(r)}(t)
\right).
$$
Then $\gamma$ is of finite type at $t = t_0$ if $\tilde{A}_r(t)$ is of rank $n+2$ for a sufficiently
large $r$.
Let ${\boldsymbol{a}} = (a_1, \dots, a_n, a_{n+1})$ be a sequence of strictly increasing natural numbers,
$1 \leq a_1 < \cdots < a_n < a_{n+1}$.
Then we call $\gamma$ of type ${\boldsymbol{a}}$ at $t = t_0 \in I$ if
$$
\begin{array}{c}
\min \{ r \mid {\mbox {\rm rank}} \tilde{A}_r(t_0) = 2\} = a_1, \quad
\min \{ r \mid {\mbox {\rm rank}} \tilde{A}_r(t_0) = 3\} = a_2, \quad \dots,
\vspace{0.2truecm}
\\
\min \{ r \mid {\mbox {\rm rank}} \tilde{A}_r(t_0) = n+2\} = a_{n+1}.
\end{array}
$$
We can define type for curves in the reduced space $P^{n+1}$ as well,
by just considering the double covering.
A point $\gamma(t_0)$ on $\gamma$ is called an {\it ordinary point} if
$\gamma$ is of type $(1, 2, \dots, n+1)$.
Otherwise it is called a {\it special point}.
The parameters of special points form discrete subset in $I$ if $\gamma$ is of finite type.
If $\gamma$ is of type ${\boldsymbol{a}}$ at $t = t_0$, then we set
$$
O_i(t_0) = \langle \gamma(t_0), \gamma'(t_0), \dots, \gamma^{(a_{i-1})}(t_0) \rangle_{\mathbf{R}},
$$
which is, by definition,
an $i$-dimensional subspace of $\mathbf{R}^{n+2}$, $(1 \leq i \leq n+1)$.
Then we have
\begin{Lem}
\label{Osculating flag}
The curve $\widetilde{\gamma} : I \to {\mathcal F}_{n+2}$ in the non-oriented
flag manifold ${\mathcal F}_{n+2}$ defined by
$$
\widetilde{\gamma}(t) : O_1(t) \subset O_2(t) \subset \dots \subset O_{n+1}(t) \subset \mathbf{R}^{n+2}
$$
is a $C^\infty$ curve.
Moreover we can give an orientation on the flag
locally near $t_0 \in I$. Namely we have local lifting of $\gamma$ in $\widetilde{\mathcal F}_{n+2}$
for the forgetful covering
$\pi: \widetilde{\mathcal F}_{n+2} \to {\mathcal F}_{n+2}$ from the manifold of oriented flags to those of non-oriented flags.
\stepcounter{Th
We call $\widetilde{\gamma}(t)$ the {\it osculating flag} of $\gamma$ at $t \in I$.
\
\noindent{\it Proof of Lemma \ref{Osculating flag}.}
Consider $(n+2)\times(n+2)$-matrix $B(t) = (\gamma(t), \gamma^{(a_1)}, \dots, \gamma^{(a_{n+1})})$.
We may suppose, after a suitable linear transformation of $\gamma$ in $\mathbf{R}^{n+2}$, that
$B(t_0)$ is the unit matrix. Then the lower triangular components of the matrix
$(\gamma(t), \gamma'(t), \dots, \gamma^{(n+1)}(t))$ provides the local representation of
$\widetilde{\gamma}$
in terms of local coordinates $\widetilde{\mathcal F}_{n+2}$ near $\widetilde{\gamma}(t_0)$.
\hfill $\Box$ \par
\
Suppose $\gamma$ is a curve in $X = E^{n+1}, S^{n+1} (\subset \mathbf{R}^{n+2})$ or $X = H^{n+1} (\subset \mathbf{R}^{1, n+1})$ and moreover suppose, in the case $X = H^{n+1}$,
the restriction of the metric to each $O_i(t)$ is non-degenerate.
If an oriented flag field along $\gamma$ is given,
then an orthonormal frame $(e_1, \dots, e_{n+1})$ along
$\gamma$ is uniquely constructed by the Gram-Schmidt's orthogonalisation
which depends on the given orientation.
We call this frame on $\gamma$ an {\it osculating frame}.
For instance, there exists the unique unit vector $e_1(t) \in O_2(t)$ normal to $e_0(t) = \gamma(t)$
such that $(e_0(t), e_1(t))$ forms an oriented basis of $O_2(t)$.
We see, by Lemma \ref{Osculating flag},
any osculating frame constructed above is $C^\infty$ along $\gamma$ which coincides, up to sign pointwise, with Frenet-Serre frame on ordinary points.
\section{Integral curves and codimension formula.}
\label{Integral curves and codimension formula.}
For an adapted framing, the lifting $\widetilde{\gamma} : I \to G$ is an integral curve
to $\widetilde{\mathcal D}$. Moreover, for an osculating framing,
$\widetilde{\gamma} : I \to G$ is an integral curve to $\widetilde{\mathcal C}$.
(See \S \ref{Pseudo-contact and canonical distributions.}.)
Thus we regard the class of adapted framed curves as the class of $\widetilde{\mathcal D}$-integral curves
in $G$ or $\widetilde{\mathcal F}_{n+2}$ or, as being locally equivalent,
the class of ${\mathcal D}$-integral curves in ${\mathcal F}_{n+2}$.
On the other hand, a curve of finite type $\gamma : I \to X$ lifts, via the osculating flag,
to a ${\mathcal C}$-integral curve
$\widetilde{\gamma} : I \to {\mathcal F}_{n+2}$ globally.
Moreover $\gamma$ lifts locally to a $\widetilde{\mathcal C}$-integral curve
$\widetilde{\gamma} : I \to \widetilde{\mathcal F}_{n+2}$, which satisfies
$e_1 = \pm\gamma'$ (arc-length differential) on immersive points pointwise.
\begin{Rem}
{\rm
A $C^\infty$ family of curves of finite types $\gamma_\lambda : I \to X$
needs not to be liftable, even locally, as a
$C^\infty$ family of ${\mathcal C}$-integral curves
$\widetilde{\gamma}_\lambda : I \to {\mathcal F}_{n+2}$.
The osculating flags do not behave smoothly under arbitrary deformation of curves.
This is why we consider also the class of ${\mathcal C}$-integral integral curves
for the bifurcation problem of envelopes.
}
\enr
Now we consider three kinds of jet spaces of curves.
First, we recall the ordinary jet space $J^r(I, X)$ consisting of $r$-jets of curves $I \to X$ or
$J^r(I, Y)$ for curves $I \to Y$. Their local description are the same as in the case $X = Y = P^{n+1}$.
O.P.Scherbak \cite{O.P.Scherbak1} shows that the codimension, called the {\it jet-codimensionin}
${\mbox{\rm{Jet-codim}}}({\boldsymbol{a}})$, in the jet space $J^r(I, P^{n+1})$ of the set $\Sigma({\boldsymbol{a}})$ of curves in $P^{n+1}$
of type
$
{\boldsymbol{a}} = (a_1, a_2, \dots, a_{n+1})
$
is given, for sufficiently large $r$, by
$$
{\mbox{\rm{Jet-codim}}}({\boldsymbol{a}}) \ = \ s({\boldsymbol{a}}) \ := \ \sum_{i=1}^{n+1} (a_i - i),
$$
the Schubert number which appears in Schubert calculus (\cite{MS}\cite{Kazarian2}).
Second, we consider the jet space of ${\mathcal D}$-integral curves,
$J^r_{\mathcal D}(I, {\mathcal F}_{n+2}) \subset J^r(I, {\mathcal F}_{n+2})$.
Each ${\mathcal D}$-integral curve
$\Gamma : I \to {\mathcal F}_{n+2}$ projects to a curve $\pi_1\circ \Gamma :
I \to P^{n+1} = {\mbox {\rm Gr}}(1, \mathbf{R}^{n+2})$ by the canonical projection
$\pi_1 : {\mathcal F}_{n+2} \to P^{n+1}$, $\pi_1(V_1, V_{n+1}) = V_1$.
Then, given type ${\boldsymbol{a}}$, we have the set of jets $\Sigma_{\mathcal D}({{\boldsymbol{a}}})$ in
$J^r_{\mathcal D}(I, {\mathcal F}_{n+2})$. We denote its codimension by
${\mbox{\rm{Jet-codim}}}_{\mathcal D}({\boldsymbol{a}})$.
Then we have
\begin{Th}
\label{codim-D}
The jet-codimension of the set of ${\mathcal D}$-integral curves
$\Gamma : I
\to {\mathcal F}_{n+2}$ such that
$\pi_1\circ\Gamma$ is of type ${\boldsymbol{a}} = (a_1, a_2, \dots, a_{n+1})$, is given by
$$
{\mbox{\rm{Jet-codim}}}_{\mathcal D}({\boldsymbol{a}}) = {\displaystyle \sum_{i=2}^{n+1}} (a_i - i) \ = \ s({\boldsymbol{a}}) - (a_1 - 1).
$$
\stepcounter{Cor
\noindent{\it Proof\,}:\
We may suppose the case point $t_0 = 0$.
Consider the integral jet space $J^r_{\rm{int}}(1, 2n+1)$ on germs of
integral curves $\Gamma : I \to
{\mathcal I}_{n+2}$ to the contact structure.
Denote by $\Sigma_{a_1} \subset J^r_{\rm{int}}(1, 2n+1)$ the set of integral jet $j^r\Gamma$ with
$\pi_1\circ \Gamma$ is of order $\geq a_1$.
Take Darboux coordinates
$x_1, \dots, x_n, z, p_1, \dots, p_n$ of ${\mathcal I}_{n+2}$ centred at $\Gamma(t_0)$ and
so that the contact structure is given by $dz - (p_1 dx_1 + \cdots + p_n dx_n) = 0$ and
$\pi_1 : {\mathcal I}_{n+2} \to {{\mbox {\rm Gr}}}(1, \mathbf{R}^{n+2})$
is given by $(x_1, \dots, x_n, z, p_1, \dots, p_n) \mapsto (x_1, \dots, x_n, z)$.
Let $\Gamma(t) = (x_1(t), \dots, x_n(t), z(t), p_1(t), \dots, p_n(t))$ and
without loss of generality, we suppose ${\mbox {\rm ord}} x_1 = a_1$.
Consider the mapping $\Pi : J^r_{\rm{int}}(1, 2n+1) \to J^{r - a_1 +1}(1, n+1)$
defined by
$$
\Pi(x_1, \dots, x_n, z, p_1, \dots, p_n) = (x_1/t^{a_1-1}, \dots,
x_n/t^{a_1-1}, z/t^{a_1-1}).
$$
Take any deformation $c(t, s) = (X_1(t, s), \dots, X_n(t, s), Z(t, s))$ of $\Pi(j^r \Gamma)$ at $s = 0$.
We set
$$
\begin{array}{ccl}
P_1(t, s) & := & \dfrac{(t^{a_1-1}Z(t, s))'}{(t^{a_1-1}X_1(t, s))'} -
\sum_{i=2}^n p_i(t)\dfrac{(t^{a_1-1}X_i(t, s))'}{(t^{a_1-1}X_1(t, s))'},
\vspace{0.2truecm}
\\
P_i(t, s) & := & p_i(t), (i = 2, \dots, n),
\end{array}
$$
for representatives at $(t, s) = (0, 0)$.
Then we get the integral deformation
$$
C(t, s) = (X_1(t, s), \dots, X_n(t, s), Z(t, s), P_1(t, s), \dots, P_n(t, s))
$$
of $\Gamma(t)$ at $s = 0$, which satisfies
$\pi(C(t, s)) = c(t, s)$. This show that any curve starting at $j^k(\pi\circ \Gamma)(0)$
in $J^{r - a_1}(1, n+1)$ lifts to a curve starting at $j^r\Gamma(0)$
in $J_{\rm{int}}^k(1, 2n+1)$.
Therefore $\Pi$ is a submersion at $j^r\Gamma(0)$.
The type ${\boldsymbol{b}}$ of $\Pi(j^r\Gamma)$ at $t = 0$ for sufficiently large is given by
$b_i = a_i - a_1 + 1$. Since the codimension $\Sigma_{a_1}$ in $J^r_{\rm{int}}(1, 2n+1)$ is given
by $n(a_1 - 1)$, we have
$$
{\mbox{\rm{Jet-codim}}}_{\mathcal D}({\boldsymbol{a}}) =
\sum_{i=1}^{n+1}(b_i - i) + n(a_1 - 1)
= \sum_{i=1}^{n+1}(a_i - a_1 + 1 - i) + n(a_1 - 1) = \sum_{i=2}^{n+1}(a_i - i).
$$
\hfill $\Box$ \par
\
Third, similarly to above, we consider the jet space of ${\mathcal C}$-integral curves,
$J^r_{\mathcal C}(I, {\mathcal F}_{n+2}) \subset J^r(I, {\mathcal F}_{n+2})$ and
$\Sigma_{\mathcal D}({{\boldsymbol{a}}})$ in
$J^r_{\mathcal D}(I, {\mathcal F}_{n+2})$. We denote its codimension by
${\mbox{\rm{Jet-codim}}}_{\mathcal C}({\boldsymbol{a}})$.
\begin{Th}
\label{codim-C}
The jet-codimension of the set of
${\mathcal C}$-integral curves
$\Gamma : I
\to {\mathcal F}_{n+2}$ such that
$\pi_1\circ\Gamma$ is of type ${\boldsymbol{a}} = (a_1, a_2, \dots, a_{n+1})$, is given by
$$
{\mbox{\rm{Jet-codim}}}_{\mathcal C}({\boldsymbol{a}}) = a_{n+1} - (n+1) \ = \ s({\boldsymbol{a}}) - s({\boldsymbol{a}}'),
$$
where ${\boldsymbol{a}}' = (a_1, a_2, \dots, a_{n})$.
\stepcounter{Cor
\noindent{\it Proof\,}:\
As is explained in \S \ref{Pseudo-contact and canonical distributions.}, a ${\mathcal C}$-integral curve
is described by the components $x_j^{j-1}(t), (1 \leq j \leq n+1)$.
In fact, by projecting to these components,
we see the diffeomorphism between the fiber $J^r_{\mathcal C}(I, {\mathcal F}_{n+2})_{(t, f)}$ over
$(t, f) \in I \times {\mathcal F}_{n+2}$ of the jet bundle
and the ordinary jet space $J^r(1, n+1)$.
To get the formula on ${\mbox{\rm{Jet-codim}}}_{\mathcal C}({\boldsymbol{a}})$, let
$\Gamma(t) = (x_{ij}(t))$ be a ${\mathcal C}$-integral curve for the coordinates introduced in
\S \ref{Pseudo-contact and canonical distributions.}
through the origin at $t = 0$.
Then we have, for the order at $t = 0$,
$$
{\mbox {\rm ord}} x_{i}^{\ j} = {\mbox {\rm ord}} x_{i}^{\ j+1} + {\mbox {\rm ord}} x_{j+1}^{\ j}, \quad (0 \leq j, j+1 < i).
$$
Therefore we have ${\mbox {\rm ord}} x_i^{\ 0} = \sum_{1\leq j \leq i} {\mbox {\rm ord}} x_j^{\ j-1}$ and hence
${\mbox {\rm ord}} x_i^{\ 0} - {\mbox {\rm ord}} x_{i-1}^{\ 0} = {\mbox {\rm ord}} x_i^{\ i-1} \geq 1, (2 \leq i \leq n+1)$.
Then the type of $\pi_1\circ\Gamma$ is of type at $t = 0$ if and only if
${\mbox {\rm ord}} x_{i}^{\ 0} = a_i, (1 \leq i \leq n+1)$. The condition is equivalent to that
${\mbox {\rm ord}} x_i^{\ i-1} = a_i - a_{i-1}, (1 \leq i \leq n+1)$.
Regarding the codimension in $J^r(1, n+1)$, we have
$$
{\mbox{\rm{Jet-codim}}}_{\mathcal C}({\boldsymbol{a}}) =
\sum_{i=1}^{n+1} ({\mbox {\rm ord}} x_i^{\ i-1} - 1) = a_{n+1} - (n+1).
$$
\hfill $\Box$ \par
\begin{Rem}
{\rm
If $\gamma = (1, x_1^{\ 0}, \dots, x_{n+1}^{\ 0}) : I \to \mathbf{R}^{n+2} \setminus \{ 0\}$
at $t \in I$ is ${\boldsymbol{a}} = (a_1, a_2, \dots, a_{n+1})$, then the dual curve $\gamma^* =
(1, x_{n+1}^{\ n}, x_{n+1}^{\ n-1}, \dots, x_{n+1}^1, x_{n+1}^{\ 0}) : I \to \mathbf{R}^{n+2} \setminus \{ 0\}$ is of type
$$
{\boldsymbol{a}}^* = (a_{n+1} - a_n, a_{n+1} - a_{n-1}, \dots, a_{n+1} - a_1, a_{n+1}).
$$
(Arnold-Scherbak's theorem \cite{O.P.Scherbak1}).
}
\enr
As a consequence, we observe that
$$
{\mbox{\rm{Jet-codim}}}_{\mathcal C}({\boldsymbol{a}}) \leq {\mbox{\rm{Jet-codim}}}_{\mathcal D}({\boldsymbol{a}})
\leq {\mbox{\rm{Jet-codim}}}({\boldsymbol{a}}).
$$
Since the transversality theorem (\cite{Mather}) hold, we see a curve of type ${\boldsymbol{a}}$ at a point
appear generically if $s({\boldsymbol{a}}) \leq 1$ in the class of curves in $P^{n+1} = {\mbox {\rm Gr}}(1, \mathbf{R}^{n+2})$, and
${\boldsymbol{a}} = (1, 2, \dots, n, n+1), (1, 2, \dots, n, n+2)$.
Moreover, a curve of type ${\boldsymbol{a}}$ at a point
appear momentarily in a generic one-parameter family of curves in $P^{n+1}$
if $s({\boldsymbol{a}}) \leq 2$. The list is given by
$$
(1, 2, \dots, n, n+1), (1, 2, \dots, n, n+2), (1, 2, \dots, n, n+3), (1, 2, \dots, n+1, n+2).
$$
For adapted framed curves we have:
\begin{Th}
\label{adapted}
For a generic one-parameter family of integral curves $\Gamma_\lambda : I \to {\mathcal F}_{n+2}$
($\lambda \in J$, $J$ being a one-dimensional manifold)
to the pseudo contact structure ${\mathcal D}$
on the flag manifold ${\mathcal F}_{n+2}$, the type of $\gamma_\lambda = \pi_1\circ \Gamma_\lambda$
and $\widehat{\gamma}_\lambda = \pi_{n+1}\circ\Gamma_\lambda$
at any point in $I$ for any parameter $\lambda \in J$ is one of the following list:
$$
(1, 2, \dots, n, n+1), (1, 2, \dots, n, n+2), (1, 2, \dots, n, n+3), (1, 2, \dots, n+1, n+2), (2, 3, 4) (n = 2).
$$
\stepcounter{Cor
\noindent{\it Proof\,}:\
For ${\mathcal D}$-integral curves, the transversality theorem holds. In fact in \cite{Ishikawa6}
the transversality theorem for integral curves to contact structure and the pseudo contact structure
is the pull-back by the submersion $\pi : {\mathcal F}_{n+2} \to {\mathcal I}_{n+2}$ of the
contact structure on the incident manifold ${\mathcal I}_{n+2}$.
By Theorem \ref{codim-D}, we see ${\mbox{\rm{Jet-codim}}}_{\mathcal D}({\boldsymbol{a}}) \leq 1$
if and only if ${\boldsymbol{a}} = (1, 2, \dots, n, n+1), (1, 2, \dots, n, n+2)$.
Moreover ${\mbox{\rm{Jet-codim}}}_{\mathcal D}({\boldsymbol{a}})
\leq 2$ if and only if ${\boldsymbol{a}}$ is one of the above list. Therefore we have the result.
\hfill $\Box$ \par
\
For osculating framed curves we have:
\begin{Th}
\label{osculating}
For a generic one-parameter family of integral curves $\Gamma_\lambda : I \to {\mathcal F}_{n+2}$,
($\lambda \in J$)
to the canonical structure ${\mathcal C}$,
the type of $\gamma_\lambda = \pi_1\circ \Gamma_\lambda$
and $\widehat{\gamma}_\lambda = \pi_{n+1}\circ\Gamma_\lambda$
at any point in $I$ for any parameter $\lambda \in J$ is one of the following list:
$$
\begin{array}{c}
(1, 2, \dots, n, n+1), (1, 2, \dots, i, i+2, \dots, n+2) (0 \leq i \leq n),
\vspace{0.2truecm}
\\
(1, 2, \dots, i, i+2, \dots, j, j+2, \dots, n+3) (0 \leq i < j \leq n+1).
\end{array}
$$
\stepcounter{Cor
\noindent{\it Proof\,}:\
It suffices to note that the transversality theorem for
integral curves $I \to {\mathcal I}_{n+2}$ for the contact structure (\cite{Ishikawa6}) implies
the transversality theorem for ${\mathcal C}$-integral curves
$I \to {\mathcal F}_{n+2}$.
Then we have the required result by Theorem \ref{codim-C}.
\hfill $\Box$ \par
\section{Singularities of envelopes and their bifurcations.}
\label{Singularities of envelopes and their bifurcations.}
Let $\gamma : I \to X$ be a framed curve with framing $(e_1, \dots, e_{n+1})$.
Then the {\it envelope} $E(\gamma)$ of $\gamma$,
generated by the family of tangent hyperplanes $e_{n+1}^{\perp}(t) \subset T_{\gamma(t)}X$,
is defined as follows (\cite{Thom}): Take the frame-dual $\widehat{\gamma} = e^{n+1} : I \to Y$
and take the fiber product
$$
W := \{ (t, z) \in I \times Z \mid \widehat{\gamma}(t) = \pi_2(z) \}
$$
of $\widehat{\gamma} : I \to Y$ and $\pi_2 : Z \to Y$. Then $W$ is an $(n+1)$-dimensional manifold.
The envelope $E(\gamma)$ is defined as the set of critical values of the projection $\Pi_1 : W \to X$ defined
by $\Pi_1(t, z) = \pi_1(z)$, for $(t, z) \in W$.
The lifting $\widetilde{\gamma} : I \to G \subset \widetilde{\mathcal F}_{n+2}$ projects to
the integral lifting $\overline{\gamma} : I \to Z \subset \widetilde{\mathcal I}_{n+2}$,
$\overline{\gamma}(t) = (e_0(t), e_{n+1})(t)$, to ${\mathcal D}$.
Note that $\pi_2\circ \overline{\gamma} = \widehat{\gamma}$.
Moreover
if we consider the osculating hyperplanes to $\widehat{\gamma} : I \to Y \subset \widetilde{{\mbox {\rm Gr}}}(1, \mathbf{R}^{n+2})$, (resp. $\widetilde{{\mbox {\rm Gr}}}(1, \mathbf{R}^{1, n+1})$), we get the dual curve
$\widehat{\gamma}^* : I \to P^{n+2}$ (resp. $I \to P^{1, n+1}$), forgetting the orientation
if necessary. In fact, $\widehat{\gamma}^* = \pi_1\circ\Gamma$ for the ${\mathcal C}$-integral lift
$\Gamma : I \to {\mathcal F}_{n+2}^{0}$ of $\widehat{\gamma}$ constructed by associated osculating flags to
$\widehat{\gamma}$, with respect to $\pi_{n+1} : {\mathcal F}_{n+2}^{0} \to P^{n+1}$ (resp. $P^{1, n}$).
\
\noindent
{\it Proof of Theorem \ref{envelope}.}
By Theorem \ref{adapted}, the type of
a curve $\widehat{\gamma}_\lambda : I \to Y$ in a generic one-parameter family of framed curves
at a point $(t, \lambda)$ is one of ${\rm (I)}: (1, 2, 3), {\rm (II)}: (1, 2, 4), {\rm (III)} : (1, 3, 4), {\rm (IV)} : (1, 2, 5)$ or ${\rm (V)} : (2, 3, 4)$.
Then we have the normal forms of singularities by the classification results in \cite{Ishikawa1}.
\hfill $\Box$ \par
\
Moreover, by Theorem \ref{osculating}, we have immediately:
\begin{Th}
For a generic one-parameter family $\gamma_\lambda$ of osculating framed curves in $E^3, S^3, H^3$,
the frame dual $\widehat{\gamma}_\lambda$
at a point $(t, \lambda)$ has one of type in the list
$$
(1, 2, 3); (1, 2, 4), (1, 3, 4), (2, 3, 4); (1, 2, 5), (1, 3, 5), (1, 4, 5), (2, 3, 5), (2, 4, 5), (3, 4, 5).
$$
Corresponding to each type in the above list,
the dual curve $\widehat{\gamma}_\lambda^*$ turns to be of type
$$
(1, 2, 3); (2, 3, 4), (1, 3, 4), (1, 2, 4); (3, 4, 5), (2, 4, 5), (1, 4, 5), (2, 3, 5), (1, 3, 5), (1, 2, 5).
$$
\stepcounter{Cor
By \cite{Ishikawa1}, the diffeomorphism class of the envelope
$E(\gamma_\lambda)$ is determined in the case
$$
(1, 2, 3), (2, 3, 4), (1, 3, 4), (1, 2, 4), (3, 4, 5), (1, 2, 5).
$$
Moreover in any case the topological class of $E(\gamma_\lambda)$ is determined (\cite{Ishikawa2}).
We will describe in the forthcoming paper in detail,
the topological bifurcations of envelopes for osculating framed curves of type
${\mbox{\rm{Jet-codim}}}_{\mathcal C}({\boldsymbol{a}}) \leq 2$, namely, for
$$
(2, 4, 5), (1, 4, 5), (2, 3, 5), (1, 3, 5), (1, 2, 5).
$$
\begin{Rem}
{\rm
Though the list of singularities is common for all of three geometries, the geometric characters are of course
distinguished. For instance, in the case $n = 2$ and for an adapted frame $e_0' = e_1$,
we have the structure equation, under the arc-length derivative,
$$
\left\{
\begin{array}{cccccc}
e_1' & = & - \delta e_0 & & + \kappa_1 e_2 & + \kappa_2 e_3, \\
e_2' & = & & - \kappa_1 e_1 & & + \kappa_3 e_3, \\
e_3' & = & & - \kappa_2 e_1 & - \kappa_3 e_2, &
\end{array}
\right.
$$
where $\delta = 0, 1, -1$ for $X = E^3, S^3, H^3$ respectively (\cite{IL}).
In general, we characterise the type ${\boldsymbol{a}}$ of the frame-dual $\widehat{\gamma}$ for a framed curve $\gamma$
by polynomials, distinguished in each gemetry,
of geometric invariants of $\gamma$ and their derivatives up to order $a_{n+1} - 1$.
For $E^3$, see \cite{Ishikawa6}.
}
\enr
{
\footnotesize
|
\section{Introduction\label{seccao 1}}
The main goal of this article is to present an alternative tool to study the
dynamics of a real rational function, using results from the Numerical Range
Theory, and has two main parts. The first, comprising Sections 2-3, is
concerned to the adequation of the Numerical Range Theory to the data
provided from rational maps. As described in Milnor\cite{Milnor 1993}, each
map $f$, in the space \textit{Rat}$_{2}$, can be expressed as a ratio
\begin{equation*}
f(z)=\frac{p(z)}{q(z)}=\frac{a_{0}z^{2}+a_{1}z+a_{2}}{b_{0}z^{2}+b_{1}z+b_{2}%
},
\end{equation*}%
where $a_{0}$ and $b_{0}$ are not both zero and $p(z)$, $q(z)$ have no
common root. Milnor\cite{Milnor 1993} states that we can obtain a roughly
description of the topology of this space \textit{Rat}$_{2}$ that can be
identified with the Zariski open subset of \textit{complex projective 5-space%
} consisting of all points
\begin{equation*}
\left( a_{0}:a_{1}:a_{2}:b_{0}:b_{1}:b_{2}\right) \in CP^{5},
\end{equation*}%
for which the \textit{resultant}
\begin{equation*}
\text{\textit{res}}(p,q)=\det \left(
\begin{array}{cccc}
a_{0} & a_{1} & a_{2} & 0 \\
0 & a_{0} & a_{1} & a_{2} \\
b_{0} & b_{1} & b_{2} & 0 \\
0 & b_{0} & b_{1} & b_{2}%
\end{array}%
\right)
\end{equation*}%
is non-zero. Taking $z=x+i0$ and $a_{0}=1$, $b_{0}=1$, $a_{2}=-a+i0$, $%
b_{2}=-b+i0$, $a_{1}=0$, $b_{1}=0$, with $x,$ $a,$ $b$ real numbers, we
obtain
\begin{equation*}
B=\left(
\begin{array}{cccc}
1 & 0 & -a & 0 \\
0 & 1 & 0 & -a \\
1 & 0 & -b & 0 \\
0 & 1 & 0 & -b%
\end{array}%
\right) \text{,}
\end{equation*}%
\textit{res}$(p,q)=\det $ $B$. This matrix $B$ is associated to the real
rational map $f(x)=\left( x^{2}-a\right) /\left( x^{2}-b\right) $. The map $%
f $ will be the one that we will use in our results associated to the matrix
$B $.
The second part of this article, comprising Sections 4-5, shows how can we
apply the Numerical Range Theory to the dynamics of the map $f$,
establishing the relation between some partitions of an ellipse $\Omega $,
and the symbolic space generated by the partition of the domain of $f$ in
real intervals. Moreover, we launch a conjecture that could be a path to
generate, in the future, an extension of the usual symbolic space applied to
rational maps, allowing us to describe much better the dynamics of this maps.
\section{Numerical Range Theory}
The classical numerical range of a square matrix $M_{n}$, with complex
numbers elements, is the set $W(M_{n})=\left\{ u^{\ast }M_{n}u,\text{ }u\in
S(\mathbb{C}^{n})\right\} $ , with $S(\mathbb{C}^{n})$ the unit sphere, $u$
is a vector in $\mathbb{C}^{n}$ and $u^{\ast }$ is the transpose conjugate
of $u$. The numerical range $W(M_{n})$ also can be defined as the the image
of the \textit{Rayleight quotient }$R_{M_{n}}(u)=u^{\ast }M_{n}u/u^{\ast }u$%
, $u\neq 0$. The set $W(M_{n})$ is closed and limited, and it is also a
subset of the Gaussian $\mathbb{C}$ plane. Toeplitz\cite{Toeplitz 1918} and
Hausdorff\cite{Hausdorff 1919} proved that $W(M_{n})$ is a convex region.
From Kippenhahn\cite{Kippenhahn 1951}, the boundary of the numerical range, $%
\partial W(M_{n})$ is a piecewise algebraic curve. In the particular case of
a square matrix $M_{2}$, with eigenvalues $\lambda _{1}$, $\lambda _{2}$, $%
W(M_{2})$ is a subset limited by an ellipse with foci in $\lambda _{1}$ and $%
\lambda _{2}$, result known as Elliptical Range Theorem, see Li\cite{Li 1996}%
.
\begin{proposition}
If $H_{M_{n}}=\left( M_{n}+M_{n}^{\ast }\right) /2$ and $S_{M_{n}}=\left(
M_{n}-M_{n}^{\ast }\right) /2$ are the Hermitian and skew-Hermitian parts of
$M_{n}$, respectively, then $\func{Re}(W(M_{n}))=W(H_{M_{n}})$ and $\func{Im}%
(W(M_{n}))=W(S_{M_{n}})$.
\end{proposition}
\begin{proof}
The proof can be found in Melo\cite{Melo 1999}.
\end{proof}
\begin{theorem}
\label{Teorema convex hull}To every complex matrix $%
M_{n}=H_{M_{n}}+S_{M_{n}} $ through the equation $k_{M_{n}}(\alpha
_{1},\alpha _{2},\alpha _{3})\equiv \det (\alpha _{1}H_{M_{n}}-i\alpha
_{2}S_{M_{n}}+\alpha _{3}I_{n})=0$ is associated a curve of class $n$ in
homogeneous line coordinates in the complex plane. The convex hull of this
curve is the numerical range of the matrix $M_{n}$.
\end{theorem}
\begin{proof}
Adapting the proof in Kippenhahn\cite{Kippenhahn 1951} we have the desired
result.
\end{proof}
\section{Merging $f(x)$ in $W(M_{n})$}
Hwa-Long Gau\cite{Gau 2006} states that we can obtain from a $4\times 4$
matrix an elliptical numerical range, thus we could use $B$ as defined in
section \ref{seccao 1}, but we can simplify our results if we use a smaller
matrix, $A_{2}$, trough a result in linear algebra. The new matrix $%
A_{2}=\left(
\begin{array}{cc}
1 & -a \\
1 & -b%
\end{array}%
\right) $ will produce equivalent results as the obtained from $B$.
\begin{lemma}
\label{Decomposicao unitaria}The matrix $B\ $is unitary decomposable in
\begin{equation*}
\left(
\begin{array}{cc}
A_{2} & \mathbb{O}_{2} \\
\mathbb{O}_{2} & A_{2}%
\end{array}%
\right)
\end{equation*}%
a block diagonal matrix.
\end{lemma}
\begin{proof}
In order to prove this result it is sufficient to find an unitary matrix $E$
such that%
\begin{equation*}
E^{\ast }BE=\left(
\begin{array}{cc}
A_{2} & \mathbb{O}_{2} \\
\mathbb{O}_{2} & A_{2}%
\end{array}%
\right) \text{.}
\end{equation*}%
With some computation we can see that
\begin{equation*}
E=\left(
\begin{array}{cccc}
1 & 0 & 0 & 0 \\
0 & 0 & 1 & 0 \\
0 & 1 & 0 & 0 \\
0 & 0 & 0 & 1%
\end{array}%
\right) .
\end{equation*}
\end{proof}
\begin{proposition}
$W(A_{2})=W(B)$
\end{proposition}
\begin{proof}
By lemma \ref{Decomposicao unitaria} $E^{\ast }BE=A_{2}\oplus A_{2}$ and
using the properties of numerical range we have
\begin{equation*}
W(E^{\ast }BE)=\text{\textit{convex hull}}\{W(A_{2})\cup W(A_{2})\}=\text{%
\textit{convex hull}}\{W(A_{2})\}.
\end{equation*}%
The \textit{convex hull} of a convex set is itself, then $W(E^{\ast
}BE)=W(A_{2})$. But the numerical range of $B$ is invariant under unitary
transformations, see Kippenhahn\cite{Kippenhahn 1951}, so $W(B)=W(A_{2})$.
\end{proof}
In our study we use $f(x)=(x^{2}-a)/(x^{2}-b)$, with $a>0$, $b>0$ and $a>b$.
Such map takes all real axis with exceptions $\pm \sqrt{b}$ on $(-\infty
,1)\cup \lbrack \frac{a}{b},+\infty )$. With $\Lambda =\mathbb{R}\backslash
\{\pm \sqrt{b}\}\times $ $(-\infty ,1)\cup \lbrack \frac{a}{b},+\infty )$ we
can define the graphic of $f$, $graph(f)$, as the pair $(x$, $f(x))\in
\Lambda $ and $\theta :\mathbb{R}\backslash \{\pm \sqrt{b}\}\longrightarrow
\Lambda $.
\begin{definition}
\label{Definicao funcoes}Let $C=\left\{ v\in \mathbb{C}^{2}:v=(x\text{, }%
if(x)\right\} $ and
\begin{equation*}
\Psi =\left\{ z\in \mathbb{C}:z=\frac{v^{\ast }A_{2}v}{v^{\ast }v},v\neq
0\right\} .
\end{equation*}%
We define $V:\Lambda \longrightarrow C$ as $(x$, $f(x))\longmapsto (x$, $%
if(x))$ and $\Xi :C\longrightarrow $ $\Psi $.
\end{definition}
By definition \ref{Definicao funcoes} the image of $(x$, $f(x))$ is $%
z=\left( v^{\ast }A_{2}v\right) /v^{\ast }v$ by $\Xi \circ V$.
\begin{proposition}
\label{criacao de psi em wa2}$\Psi $ $\subset W(A_{2})$
\end{proposition}
\begin{proof}
By the definition of $\Psi $ and $W(A_{2})$, using \textit{Rayleight quotient%
}, the result follows.
\end{proof}
From proposition \ref{criacao de psi em wa2} we know that $\Psi $ is a
subset of $W(A_{2})\subset \mathbb{C}$, and using definition \ref{Definicao
funcoes}, we can calculate the elements $z\in \Psi $, and as they were
defined, they will become a function of $x$. After some calculations we have
\begin{equation*}
z(x)=\frac{-a^{2}b+\left( 2a+b\right) bx^{2}-3bx^{4}+x^{6}+i\left(
a+1\right) \left( -abx+(b+a)x^{3}-x^{5}\right) }{a^{2}+\left(
b^{2}-2a\right) x^{2}+\left( 1-2b\right) x^{4}+x^{6}}.
\end{equation*}%
So, the $z\in \Psi $ is a function such that $z(x)=g(x)+ih(x)$, with $x\in
\mathbb{R}$. Some elementary calculus show us that $g(x)$ and $h(x)$ are
real rational continuous functions in $\mathbb{R}$,$\ $therefore $z(x)$ is
continuous in $\mathbb{C}$. We call some attention to the fact that $z(\pm
\sqrt{b})$ and $z(\infty )$ exists and are well defined in $\mathbb{C}$.
\begin{observation}
We have $z(x+1)=z(x)$ for
\begin{equation*}
x=\frac{1}{2}\left( -1\pm \sqrt{1+6a-2b\pm 2\sqrt{4a+9a^{2}-10ab+b^{2}}}%
\right) \text{.}
\end{equation*}
\end{observation}
Let $x_{2}$ and $x_{12}$ be the values where $f(x)=0$. If we calculate $%
z(0,a/b)$; $z(x_{2},0)$; $z(x_{12},0)$; $z(x,x)$; $z(x,-x)$ we obtain four
different points of $\Psi $. With some elementary algebra we calculated the
ellipse that contain this four points. This ellipse is
\begin{equation*}
\Omega =\left\{ \frac{\left( x-\frac{1-b}{2}\right) ^{2}}{\left( \frac{1+b}{2%
}\right) ^{2}}+\frac{y^{2}}{\left( \frac{1+a}{2}\right) ^{2}}=1\text{, }(x%
\text{, }y)\in \mathbb{R}^{2}\text{, }a>b,a>0,b>0\right\} \text{,}
\end{equation*}%
with $1+b$ and $1+a$ the minor and major axis length of $\Omega $,
respectively.
Moreover, when we use all points $(x,$ $f(x))$ they will fall in $\Omega $
under transformation by $z$.
\begin{lemma}
\label{lema de z a formar omega}Let $z(x)=g(x)+ih(x)$, then the pair $(\func{%
Re}(z),\func{Im}(z))$ satisfies $\Omega .$
\end{lemma}
\begin{proof}
We obtain this result replacing in the equation of $\Omega $, $x$ by $\func{%
Re}(z)$ and $y$ by $\func{Im}(z).$
\end{proof}
\begin{proposition}
\label{proposicao do si em psi}If $S_{i}\in \Psi $ there are, at least one $%
x_{i}$ such that $z(x_{i})=S_{i}$.
\end{proposition}
\begin{proof}
Since $z(x)=g(x)+ih(x)$ is a continuous function in $\mathbb{C}$ and by the
lemma \ref{lema de z a formar omega} the result follows.
\end{proof}
Then we conclude that $\Psi $ can be represented by the ellipse with
equation $\Omega $.
Since $\Omega $ is constructed in the space $\mathbb{R}^{2}$ and this space
is isomorphic to $\mathbb{C}$, when we refer to an element $z$ $\in \Omega $
it can understood has a vector in $\mathbb{R}^{2}$ or a complex number in
the plane $\mathbb{C}$.
There are relations between the functions $f$ and $g$ that we can observe,
described in the following lemmas. The proofs are omitted because they
result from straight calculus.
\begin{lemma}
\label{lema do zero de f e max relat g}If $x_{0}$ is a zero of $f(x)$, then $%
g(x_{0})$ is a relative maximum of $g(x)$.
\end{lemma}
\begin{lemma}
\label{lema das decont f e min de g}If $x_{0}$ is a relative minimum of $%
f(x) $ or $x_{0}$ is a discontinuity value of $f(x)$, then $g(x_{0})$ is a
relative minimum of $g(x)$.
\end{lemma}
There are similar relations between $h(x)$ and $f(x)$.
Follows some results relating $f(x)$ to $\Omega $.
\begin{lemma}
\label{lema do ponto fixo e do vertice}Let $f(x_{0})=\pm x_{0}$, then $\Xi
\circ V(x_{0},f(x_{0}))$ is vertex of $\Omega $.
\end{lemma}
\begin{proof}
If $f(x_{0})=x_{0}$, by $V$ we have $\left( x_{0}\text{, }ix_{0}\right) $.
So
\begin{equation*}
\Xi (\left( x_{0}\text{, }ix_{0}\right) )=\frac{\left(
\begin{array}{cc}
x_{0} & -ix_{0}%
\end{array}%
\right) \left(
\begin{array}{cc}
1 & -a \\
1 & -b%
\end{array}%
\right) \left(
\begin{array}{c}
x_{0} \\
ix_{0}%
\end{array}%
\right) }{\left(
\begin{array}{cc}
x_{0} & -ix_{0}%
\end{array}%
\right) \left(
\begin{array}{c}
x_{0} \\
ix_{0}%
\end{array}%
\right) }=\frac{1-b}{2}-i\frac{1+a}{2}
\end{equation*}
And if we look at the equation of $\Omega $, we see that $\left( \frac{1-b}{2%
},-\frac{1+a}{2}\right) $ is a vertex of $\Omega $.
If $f(x_{0})=-x_{0}$ we obtain another vertex of $\Omega $ in a similar way,
which is $\left( \frac{1-b}{2},\frac{1+a}{2}\right) $.
\end{proof}
\begin{lemma}
\label{lema das descontinuidades e do vertice}The discontinuities of $f(x)$
and the values where $f(x)$ has a minimum are transformed by $\Xi \circ
V\circ \theta $ in the vertex $(-b,0)$ of $\Omega $, and the roots of $f(x)$
and the $\infty $ are transformed by $\Xi \circ V\circ \theta $ in the
vertex $(1,0)$ of $\Omega $.
\end{lemma}
\begin{proof}
Since $z(x)=g(x)+ih(x)=v^{\ast }A_{2}v/v^{\ast }v$, $v=(x,if(x))$ is a
continuous function in $\mathbb{C}$, we have $z(\pm \sqrt{b})=-b$ and $%
z(\infty )=1$.
\end{proof}
\section{Partitions of $\Omega $}
Let $x_{1}$, $x_{2}$, $x_{5}$, $x_{6}$, $x_{7}$, $x_{8}$, $x_{9}$, $x_{12}$,
$x_{13}$ be the solutions of $g^{\prime }(x)=0$. By lemma \ref{lema das
decont f e min de g}, and lemma \ref{lema do zero de f e max relat g}, and
considering the order of real axis, we will have $x_{2}$ and $x_{12}$ as
zeros of $f(x)$; $x_{5}$ and $x_{9}$ as the discontinuities of $\ f(x)$ and $%
x_{7}=0$. All this values have image from $z(x)$, lemma \ref{lema de z a
formar omega}, including the infinity, being related by $%
z(x_{2})=z(x_{12})=z(\infty )$ and equal to vertex $(1,0)$ in $\Omega $, see
lemma \ref{lema das descontinuidades e do vertice}, $z(x_{5})=z(x_{9})=z(0)$
and equal to vertex $(-b,0)$ in $\Omega $. Related to the real axis, $%
z(x_{1})$ is symmetric to $z(x_{13})$ and $z(x_{6})$ is symmetric to $%
z(x_{8})$ in $\Omega $. Where are the missing $x_{3}$, $x_{4}$, $x_{10}$, $%
x_{11}$ ? They will be the values such that $z(x_{1})=z(x_{11})$, $%
z(x_{6})=z(x_{10})$, $z(x_{8})=z(x_{4})$ and $z(x_{13})=z(x_{3})$.
Using this special values $x_{i}$, $i=1,..,13$, with order $x_{i}<x_{i+1}$,
we can define a partition function $pa$, as
\begin{equation*}
pa(x)\underset{x\in \mathbb{R}}{=}\left\{
\begin{array}{ll}
I_{1}\text{,} & if\text{ }x<x_{1} \\
I_{i}\text{,} & if\text{ }x_{i-1}<x<x_{i}\text{ with }2\leq i\leq 13 \\
I_{14}\text{,} & if\text{ }x>x_{13}%
\end{array}%
\right. .
\end{equation*}
Now we will create partitions in $\Omega $ using the images $z(x_{i})$, $%
i=1,..,13$ in $\Omega $. Here, we ask attention for one particular aspect of
$z$, see proposition \ref{proposicao do si em psi}. Some intervals $I_{i}$
will be transformed in the same arc of $\Omega $. The only thing that will
distinguish them is the orientation and the origin of its end points.
\begin{definition}
Let $S_{i}=z(x_{i})$, we define $arc(S_{i},S_{i+1})$ as the arc of $\Omega $
starting at $S_{i}$ and ending at $S_{i+1}$, with counterclockwise
orientation.
\end{definition}
We define $pa_{\Omega }$, a partition function, as:
\begin{equation*}
pa_{_{\Omega }}(w)\underset{w\in \mathbb{C}}{=}\left\{
\begin{array}{ll}
J_{1}\text{,} & if\text{ }w\in \text{ }arc(z(\infty )\text{, }z(x_{1}))\text{
} \\
J_{i}\text{,} & if\text{ }w\in \text{ }arc(z(x_{i-1})\text{, }z(x_{i}))\text{
with }2\leq i\leq 13 \\
J_{14}\text{,} & if\text{ }w\in \text{ }arc(z(x_{14})\text{, }z(\infty ))%
\end{array}%
\right. .
\end{equation*}
The functions $pa(x)$ and $pa_{\Omega }(w)$ are related by
\begin{equation*}
z(I_{i})=\left\{
\begin{array}{ll}
J_{1} & if\text{ }i=1 \\
-J_{i} & if\text{ }2\leq i\leq 6 \\
J_{i} & if\text{ }i=7\text{ or }i=8 \\
-J_{i} & if\text{ }9\leq i\leq 13 \\
J_{14} & if\text{ }i=14%
\end{array}%
\right. \text{,}
\end{equation*}%
thus we can build a matrix $T_{14}$ of the transformation $pa_{\Omega
}(z(pa))$,%
\begin{equation*}
T_{14}=\left[
\begin{array}{cc}
N_{7} & \mathbb{O}_{7} \\
\mathbb{O}_{7} & N_{7}%
\end{array}%
\right] \text{,}
\end{equation*}%
with
\begin{equation*}
N_{7}=\left[
\begin{array}{ccccccc}
1 & 0 & 0 & 0 & 0 & 0 & 0 \\
0 & -1 & 0 & 0 & 0 & 0 & 0 \\
0 & 0 & -1 & 0 & 0 & 0 & 0 \\
0 & 0 & 0 & -1 & 0 & 0 & 0 \\
0 & 0 & 0 & 0 & -1 & 0 & 0 \\
0 & 0 & 0 & 0 & 0 & -1 & 0 \\
0 & 0 & 0 & 0 & 0 & 0 & 1%
\end{array}%
\right] .
\end{equation*}%
It is easy to see that $T_{14}.T_{14}=I_{14}$, $det(T_{14})=1$, and it is an
involutary matrix.
\section{Dynamics of $f(x)$}
Now we have a new tool to study the dynamics of $f$ using a symbolic space.
Using $\Omega $ to study the behavior of $f$ we will have the same
advantages that we would have when studying the behavior of second degree
polynomials functions in the unit circle.
If we define a symbolic space using the partitions created by the function $%
pa$ in the real axis we will have the problem of dealing with the
discontinuities of the function $f$ and the infinity itself. So, profiting
that $z$ is a continuous complex function in $\mathbb{C\cup \{\infty \}}$
this problem will vanish.
We can build two distinct symbolic spaces. The first will be the classical
association between the intervals produced by $pa$ in the real axis, see
Milnor\cite{Milnor 1993} for further reference, using the domain of the
function $f$, and considering an alphabet $\mathcal{A}$ with designations $%
I_{i}$ for each interval, we will have a symbolic space $\Sigma _{c}=$ $%
\mathcal{A}^{\mathbb{N}}$. The second will be constructed as we consider the
alphabet $\mathcal{B}=\{J_{1},...,J_{14}\}$, and the set $\Sigma =\mathcal{B}%
^{\mathbb{N}}$ of symbolic sequences on the elements of $\mathcal{B}$,
introducing the map $spa:\mathbb{R\cup \{\infty \}}\longrightarrow \mathcal{B%
}$.
\begin{conjecture}
The symbolic dynamics of $f$ does not change if we use $\Sigma $ instead of $%
\Sigma _{c}$.
\end{conjecture}
Both spaces are connected by the transformation matrix $T_{14}$ and doing
some calculus in matrix algebra, since this matrix is an involutary matrix,
we could get the result. All computations in our work points in that
direction. But we are still working in a suitable proof of this result.
Moreover, $\Sigma $ will work as an extension of $\Sigma _{c}$.
It means that we can identify the periodic orbits in the same values of $a$
and $b$ as we use both spaces\ $\Sigma $ and $\Sigma _{c}$. For example for
the values $a=4.01$ and $b=2.5$ the critical orbit of $f$ is periodic in
both spaces, such as all the others values of periodicity found in our
research. But they are many new sequences in $\Sigma $ that needs more work
to full understood slightly changes caused by the obliteration of $\infty $.
We are in the way.
This work, when finished, will imply a sequential work in the kneading
theory and further study of the entropy of rational real maps.
|
\section{Introduction}
This paper studies minimal prime ideals of an Ore extension over
a commutative Dedekind domain. Ore extensions are widely used as
the underlying rings of various linear systems investigated in the area Algebraic
system theory. These systems may represent systems coming from mathematical physics, applied mathematics
and engineering sciences which can be described by means of systems of
ordinary or partial differential equations, difference equations,
differential time-delay equations, etc. If these systems are linear,
they can be defined by means of matrices with entries in
non-commutative algebras of functional operators such as the ring of
differential operators, shift operators, time-delay operators, etc.
An important class of such algebras is called Ore extensions (Ore
Algebras).
The structure of prime ideals of various kind of Ore extensions have
been investigated during the last few years. In \cite{IRV2},
\cite{LMA} primes of Ore extensions over commutative noetherian
rings were considered. In \cite{CHI}, \cite{MFM}, and \cite{DSP},
prime ideals of Ore extensions of derivation type were described.
These result recently were exploited in \cite{MAW} to investigate
properties of minimal prime rings of Ore extensions of derivation
type
in term of their contraction on the coefficient ring.
In this note we extend this result to a general Ore extension of
automorphism type .
\section{Ore Extension}
We recall some definitions, notations, and more or less well known
facts concerning. A $\textit{(left) skew derivation}$ on a ring $D$
is a pair $(\sigma, \delta)$ where $\sigma$ is a ring endomorphism
of $D$ and $\delta$ is a $\it{(left)}$
~$\sigma$-$\textit{derivation}$ ~on $D$; that is, an additive map
from $D$ to itself such that $\delta(ab) = \sigma(a) \delta(b) +
\delta(a)b$ for all $a,b \in D$. For $(\sigma, \delta)$ any skew
derivation on a ring $D$, we obtain
\[
\delta(a^m) = \sum^{m-1}_{i=0} \sigma(a)^i \delta(a) a^{m-1-i}
\]
for all $a \in D$ and $m=1,2, \cdots $. (See \cite[Lemma 1.1]{GOO})
\begin{defi}
Let $D$ be a ring with identity 1 and $(\sigma, \delta)$ be a
(left) skew derivation on the ring $D$.
\noindent \textit{The Ore Extension} over $D$ with respect to the skew derivation
$(\sigma, \delta)$
is the ring consisting of all polynomials over $D$ with
an indeterminate $x$ denoted by:
\[ D[x;\sigma ,\delta ]\, \, =\, \, \{ \, f(x)\, =\, a_{n}
x^{n} \, +\, \, \cdots \, \, +\, a_{0} \, \, \mid \, \, a_{i} \, \in
\, D\, \, \} \]
satisfying the following equation, for all $a \in D$
\[xa\, \, =\, \sigma (a)x\, +\, \delta (a).\]
\end{defi}
The notations $D[x; \sigma]$ and $D[x; \delta]$ stand for the particular Ore extensions
where respectively $\delta = 0$ dan $\sigma$ the identity map.
For the case $\delta = 0$, Marubayashi et. al. \cite{MAW} studied the factor rings of
$D[x; \sigma]$ over minimal prime ideals where $D$ is a commutative Dedekind domain.
In order to extend their results to general cases, this paper investigates the class of
minimal prime ideals in $D[x; \sigma, \delta]$.
The Ore extension $D[x; \sigma, \delta]$
is a free left $D$-module with basis $1, x, x^2, \cdots$
To abbreviate the assertion, the symbol $R$ stands for
the Ore extension $D[x; \sigma, \delta]$ constructed from
a ring $D$ and a skew derivation $(\sigma, \delta)$ on $D$.
The $\textit{degree}$ of a nonzero element $f \in R$ is defined in
the obvious fashion. Since the standard form for elements of $R$ is
with left-hand coefficients, the $\textit{leading coefficient}$ of
$f$ is $f_n$ if
\[
f= f_0 + f_1x + \cdots + f_{n-1}x^{n-1} + f_nx^n
\]
with all $f_i \in D$ and $f_n \neq 0$. If $\sigma$ is an
automorphism, $f$ can also be written with right-hand coefficients,
but then its $x^n$-coefficient is $\sigma^{-n}(f_n)$.
While a general formula for $x^na$ where $a \in D$ and $n \in
\mathbb{N}$ is too involved to be of much use, an easy induction
establishes that
\[
x^na = \delta^n(a) + a_x + \cdots + a_{n-1}x^{n-1} + \sigma^n(a)x^n
\]
for some $a_1, \cdots , a_{n-1} \in D$.\\
In preparation for our analysis of the types of ideals occured
when prime ideals of an Ore extension $D[x; \sigma, \delta]$ are
contracted to the coefficient ring $D$, we consider $\sigma$-prime,
$\delta$-prime, and $(\sigma, \delta)$-prime ideals of $D$.
\begin{defi}
Let $\Sigma$ be a set of map from the ring $D$ to itself. A
$\Sigma$-ideal of $D$ is any ideal $I$ of $D$ such that $\alpha(I)
\subseteq I$ for all $\alpha \in \Sigma$. A $\Sigma$-prime ideal is
any proper $\Sigma$-ideal $I$ such that whenever $J, K$ are
$\Sigma$-ideals satisfying $JK \subseteq I$, then either $J
\subseteq I$ or $K \subseteq I$.
\end{defi}
In the context of a ring $D$ equipped with a skew derivation
$(\sigma, \delta)$, we shall make use of the above definition in
the cases $\Sigma = \{\sigma \}, \Sigma = \{\delta \}$ or $\Sigma =
\{\sigma, \delta \}$; and simplify the prefix $\Sigma$ to respectively $\sigma,
\delta,$ or $(\sigma, \delta)$. Concerning the contraction of
prime ideals in an Ore Extension to its coefficient ring,
Goodearl \cite{GOO} obtained the following theorem which will be of use later.
\begin{theo}
Let $R=D[x; \sigma, \delta]$ where $D$ is a commutative Dedekind
domain and $\sigma$ is an automorphism. If $\mathfrak{p}$ is any
ideal of $D$ which is $(\sigma, \delta)$-prime, then $\mathfrak{p}=P
\cap R$ for some prime ideal $P$ of $R$ and more specially
$\mathfrak{p}R \in$ Spec($R$) where $Spec(R)$ denotes the set of all Prime ideal in $R$.
\end{theo}
\section{Minimal Prime Ideals of Ore Extensions}
Throughout this section, let $D$ be a commutative Dedekind domain and
$R = D[x;\sigma, \delta]$ be the Ore
extension over $D$, for $(\sigma, \delta)$ is a skew derivation, $\sigma \neq 1$ is an automorphism of $D$ and $\delta \neq 0$.
Marubayashi et. al. \cite{MAW} already investigated the class of minimal prime ideals in $R$ where
$\delta=0$. For the case $\delta \ne 0$ but it is inner, Goodearl \cite{GOO} showed the
existance of an isomorphism between $D[x;\sigma, \delta]$ and $D[y;\sigma]$ as described in the following.
The $\sigma$-derivative $\delta$ is called inner if there exists an element $a \in D$ such that $\delta(b) = ab
- \sigma(b)a$ for all $b \in D$.
\begin{theo} \label{t-isomorfisma}
Let $D[x;\sigma, \delta]$ be an Ore extension where $\sigma \neq 1$
and $\delta \neq 1$. If $\delta$ is an inner $\sigma$-derivation,
i.e, there exists $a \in D$ such that $\delta(b) = ab - \sigma(b)a$
for all $b \in D$, then $D[x;\sigma,\delta]$ and $D[y;\sigma ]$, where $y=x-a$, are isomorph.
\end{theo}
Hence, by combining Theorem \ref{t-isomorfisma} and the class of minimal prime ideals in Ore extensions $R$ obtained in \cite{MAW}
we can derive the class of minimal prime ideals in Ore extension $R=D[x; \sigma, \delta]$ for $\sigma$ and $\delta$ being
respectively an automorphism and inner as the following.
Notation $\textnormal{Spec}_0 (R)$ stands for the set of all prime ideals in $R= D[x;\sigma,\delta]$ having
zero intersection with $D$.
\begin{theo} \label{t-utkdeltainner}
Let $R=D[x;\sigma,\delta]$ be an Ore extension.\\
\textnormal{(1)} The set \\
$ \{ \mathfrak{p}[x;\sigma,\delta], ~P \mid \mathfrak{p}\ \hbox{is
a}\ \sigma-\hbox{prime ideal of}\ D \hbox{and}\ P \in
\textnormal{Spec}_0 (R) ~\hbox{with} ~P \neq (0) \}$
\\
consists of all
minimal prime ideals of $R$.\\
\textnormal{(2)} Let $P \in \textnormal{Spec}(R)$ with $P \neq (0)$.
Then $P$ is invertible if and only if it is a minimal prime ideal of
$R$.
\end{theo}
Now we shall investigate the class of minimal prime ideals for general $\delta \ne 0$.
For this general case, we need the following lemma.
\begin{lem} \label{lema1}
If $P = \mathfrak{p}[x;\sigma,\delta]$ is a minimal prime ideal of
$R$ where $\mathfrak{p}$ is a $(\sigma,\delta)$-prime ideal of
$D$, then $\mathfrak{p}$ is a minimal $(\sigma,\delta)$-prime ideal
of $D$.
\end{lem}
\textit{Proof}. Assume that $\mathfrak{p}$ is a
$(\sigma,\delta)$-prime ideal of $D$ but is not minimal
$(\sigma,\delta)$-prime. Let $\mathfrak{q}$ be a
$(\sigma,\delta)$-prime ideal of $D$ such that $\mathfrak{q}
\varsubsetneq \mathfrak{p}$ and $\mathfrak{q}\ne (0)$. Then applying \cite[Theorem 3.1]{CHI} we have
$\mathfrak{q}R \in$ Spec($R$). So,
$\mathfrak{q}R \varsubsetneq \mathfrak{p}R = P$. This is a
contradiction because $P$ is a minimal prime ideal. $\square $
\begin{theo}
Let $P$ be a prime ideal of $R$ and $ P \cap D = \mathfrak{p} \ne
(0)$. Then $P$ is a minimal prime ideal of $R$ if and only if either
$P =\mathfrak{p}[x;\sigma,\delta]$ where $\mathfrak{p}$ is a minimal
$(\sigma,\delta)$-prime ideal of $D$ or $(0)$ is the largest
$(\sigma,\delta)$-ideal of $D$ in $\mathfrak{p} $.
\end{theo}
\noindent \textit{Proof}.~Let $P$ be a minimal prime ideal of $R$
and $ P \cap D = \mathfrak{p} \ne (0)$.
Since $\mathfrak{p} \neq (0)$, then there are two cases \cite[Theorem 3.1]{GOO}; namely, either $\mathfrak{p}$ is a $(\sigma,
\delta)$-prime ideal of $D$ or $\mathfrak{p}$ is a prime ideal of
$D$ and
$\sigma(\mathfrak{p}) \ne \mathfrak{p}$.
If $\mathfrak{p}$ is a $(\sigma, \delta)$-prime ideal of $D$, then
$\mathfrak{p}R \in$ Spec($R$), by \cite[Theorem 3.1]{GOO}. So,
$\mathfrak{p}R = P$ because $\mathfrak{p}R \subseteq P$ and $P$ is a
minimal prime ideal. On the other hand $\mathfrak{p}R =
\mathfrak{p}[x; \sigma, \delta]$. From here, we get $P =
\mathfrak{p}[x; \sigma, \delta]$ and $\mathfrak{p}$ is a minimal
$(\sigma, \delta)$-prime ideal of $D$,
by Lemma \ref{lema1}.
Suppose $\mathfrak{p}$ is a prime ideal of $D$ and $\sigma(\mathfrak{p})
\ne \mathfrak{p}$. Let $\mathfrak{m}$ be the largest $(\sigma,
\delta)$-ideal contained in $\mathfrak{p}$ and assume that
$\mathfrak{m} \ne (0)$. Then by primeness of $\mathfrak{p}$ it can be shown
that $\mathfrak{m}$ is a $(\sigma, \delta)$-prime ideal of
$D$. So, $\mathfrak{m}R$ is a prime ideal of $R$ by \cite[Proposition 3.3]{GOO}. On the other hand, since $\sigma(\mathfrak{p}) \ne
\mathfrak{p}$, we have $\mathfrak{m} \varsubsetneq \mathfrak{p}$.
So, $\mathfrak{m}R \varsubsetneq \mathfrak{p}R \subseteq P$, i.e,
$P$ is not a minimal prime. This is a contradiction. So, $(0)$ is
the largest $(\sigma,\delta)$-ideal of $D$ in $\mathfrak{p}. $
Conversely, let $ P = \mathfrak{p}[x; \sigma, \delta]$, where $\mathfrak{p}$
is a minimal $(\sigma, \delta)$-prime ideal of $D$. Then, according to \cite[Theorem 3.3]{GOO}, $ P =
\mathfrak{p}[x; \sigma, \delta]$ is a prime ideal of $R$. Let $Q$ be a prime ideal of $R$ where $Q \subseteq P$.
Set $\mathfrak{q} = Q \cap D$, then $\mathfrak{q} = Q \cap D
\subseteq P \cap D = \mathfrak{p}$. Applying \cite[Theorem 3.1]{GOO}, we have
two cases; namely, either $\mathfrak{q}$ is a $(\sigma,
\delta)$-prime ideal of $D$ or $\mathfrak{q}$ is a prime ideal of
$D$ and $\sigma(\mathfrak{q}) \ne \mathfrak{q}$.
For the first case, suppose $\mathfrak{q}$ is a $(\sigma, \delta)$-prime ideal of $D$. Then
$\mathfrak{q}=\mathfrak{p}$ because $\mathfrak{q}\subseteq
\mathfrak{p}$ and $\mathfrak{p}$ is a minimal $(\sigma,
\delta)$-prime ideal of $D$. So, $ P = \mathfrak{p}[x; \sigma,
\delta]= \mathfrak{q}[x; \sigma, \delta] \subseteq Q$. This implies
$P=Q$.
For the other case, if $\mathfrak{q}$ is a prime ideal of $D$, then
$\mathfrak{q}=\mathfrak{p}$ because $D$ is a Dedekind domain. So, $
P = \mathfrak{p}[x; \sigma, \delta]= \mathfrak{q}[x; \sigma, \delta]
\subseteq Q$. This implies $P=Q$. Thus
$P$ is a minimal prime ideal of $R$.
Let $(0)$ be the largest $(\sigma, \delta)$-ideal of $D$ in
$\mathfrak{p}$. Let $Q$ be a prime nonzero ideal of $R$ satisfying $Q \subseteq
P$. Set $\mathfrak{q} = Q \cap D$, then $\mathfrak{q} = Q \cap D
\subseteq P \cap D = \mathfrak{p}$. Similar to the above explanation, we have
two cases; namely, either $\mathfrak{q}$ is a $(\sigma,
\delta)$-prime ideal of $D$ or $\mathfrak{q}$ is a prime ideal of
$D$ and $\sigma(\mathfrak{q}) \ne \mathfrak{q}$ but the first case will not happen. If it is so,
then, because of $(0)$ being the largest $(\sigma,
\delta)$-ideal of $D$ in $\mathfrak{p}$,
$\mathfrak{q}= (0)$ implying a contradiction $Q \cap D = 0$ (see \cite[Lemma 2.19]{GOW}).
Thus $\mathfrak{q}$ is a prime ideal of $D$ with $\sigma(\mathfrak{q})
\ne \mathfrak{q}$. As a result, $\mathfrak{q}= \mathfrak{p}$. So $Q \cap D =
P \cap D$, which, according to \cite[Proposition 3.5]{GOW}, implies $Q = P$.
Thus $P$ is the
minimal prime ideal of $R$. $\square $
\section{Concluding Remark}
In this paper we identify all minimal prime ideals in an Ore extension over
a Dedekind domain according to their contraction on the coefficient ring.
Studying the results in Marubayashi et. al \cite{MAW} it is expected that this identification can be used to study the structure of the corresponding factor rings which is currently under investigation.
\section*{Acknowledgement}
This research was supported by ITB according to Surat Perjanjian Pelaksanaan Riset Nomor: 265/K01.7/PL/2009, 6 February 2009.
|
\section{Introduction}
For many purposes atomic Bose-Einstein condensates (BECs) can be accurately described by
means of a mean-field theoretical model, the Gross-Pitaevskii (GP) equation \cite{BECBook,book2,revnonlin}. Importantly the GP equation describes
not only the ground-state, but also macroscopic nonlinear excitations of BECs, such as matter-wave dark solitons, which have been studied theoretically
(see, e.g., Refs.~\cite{revnonlin,ourbook} and references therein) and were observed in a series of experiments
\cite{han1,nist,dutton,bpa,ginsberg2005,engels,hamburg,hambcol,kip,technion,draft6}. In fact, these localized nonlinear
structures have attracted much attention as they arise spontaneously upon crossing the BEC phase-transition \cite{zurek2,zurek3},
while their properties may be used as diagnostic tools for probing properties of BECs \cite{anglin}.
Additionally, applications of matter-wave dark solitons have been proposed: the dark soliton
position can be used to monitor the phase acquired in an atomic matter-wave interferometer in the nonlinear regime \cite{appl1,appl2}.
As matter-wave dark solitons are known to be more robust in the quasi one-dimensional (1D) geometry, the majority of relevant theoretical
studies have been performed in the framework of the 1D GP equation and, particularly, in the so-called
Thomas-Fermi (TF)-1D regime (see, e.g., Ref.~\cite{kip}). More specifically, many works have been devoted to the stability
\cite{fms,mprizolas,muryshev} and dynamical properties of dark solitons, such as their oscillations
\cite{fms,mprizolas,muryshev,motion1,motion2,motion3,motion4,motion5,motion6,motion7} and sound emission \cite{motion5,nppsound} in the presence
of a harmonic trapping potential. In the TF-1D regime, matter-wave dark solitons have also been studied in periodic (optical lattice)
potentials \cite{bbbb1,yulin,bbbb2,bbbb3,dsolyuri}, as well as in combinations of harmonic traps and optical lattices \cite{weol,nppOL,gt2,Theocharis}.
Notice that upon properly choosing the harmonic trap and optical lattice parameters, it is possible to create a {\it double-well potential},
as was demonstrated experimentally in Ref.~\cite{albiez}. It may also be formed by combining a harmonic trap with a repulsive
barrier potential, induced by a blue-detuned laser beam \cite{interf}. Double-well potentials have been studied in theory
\cite{smerzi,kiv2,mahmud,bam,Bergeman_2mode,infeld,todd,Theo06}, while they have played a key role in important experimental observations:
these include Josephson oscillations and tunneling (for a small number of atoms) or macroscopic quantum self-trapping and an asymmetric partition
of the atoms between the wells (for sufficiently large numbers of atoms) \cite{albiez}, as well as nonlinear matter-wave interference
leading to the formation of matter-wave vortices \cite{bpa2,Nate2} and dark solitons \cite{kip,draft6}.
So far, there exist only a few studies on matter-wave dark solitons in double-well potentials \cite{yuridw,ichihara}. In particular,
Ref.~\cite{yuridw} was devoted to the study of nonlinearity-assisted quantum tunneling and formation of
dark solitons in a matter-wave interferometer, which is realized by the adiabatic transformation of a double-well potential into a
single-well harmonic trap. On the other hand, in Ref.~\cite{ichihara}, the stability of the first excited state of a quasi-1D BEC
trapped in a double-well potential was studied and regimes of (in)stability were found; note that in the nonlinear regime, the first excited state
of the BEC is nothing but the stationary ``black'' soliton --- alias ``kink'' --- solution of the pertinent 1D GP equation.
At this point, it is relevant to mention that both Refs.~\cite{yuridw,ichihara} follow the majority of the theoretical works on
dark solitons in BECs, as they have also been performed in the TF-1D regime. Nevertheless, this regime has not been
practically accessible in real experiments: in fact, in most relevant experiments the condensates were usually elongated
(alias ``cigar-shaped'') three-dimensional (3D) objects, while only two recent experiments \cite{kip,draft6} were conducted
in the so-called dimensionality crossover regime from 1D to 3D \cite{str}. The observations of these experiments were found to be in
very good agreement with the theoretical predictions \cite{kip,draft6} (see also Ref.~\cite{Theo2007}, based on the use of
effective 1D mean-field models devised in Refs.~\cite{gerbier,npse,Delgado}).
In the present
work we study dark soliton states in cigar-shaped BECs --- being in the
dimensionality crossover regime from 3D to 1D --- confined in double-well potentials that are formed by a combination of a harmonic
trap and an optical lattice. Our analysis relies on the use of an effective 1D mean-field model,
which was first introduced in Ref.~\cite{gerbier} and later was rederived and analyzed in detail in Refs.~\cite{Delgado}. This model,
which has the form of a GP-like equation with a non-cubic nonlinearity, was also successfully used in Ref.~\cite{draft6} to analyze
dark soliton statics and dynamics observed in the experiment. Our aim is to systematically study, apart from the single stationary dark soliton
(see Ref.~\cite{ichihara} for a study in the TF-1D regime), multiple dark soliton states as well; such a study is particularly relevant, as
many dark solitons can be experimentally created by means of the matter-wave interference method, as demonstrated in Refs.~\cite{kip,draft6}.
Our study starts from the non-interacting (linear) limit, where we obtain the pertinent excited states of the BEC and, then, using continuation
in the chemical potential (number of atoms), we find all branches of purely nonlinear solutions, including the ground state and
single or multiple dark solitons, and their bifurcations. An important finding of our analysis is that the optical lattice (which sets the
barrier in the double-well setting) results in the emergence of nonlinear states that do not have a linear counterpart, such as
dark soliton states, with solitons located in one well of the double-well potential. As concerns the bifurcations of the various branches of solutions,
we show that particular states emerge or disappear for certain chemical potential thresholds, which are determined analytically,
in some cases, by using a Galerkin-type approach \cite{Theo06}. For each branch, we also study the stability of the pertinent solutions
via a Bogoliubov-de Gennes (BdG) analysis, which reveals the role of the anomalous
(negative energy) modes in the excitation spectra as concerns the emergence of oscillatory instabilities of dark solitons: it is found that such instabilities
occur due to the collision of an anomalous mode with a positive energy mode. Furthermore, based on the weakly-interacting limit of the model,
we study analytically small-amplitude oscillations of the one- and multiple-dark-soliton states. Particularly, we obtain equations of motion
for the single and multiple solitons from which we determine the characteristic soliton frequencies; the latter, are found to be, in many cases
in good agreement with the eigenfrequencies found in the framework of the BdG analysis.
The paper is organized as follows. We present our model in section II and provide the theoretical framework of our analysis.
In Sec. III, we study the bifurcations of the branches of the solutions of the model and study their stability via the BdG equations.
In Sec. IV, we provide some analytical estimates for the soliton frequencies based on the dynamics of solitons and compare our findings to
the ones obtained by the BdG analysis. Sec. V is devoted to a study of the soliton dynamics numerically, and shows the manifestation of
instabilities when they arise. Finally, in Sec.VI, we present our conclusions.
\section{Model and theoretical Framework}
\label{sec2}
\subsection{The effective 1D mean-field model and BdG analysis}
We consider a BEC confined in a highly elongated trap, with longitudinal and transverse confining frequencies (denoted by $\omega_z$ and
$\omega_{\perp}$, respectively) such that $\omega_z \ll \omega_{\perp}$. In such a case, use of the adiabatic approximation, together with
a variational approach for determining the local transverse chemical potential, leads to the following 1D GP equation for the longitudinal
condensate's wave function \cite{gerbier,Delgado}:
\begin {equation}
i\hbar \partial_t \psi=
\left[-\frac{\hbar^{2}}{2m} \partial_z^{2}+ V_{\rm ext}(z)
+\hbar \omega_\perp\sqrt{1+4 a |\psi|^2} \right] \psi,
\label{gerbier}
\end {equation}
where $\psi(z,t)$ is normalized to the number of atoms, i.e., $N=\int_{-\infty}^{+\infty} |\psi|^2 dx$, $a$ is the $s$-wave scattering length,
$m$ is the atomic mass, and $V_{ext}(z)$ is the longitudinal part of the external trapping potential. As demonstrated in Refs.~\cite{Delgado},
Eq.~(\ref{gerbier}) provides accurate results in the dimensionality crossover regime and the TF limit, thus describing the axial dynamics of
cigar-shaped BECs in a very good approximation to the 3D GP equation. It is worth mentioning that in the weakly-interacting limit, $4a |\psi|^2 \ll 1$,
Eq.~(\ref{gerbier}) is reduced to the usual 1D GP equation with a cubic nonlinearity, characterized by an effective 1D coupling constant
$g_{1D} = 2 a \hbar \omega_{\perp}$ (see, e.g., discussion and references in Ref.~\cite{revnonlin}).
Let us now assume that the trapping potential is a superposition of a harmonic trap and an optical lattice, characterized by an
amplitude $V_0$ and a wavenumber $k$, namely,
\begin{equation}
V_{\rm ext}(z) = \frac{1}{2}m\omega_z^2 z^2+V_0\cos^2(kz).
\end{equation}
It is clear that a proper choice of the potential parameters (and the number of atoms) may readily lead to a trapping potential that has the
form of an effective double well potential; in such a setting, the height of the barrier can be tuned by the amplitude of the optical lattice.
In our considerations below, we consider the case of a $^{87}$Rb condensate, confined in a harmonic trap with frequencies
$\omega_\perp= 10\omega_z=2\pi \times 400$Hz (these values are relevant to the experiments of Refs.~\cite{kip,draft6}); furthermore,
we will vary the optical lattice strength in the range of $V_0=0 $ to $V_0=1.16 \times 10^{-12}$~eV, thereby changing
the trapping potential from a purely harmonic form to a double-well potential. Finally, we will assume a fixed value of the optical
lattice wavenumber, namely $k=\pi/5.37$ $\mu\text{m}^{-1}$.
Next, assuming that the density, length, time and energy are measured, respectively, in units of $a$,
$\alpha_{\perp} \equiv \sqrt{\hbar/m\omega_{\perp}}$ (transverse harmonic oscillator length),
$\omega_{\perp}^{-1}$, and $\hbar \omega_{\perp}$, we express Eq.~(\ref{gerbier}) in the following dimensionless form,
\begin{equation}
i \partial_t \psi =\hat{H}\psi+ \sqrt{1+4|\psi|^2}\psi,
\label{1dDelgado}
\end{equation}
where $\hat{H}=-(1/2)\partial_{z}^{2} + (1/2)\Omega^2 z^2 + V_0\cos^2(kz)$, is the usual single-particle Hamiltonian and
$\Omega \equiv \omega_z/\omega_{\perp}$ is the normalized harmonic trap strength (note that the normalized optical lattice strength $V_0$
is measured in the units of energy $\hbar \omega_{\perp}$).
Below, we will analyze the stability of the nonlinear modes of Eq.~(\ref{1dDelgado}) by means of the Bogoliubov-de Gennes (BdG) equations
(see, e.g., Ref.~\cite{BECBook}). In particular, once a numerically exact --- up to a prescribed tolerance --- stationary state, $\psi_{0}(z)$,
is found (e.g., by a Newton-Raphson method), we consider small perturbations of this state of the form,
\begin{equation}
\psi(z,t)=\left[\psi_{0}(z)+ \left(u(z)e^{-i\omega t}+\upsilon^{\ast}(z)e^{i\omega^{\ast} t}\right)\right]e^{-i\mu t},
\label{ansatz}
\end{equation}
where $\ast$ denotes complex conjugate. This way, we derive from Eq.~(\ref{1dDelgado})
the following BdG equations:
\begin{eqnarray}
&&[\hat{H} - \mu + f] u + g\upsilon = \omega u,
\label{BdG1} \\
&&[\hat{H} - \mu + f] \upsilon + gu = -\omega \upsilon,
\label{BdG2}
\end{eqnarray}
where
$\mu$ is the chemical potential, and the functions $f$ and $g$ are given by $f= g +\sqrt{1+4n_{0}}$, and
$g = 2 \psi_{0}^2/ \sqrt{1+4n_{0}}$ (with $n_{0} \equiv |\psi_{0}|^2$). Then, solving
Eqs.~(\ref{BdG1})-(\ref{BdG2}), we determine the eigenfrequencies
$\omega \equiv \omega_{r}+i \omega_{i}$ and the amplitudes $u$ and
$\upsilon$ of the normal modes of the system. Note that if $\omega$ is an eigenfrequency
of the Bogoliubov spectrum, so are $-\omega$, $\omega^{\ast}$ and
$-\omega^{\ast}$ (due to the Hamiltonian nature of the system), and
consequently
the occurrence of a complex eigenfrequency always leads to a dynamic instability.
Thus, a linearly stable configuration is
tantamount to $\omega_i =0$, i.e., all eigenfrequencies being real.
An important quantity resulting from the BdG analysis is the amount of energy carried by the normal mode with eigenfrequency $\omega$, namely
\begin{equation}
E=\int{dz(|u|^2-|\upsilon|^2)}
\omega.
\label{energy}
\end{equation}
The sign of this quantity, known as {\it Krein sign} \cite{MacKay},
is a topological property of each eigenmode.
Importantly, if the normal mode eigenfrequencies with
opposite energy (Krein) signs become resonant then, in most cases,
there appear complex frequencies in the excitation spectrum, i.e., a dynamical instability occurs \cite{MacKay}. In order to further elaborate on
such a possibility, it is relevant to note that modes with complex or imaginary frequencies carry zero energy (due to the condition $(\omega-\omega^{\ast})\int dz (|u|^2-|\upsilon|^2)=0$, which must hold for each BdG mode), while {\it anomalous modes} have
negative energy (see, e.g., Sec.~5.6 of Ref.~\cite{BECBook}). The presence of anomalous modes in the excitation spectrum is a direct signature of
an energetic instability, which is particularly relevant to the case of dark solitons \cite{fms}; such excited states of the system are discussed below.
\subsection{Dark soliton states}
Exact analytical dark soliton solutions of NLS equations with a generalized defocusing nonlinearity, such as Eq.~(\ref{1dDelgado}), are not available.
In fact, in such cases, dark solitons may only be found in an implicit form (via a phase-plane analysis) or in an approximate form (via the
small-amplitude approximation) \cite{bass}. Nevertheless, exact analytical dark soliton solutions of Eq.~(\ref{1dDelgado}) can be found in the
weakly-interacting limit ($4|\psi|^2 \ll 1$), and in the absence of the external potential: in this case, Eq.~(\ref{1dDelgado}) is reduced to the
completely integrable cubic NLS model, which possesses single- and multiple-dark-soliton solutions on top of a background with constant density
$n=n_0=\mu$ (with $\mu$ being the chemical potential). A single dark soliton solution has the form \cite{zsd},
\begin{equation}
\psi(z,t) = \sqrt{n_0}\left[ i\nu +B\tanh(\eta) \right] \exp(-i\mu t),
\label{single}
\end{equation}
where $\eta = \sqrt{n_0}B[z-z_0(t)]$, $z_0(t)$ is the soliton center, the parameter $B\equiv \sqrt{1-\nu^2}$ sets the soliton depth given
by $\sqrt{n_0} B$, while the parameter $\nu$ sets the soliton velocity, given by $dz_0/dt=\sqrt{n_0}\nu$. Note that for $\nu=0$ the dark
soliton becomes a black soliton (alias a stationary kink), with a $\pi$ phase jump across its density minimum.
Aside from the single-dark-soliton, multiple dark soliton solutions of the cubic NLS equation are also available \cite{zsd,Blow,AA}.
In the simplest case of a two-soliton solution, with the two solitons moving with opposite velocities, $\nu_1=-\nu_2=\nu$, the wave function can
be expressed as \cite{AA} (see also Ref.~\cite{gagnon}):
\begin{equation}
\psi(z,t) = \frac{F(z,t)}{G(z,t)}\exp(-i\mu t),
\label{double}
\end{equation}
where $F=2(n_0-2n_{\rm min})\cosh(T)-2n_0 \nu \cosh(Z)+i\sinh(T)$, $G=2\sqrt{n_0}\cosh(T)+2 \sqrt{n_{\rm min}} \cosh(Z)$,
while $Z=2\sqrt{n_0} B z$, $T=2\sqrt{n_{\rm min}(n_0 - n_{\rm min})}t$, and $n_{\rm min} =n_0 -n_0 B^2= n_0 \nu^2$ is the minimum density
(i.e., the density at the center of each soliton).
Generally, the single dark soliton, as well as all higher-order dark soliton states, can be obtained in a stationary form from the
{\it non-interacting} (linear) limit of Eq.~(\ref{1dDelgado}), corresponding to $N \rightarrow 0$ \cite{KivsharPLA,konotop1}. In this limit,
Eq.~(\ref{1dDelgado}) is reduced to a linear Schr{\"o}dinger equation for a confined single-particle state, namely
the equation of the quantum harmonic oscillator, together with the contribution
of the optical lattice. However, due to the presence of the harmonic trap, the problem
is characterized by discrete energy levels and corresponding localized eigenmodes. In the
{\it weakly-interacting} limit, where Eq.~(\ref{1dDelgado}) is reduced to the cubic NLS equation, all these eigenmodes exist for
the nonlinear problem as well \cite{KivsharPLA,konotop1}, describing an analytical continuation of the linear modes to a set of nonlinear
stationary states. Recent analysis and numerical results \cite{AZ} (see also Ref.~\cite{draft6}) suggest that there are no solutions of
Eq.~(\ref{1dDelgado}) without a linear counterpart, at least in the case of a purely harmonic potential.
However, as we will show below, the presence of the optical lattice in the model results in the occurrence of
additional states, for sufficiently large nonlinearity (large atom numbers), which do not have a linear counterpart.
Next, let us discuss the dynamics of dark solitons in the considered setup with an external potential in the weakly-interacting limit ($4|\psi|^2 \ll 1$). Therefore, we factorize the total wavefunction $\psi=n_0 u$ into the background density $n_0$ and the soliton wavefunction $u$. We assume that the background density is given by the ground state wavefunction of the condensate and that we can use the TF approximation to describe the latter. Then Eq.~(\ref{1dDelgado}) may be expressed as a perturbed NLS equation for the soliton wavefunction $u$, namely,
\begin{equation}
i \partial_t u + \frac{1}{2}\partial_z^2 u -\mu (|u|^2-1)u = P(u),
\end{equation}
where the effective perturbation $P(u)$ is given by:
\begin{equation}
P(u)=(1-|u|^2)Vu+ \frac{1}{2}\frac{1}{\mu - V(z)}\frac{dV}{dz}\frac{\partial u}{\partial z}.
\end{equation}
Here, $V(z)=(1/2)\Omega^2 z^2+ V_0\cos^2(kz)$ is the trapping potential, and all terms of $P(u)$
are assumed to be of the same order.
Furthermore, assuming that the length scale of the trap is much larger than the width of the soliton,
one can apply the adiabatic perturbation theory for dark solitons devised in Ref.~\cite{KivsharPRE,motion2} to derive
the following equation of motion for the slowly-varying dark soliton center $z_0(t)$
\cite{Theocharis}:
\begin{equation}
\frac{d^2 z_0}{dt^2} = -\frac{1}{2}\frac{dV_{\rm eff}}{d z_0},
\label{eqofm}
\end{equation}
where the effective potential felt by the soliton is given by:
\begin{equation}
V_{\rm eff}(z)=\frac{1}{2}\Omega^2 z^2 + \frac{1}{2} V_0\left[1-\left(\frac{\pi^2}{3}-2\right)\frac{k^2}{6\mu}\right] \cos(2kz).
\label{Veff}
\end{equation}
At this point we should make the following remarks.
First, for the derivation of Eq.~(\ref{eqofm}), it was assumed that the background density of the condensate can be described via the TF
approximation; this actually means that Eq.~(\ref{eqofm}) is valid only for sufficiently large number of atoms.
At the same time, however, the derivation of Eq.~(\ref{eqofm}) was done in the framework of the cubic NLS model, i.e., the weakly-interacting
limit of Eq.~(\ref{1dDelgado}), which is valid for sufficiently small number of atoms. Nevertheless, Eq.~(\ref{eqofm}) may still be relevant to
provide some estimates. It is known that the soliton oscillation frequency (for a BEC confined in a harmonic trap) is up-shifted for large number of atoms due to the dimensionality of the system (i.e., the effect of transverse dimensions, which are taken into regard in the derivation
of Eq.~(\ref{1dDelgado})) \cite{Theo2007}. Particularly, one should expect that in the case of a purely harmonic trap the soliton oscillates with
a frequency
$\omega_{\rm ds}>\Omega/\sqrt{2}$,
which can be found numerically in the absence of the optical lattice (note that $\Omega/\sqrt{2}$ is the characteristic value of
the soliton oscillation frequency in a 1D harmonic trap of strength $\Omega$
in the TF limit~\cite{fms,mprizolas,muryshev,motion1,motion2,motion3,motion4,motion5,motion6}). For a more detailed discussion see Sec.~\ref{sec 4a}.
Moreover, for sufficiently small optical lattice wavenumbers (such as the chosen value
of $k=\pi/5.37 \mu m^{-1}$),
and large chemical potentials (TF limit)
the effective potential of Eq.(\ref{Veff}) can be simplified to the following form:
\begin{equation}
V_{\rm eff}(z)=
\omega_{\rm ds}^2 z^2 + \frac{1}{2} V_0 \cos(2kz).
\label{simpeffpo}
\end{equation}
We complete this Section by mentioning that the case of two (or more) spatially well separated solitons can also be treated analytically, taking into account the repulsive
interaction potential between two solitons. In the case of two spatially separated solitons located at $ z_1$ and $z_2$, on top of a constant background with density
$n_0$, the interaction potential takes the form \cite{draft6},
\begin{equation}
V^{\rm int}(z_1, z_2)=\frac{n_0}{\sinh^2(\sqrt{n_0}(z_2-z_1))}.
\end{equation}
\section{Bifurcation and BdG analysis}
\label{sec3}
Let us now proceed by investigating the existence, the stability and possible bifurcations of the lowest macroscopically excited states
of Eq.~(\ref{1dDelgado}). We fix the parameter values as follows:
$\Omega=0.1$, $k=\pi/5.37$ $\mu\text{m}^{-1}$ and $V_0=1.16 \times 10^{-12}$ eV; for such a choice,
the four lowest eigenvalues of the operator $1+\hat{H}$ (corresponding to the linear problem)
are found to be:
\begin{equation}
\omega_{0}=2.138 \times 10^{-12} {\rm eV}, \,\,
\omega_{1}= 2.142\times 10^{-12} {\rm eV}, \,\,
\omega_{2}=2.634\times 10^{-12} {\rm eV}, \,\,
\omega_{3}= 2.700\times 10^{-12} {\rm eV}, \,\,
\label{eigenvalues}
\end{equation}
and the energy of the first excited state is almost degenerate with that of the ground state, i.e., $\omega_{0}\sim\omega_{1}$.
\begin{figure}[htbp]
\includegraphics[width=10cm,height=7cm]{./fig1.eps}
\caption{(Color online) Number of atoms as a function of the chemical potential for the different states. The potential parameters are
$\Omega=0.1$, $k=\pi/5.37$ $\mu\text{m}^{-1}$ and $V_0=1.16 \times 10^{-12}$ eV.
The insets show the densities of the different states for $\mu=3 \times 10^{-12}$ eV.}
\label{Fig: 1}
\end{figure}
Fig.~\ref{Fig: 1} shows the number of atoms as a function of the chemical potential for different branches of solutions. The insets show
the spatial density profiles for $\mu=3\times 10^{-12}$ eV, with the units of the horizontal
and vertical axes being given by
$\mu m$ and
$1/\mu m$, respectively.
Branch (1) corresponds to the states with the largest number of atoms, for fixed chemical potential. The respective states are
symmetric and have no nodes. Continuation of this symmetric branch to the linear limit ends at the eigenvalue $\omega_{0}$, corresponding to the
ground state of the linear problem.
The second branch [branch (2)] starts from the first excited state in the linear limit, $\mu \rightarrow \omega_1$ as $N \rightarrow 0$.
The wavefunctions of the states of this branch are {\it antisymmetric} and have a node at the center of the barrier, while their densities are
symmetric. From this density symmetric one-soliton branch, two asymmetric one-soliton branches, namely branch (3) and its mirror image with
respect to the $z=0$ axis, bifurcate close to the linear limit, at $\mu \simeq 2.144 \times 10^{-12}$~eV -- see the close-up inset.
Let us define a local occupation number, i.e., number of atoms in the different wells,
as the integral over the density up to the center of the barrier (see also Sec. V below). The occupation number in the well
with the node is then smaller than the occupation number without the node.
This can be explained by
the fact that the state with one node is the first excited state, characterized by a higher energy than the one without a node, which is the
ground state. Thus, in order to balance the chemical potential in both wells
one needs a larger interaction energy (i.e., more atoms) in the well without a node.
Note that the macroscopic quantum self-trapping (MQST) state as predicted in Ref \cite{smerzi} is not considered in what follows.
The reason for that is that the MQST is a running phase state while our bifurcation analysis below focuses on the stationary states of the system.
Branches (4) and (5) have no linear counterparts since at $\mu \simeq 2.828\times10^{-12}$~eV they
``collide" and disappear. The states that belong to these branches are asymmetric, exhibiting two nodes. The state with larger number of atoms
(for a fixed chemical potential) has one node approximately at the barrier and one node in one well. The other state has both nodes in one well. Once again,
the occupation number in the well with two nodes is less than the one of the well with one node, so as to balance the chemical potential.
Branch (6) starts from the second excited state of the linear problem, hence
$\mu \rightarrow \omega_2$ as $N\rightarrow 0$. Therefore, the density of
the respective state is symmetric and there is one node in each well.
Branch (7) starts from the third excited state of the linear problem,
and $\mu\rightarrow \omega_3$ as $N\rightarrow 0$. The states belonging to
this branch have three nodes, one located in each well and one at the barrier, and are symmetric. Close to the linear limit, at
$\mu \simeq 2.795\times10^{-12}$~eV, two asymmetric three-soliton branches, namely branch (8) and its mirror image
with respect to the $z=0$ axis, bifurcate from this state. This state has two nodes in one well and one in the other. The occupation number in
the well with more nodes is smaller than in the well with just one node in order to balance the chemical potential. Notice that there exist two
more three-soliton branches without linear counterparts, but they only occur at higher chemical potentials --- where the BEC has occupied
four-wells rather than two --- so they will not be considered here.
\begin{figure}[htbp]
\begin{minipage}[c]{8 cm}
\includegraphics[width=5.5cm,angle=270]{./fig2a.eps}
\end{minipage}
\begin{minipage}[c]{8 cm}
\includegraphics[width=5.5cm,angle=270]{./fig2b.eps}
\end{minipage}
\begin{minipage}[c]{8 cm}
\includegraphics[width=5.5cm,angle=270]{./fig2c.eps}
\end{minipage}
\begin{minipage}[c]{8 cm}
\includegraphics[width=5.5cm,angle=270]{./fig2d.eps}
\end{minipage}
\caption{(Color online) BdG Spectrum of the states shown in Fig. \ref{Fig: 1}. First panel shows the spectrum of the first branch, second panel the spectrum of the second branch and so forth. The (blue) $\divideontimes$ symbol denotes an eigenmode of positive Krein sign (or zero Krein sign for a vanishing eigenfrequency),
the (green) $\times$ symbol an eigenmode of negative Krein sign, and the (red) $+$ symbol an eigenmode associated to a vanishing Krein sign with complex/imaginary eigenfrequency.}
\label{Fig: 2}
\end{figure}
Next, let us study the
stability of the states belonging to the above
mentioned branches by considering the respective excitation spectra,
shown in Fig.~\ref{Fig: 2}.
In this figure, the (blue) $\divideontimes$ symbol denotes a positive Krein
sign mode (or a zero one for a vanishing frequency), the (green) $\times$ symbol a negative Krein sign mode,
i.e., a negative energy {\it anomalous mode}, and the (red) ($+$) symbol a vanishing Krein sign, i.e., an eigenmode associated with complex/imaginary
eigenfrequency. Generally speaking,
there are two different kinds of instability, corresponding to the cases of either a purely imaginary eigenfrequency, or a
genuinely complex
eigenfrequency. The latter case gives rise to the so-called oscillatory instability,
stemming from the collision of a negative energy mode with one of positive energy.
The first panel of Fig.~\ref{Fig: 2} shows the BdG spectra along the first branch. The imaginary part is zero for every value of the
chemical potential, a fact reflecting the stability of this state. The absence of negative energy modes is expected, as this state is
actually the ground state of the system. Since the BdG spectrum refers to all excitations of the state, one can attribute in the linear
limit to each mode an excited state of the linear problem. The energy gap between the ground state and the first excited state corresponds
to the energy of the first mode. Since the frequency of the so-called \textit{bosonic Josephson junction}, namely the oscillation frequency
characterizing the transfer of atoms between the two wells, is equal to this frequency in the linear limit, this mode is connected to
the oscillation of atoms between the two wells. We investigate this fact in more detail below (see
Sec.~\ref{section: Dynamics}) by direct simulations in the framework of
Eq.~(\ref{gerbier}). The frequency value of the second nonzero BdG mode of the ground state spectrum is correspondingly equal to the energy gap between the ground state and the second excited
state in the linear limit, and so forth. We define the difference in the way that these frequency differences are positive.
The second panel shows the BdG spectra along the second branch, solutions of which represent states with a density minimum at the center of
the barrier. In the linear limit one can assign to each mode an energy difference of the linear problem similar as in the BdG spectrum of the ground state. However in this case the frequency difference of the first excited state to the ground state is negative (according to our definition) leading to a negative energy mode. As one can see in the inset, the lowest mode has negative energy close to the linear limit. However, increasing the chemical
potential the eigenfrequency of this mode is moving rapidly towards the origin of the spectrum and, at $\mu= 2.1437\times10^{-12}$~eV,
becomes and remains purely imaginary, thus carrying zero energy (red $+$ symbols). This happens exactly at the point where the third branch
bifurcates as one can see in the inset of Fig. \ref{Fig: 1}. For the above mentioned potential parameters, this bifurcation happens in the weakly-interacting regime, where Eq.~(3) reduces
to the cubic NLS equation. Thus, one can apply the Galerkin-type approach of Ref.~\cite{Theo06} and find that the critical value of the norm
$N$ at which the bifurcation occurs is given by
\begin{equation}
N_\text{cr}=\frac{\omega_1-\omega_0}{3B-A_1},
\label{Ncr}
\end{equation}
where $\omega_{0,1}$ are the first two lowest eigenvalues of the operator $1+\hat{H}$, $A_1=\int{\psi_{1}^4dz}$,
$B=\int{\psi_{0}^2\psi_{1}^2dz}$, while $\psi_{0,1}$ denote the ground state and the first excited state of the operator $1+\hat{H}$,
respectively. One can also determine the critical chemical potential at which the symmetry breaking is expected to occur,
namely $\mu_{cr}=1+\omega_{0}+3BN_{\text{cr}}$ \cite{Theo06} leading for (these parameter values) to
$\mu_{\text{cr}}= 2.1436\times10^{-12}$~eV, which is in very good agreement with our numerical findings.
The third panel shows the BdG spectra along the third branch which bifurcates at
$\mu= 2.1437\times10^{-12}$~eV from the second one. All the eigenfrequencies of the arising asymmetric state are real close to the linear limit,
thus it is concluded that this state is linearly stable and that the bifurcation is a supercritical pitchfork. This state has one negative
energy mode as well, since it is a one-soliton state.
The eigenfrequency of the negative energy mode increases with increasing chemical potential,
passes through the mode
corresponding to the second excited state of the linear limiting case, and finally collides at some critical value of the chemical potential, $\mu = 2.76 \times 10^{-12}$~eV,
with the mode
corresponding to the third excited state; this results in the generation of a quartet of unstable eigenfrequencies and,
at this point, the state becomes oscillatory unstable. However, the magnitude of the instability is much smaller than in the previous case of
the purely imaginary eigenfrequency.
The fourth and fifth panel show the spectra along the fourth and fifth branch, the solutions of which correspond to asymmetric two-soliton states
having no linear counter-part. The previous one has got one imaginary and one negative energy mode and the latter one has two negative energy modes. These modes result from the presence of two dark solitons. The state considered
in the fourth panel has two dark solitons, one soliton located approximately
at the barrier and one soliton in one well. The eigenfrequency corresponding to the former one is purely imaginary and looks similar to the mode of the symmetric one soliton state of the second panel.
The eigenfrequency of the other anomalous mode decreases with increasing chemical potential and collides with a mode with positive energy, thus
generating a quartet of complex eigenfrequencies. The magnitude of the imaginary part is again much smaller than the magnitude of the purely imaginary
eigenfrequency. The generation of dynamic instabilities is illustrated e.g. in panel five.
The collision of one of the anomalous modes with a mode of positive Krein sign could lead to the emergence of a quartet of complex eigenfrequencies, resulting in the concurrent presence of an apparent
level crossing in the real part of the eigenfrequency along with a nonzero imaginary part of the relevant eigenfrequency mode.
At $\mu \simeq 2.828\times10^{-12}$~eV, the states of the fourth and fifth branch collide and disappear as can be seen in Fig. \ref{Fig: 1}. Just before the collision (for slightly larger $\mu$), the state of the
fourth branch is unstable due to the presence of a mode with imaginary eigenfrequency in the BdG spectrum, while the state of the fifth branch
is linearly stable since all the BdG eigenfrequencies have only real parts. Thus, it can be concluded that, at this critical point, a saddle-node
bifurcation occurs, readily destroying these two states.
The sixth panel shows the BdG spectra along the sixth branch, the solutions
of which correspond to symmetric two-soliton states. This branch
has a linear counterpart, and we can assign to each mode in the linear limit an energy difference of the linear problem. The negative energy modes
occur for negative energy differences, e.g., at the energy difference to the ground state and the first excited state.
The anomalous modes here are close to being degenerate. Physically speaking,
this reflects the fact that the dark solitons are
essentially decoupled at the linear limit due to the presence of the optical lattice. The negative energy modes are not the
modes with the lowest eigenfrequency.
The energy gaps between the second excited state and
the third and fourth excited state, respectively, are smaller than the gaps between the
ground state and the first excited state. Therefore, these modes with
positive energy occur in the linear limit at lower frequencies.
One of the negative energy modes collides at some critical value
of the chemical
potential with the mode stemming from the fourth excited state and forms a
quartet of complex frequencies, thus leading to an oscillatory instability.
The seventh panel shows the BdG spectra along the seventh branch,
the solutions of which corresponds to the symmetric
three-soliton state exhibiting
a linear counterpart. The negative energy modes occur at the energy gaps to the ground state, the first and the second excited states. The
eigenfrequency of the latter becomes, for increasing $\mu$, purely imaginary at the point where the eighth branch bifurcates. The other two anomalous modes
are almost degenerate, reflecting again the fact that the solitons in the
different wells are almost decoupled. One of the negative energy modes
collides with a mode with positive energy and forms a quartet of
complex frequencies. However, the magnitude of the imaginary part is much
smaller than the magnitude of the purely imaginary mode. The behavior of the purely imaginary mode is similar to the behavior of the mode in
panel two (describing a single soliton at the center of the barrier). The behavior of the other two modes is similar to the modes in panel
six describing one soliton in each well.
Finally, the eighth panel shows the BdG spectra along the eighth branch, the solutions of
which corresponds to the asymmetric three-soliton state.
This state, which has two solitons in one well and one soliton in the other well, emerges at $\mu\sim 2.74\times10^{-12}$~eV, where the seventh
state (i.e., the one corresponding to the nonlinear continuation of the third excited state),
becomes unstable. The critical value of the number of atoms for which the bifurcation occurs can be predicted in the same way as for the bifurcation
of the one soliton branch; the result is:
\begin{equation}
N_\text{cr}^{(3)}=\frac{\omega_3-\omega_2}{3D-C_1},
\label{Ncr2}
\end{equation}
where $\omega_{2,3}$ are the third and fourth lowest eigenvalues of operator $1+\hat{H}$, $C_1=\int{\psi_{3}^4dz}$,
$B=\int{\psi_{2}^2\psi_{3}^2dz}$, while $\psi_{2,3}$ denote the second and third excited state of the operator $1+\hat{H}$, respectively.
The corresponding critical chemical potential is then given by $\mu_{cr}^{(3)}=1+\omega_{2}+3BN_{\text{cr}}^{(3)}$ leading (for these parameters)
to $\mu_{\text{cr}}^{(3)}= 2.7420\times10^{-12}$~eV, which is in very good agreement with the numerical critical point given above.
Right after the bifurcation, the asymmetric three soliton states are stable and, thus, we can conclude that the bifurcation
is of the supercritical pitchfork type. Two modes look similar to the negative energy modes in panel five describing a two-soliton state
in one well and the third mode looks similar to a mode of one-soliton in a harmonic trap.
From the above analysis, we can readily derive some general conclusions concerning dark soliton states in a double well potential. The sum of
negative energy modes and imaginary eigenfrequency modes
is equal to the number of nodes in the wavefunction profile
which, for large values of the chemical potential, represent
dark solitons. In the linear limit, the BdG eigenfrequencies of a state
correspond to the energy differences between that and other states of the
linear system.
For a negative energy difference the corresponding mode has a negative energy.
Dark solitons located at the center of the barrier are known to
be linearly unstable (for sufficiently high chemical potential)
associated with a purely imaginary
eigenfrequency \cite{weol,ichihara}. Thus, symmetric states with an
odd number of nodes always
have a mode which, above a critical value of the chemical potential,
becomes purely imaginary. At this critical point additional stable
asymmetric dark-soliton
states emerge through supercritical pitchfork bifurcations.
\section{Statics vs. dynamics of dark soliton states}
\subsection{The one-soliton state} \label{sec 4a}
Having investigated in detail the statics of matter-wave dark solitons in double-well potentials, we now proceed to compare these results to the
soliton dynamics, using the theoretical background exposed in Sec.~\ref{sec3}.B. Particularly, considering the case of a single dark soliton,
we can readily obtain from Eqs.~(\ref{eqofm}) and (\ref{simpeffpo}) the following approximate equation of motion for the dark soliton center:
\begin{equation}
\frac{d^2 z_0}{dt^2} = -\omega_{\rm ds}^2 z_0 + \frac{1}{2}k V_0\sin(2kz_0).
\label{EOM}
\end{equation}
It is straightforward to find the fixed points $z_{\rm o}, p_{\rm o} \equiv dz_{\rm o}/dt$
of the above dynamical system, which are given by
\begin{equation}
p_{\rm o} =0, \qquad z_{\rm o}=\frac{kV_{0}}{2\omega_{\rm ds}^2}\sin(2kz_{\rm o}).
\label{fixpoint}
\end{equation}
It is clear that the fixed points correspond to intersections of the straight line
$f_1(z_{\rm o})=z_{\rm o}$ and the sinusoidal curve
$f_2(z_{\rm o}) =\frac{kV_{0}}{2\omega_{ds}^2}\sin(2kz_{\rm o})$.
The fixed point at $(z_{\rm o},p_{\rm o})=(0,0)$
exists for all choices of ($\Omega,V_{0},k$). This equilibrium corresponds to the symmetric one-soliton state, namely a stationary black
soliton located at the center of the potential.
In the absence of the optical lattice (i.e., for $V_0=0$) this is the only existing fixed point, i.e., in the case of a purely harmonic trap, only a
symmetric one-soliton state can exist, which originates from the first
excited state of the linear problem.
On the other hand, when the optical lattice is
present (i.e., for $V_0 \ne 0$),
this situation changes and more fixed points arise when the slope of the curve $f_2(z_{\rm o})$ at the origin becomes larger than the one of the
straight line $f_1(z_{\rm o})$, namely when $k^2 V_{0}/\omega_{\rm ds}^2 > 1$, two new fixed points appear through a supercritical pitchfork, with the
newly emerging states corresponding to the bifurcating
asymmetric one-soliton states. Therefore, the bifurcating state appears only for a sufficiently strong optical lattice, such that
$V_0 > V_{0,cr} \equiv \omega_{\rm ds}^2/k^2$. Note that for even larger values of the parameter $k^2 V_{0}/\omega_{\rm ds}^2$ more fixed
points appear, corresponding to solitons located in further (i.e., more
remote from the trap center) wells of the optical lattice; for this reason,
these states are not considered herein.
Let us now investigate the spectral stability of the fixed points employing the linearization technique (see, e.g., Ref.~\cite{Strogatz}).
Particularly, we first find the corresponding Jacobian matrix (evaluated at the fixed points), which is given by:
\begin{equation*}
{\bf J} = \left(
\begin{array}{cc}
0& 1 \\
-\omega_{\rm ds}^2+k^2V_{0}\cos(2kz_0) & 0\\
\end{array} \right).
\end{equation*}
When evaluated at the trivial fixed point $(z_{\rm o},p_{\rm o})=(0,0)$, the corresponding eigenvalue problem for the Jacobian,
$\det({\bf J}-\lambda{\bf I})=0$, leads to eigenvalues $\lambda^2=-\omega_{\rm ds}^2+k^2V_{0}$.
It is clear that in the case of $k^2 V_{0}/\omega_{\rm ds}^2<1$, the eigenvalues are imaginary (note that, accordingly, the eigenfrequencies
are real since $\omega=i\lambda$) and, thus, the fixed point
$(0,0)$ is a center.
In this case, if the dark soliton is weakly displaced from the center of the
trap, it performs harmonic oscillations around the center
with an oscillation frequency given by $\omega_{\rm osc}=\sqrt{\omega^2_{\rm ds}-k^2V_{0}}$.
On the other hand, if $k^2\frac{V_{0}}{\omega_{\rm ds}^2}>1$ then the eigenvalues become real (and, accordingly, the eigenfrequencies
are imaginary), hence
$(0,0)$ becomes a saddle point, while the respective
imaginary eigenfrequency is given by
$\omega_{\rm osc}=i\sqrt{|\omega^2_{\rm ds}-k^2V_{0}|}$.
For the newly bifurcating (stable, since $V_{0}>\omega_{\rm ds}^2/k^2$)
fixed points
one can obtain the following eigenfrequency
\begin{equation}
\omega_{\rm osc}=\sqrt{\omega_{ds}^2-k^2V_{0} \cos(2kz_0)}.
\label{omega ODE}
\end{equation}
From the above analysis it is clear that the
supercritical pitchfork occurs at
\begin{equation}
V_{0,cr}=\frac{\omega_{\rm ds}^2}{k^2}.
\label{V0cr}
\end{equation}
Based on the above, one can make the following estimates and
comparisons to numerical results:
in the TF limit, and in the weakly-interacting case
(i.e., in the framework of the 1D cubic GP
equation), the dark soliton oscillation frequency
is predicted to be $\omega_z/\sqrt{2}=28.28$Hz. However, as mentioned in
Sec.~\ref{sec3}.B, a larger value of this frequency is expected in the framework of Eq.~(\ref{gerbier}), due to the effect of the dimensionality of the system.
Employing a BdG analysis in the case of a pure harmonic trap, we find that the anomalous mode eigenfrequency
(i.e., the soliton oscillation frequency) decreases with increasing chemical potential and asymptotically reaches the value
$\omega_\text{ds}=30.36$Hz (see first panel of Fig. \ref{Fig: 3} below). Using this value (and for
$k=\pi/5.37 \mu m^{-1}$), one obtains $V_{0,cr}=9.6\times10^{-14}$~eV.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=8.5cm]{./fig3a.eps}
\end{center}
\vspace{-0.5cm}
\includegraphics[width=6cm,angle=270]{./fig3b.eps}
\caption{(Color online) Top panel: Fixed points (lines) of the
dynamical system (\ref{EOM}) and position of the
soliton (open circles) calculated by the Eq. (\ref{gerbier}) with $\mu=3.8 \times 10^{-12}$~eV. Bottom left (right) panel:
BdG spectra of the symmetric (asymmetric) one-soliton state for $\mu=3.8 \times 10^{-12}$~eV as a function of $V_0$.}
\label{Fig: 4}
\end{figure}
In the top panel of Fig. \ref{Fig: 4}, we show the fixed points of the dynamical system (\ref{EOM})
as a function of the optical lattice strength $V_{0}$. The dash-dotted (green) line denotes the non-zero fixed point calculated from
Eq.~(\ref{fixpoint}) with $\omega_{\rm ds}=\omega_{z}/\sqrt{2}$, while the solid (black) line denotes the non-zero fixed point
calculated from Eq.~(\ref{fixpoint}) with $\omega_\text{\rm ds}=30.36$ Hz. Furthermore, the circles denote
the position of the node of the nonlinear stationary state --- i.e., the position of the dark soliton --- obtained by numerical integration
of Eq.~(\ref{gerbier}) with $\mu=3.8 \times 10^{-12}$~eV (in the TF limit). Clearly, at $V_{0,cr}\simeq 9.6\times10^{-14}$~eV, the asymmetric
one-soliton branch bifurcates through a supercritical pitchfork bifurcation; notice the excellent agreement between the two approaches.
The bottom left and right panels of Fig. \ref{Fig: 4} show the BdG spectra of the symmetric and asymmetric one soliton state, respectively,
as a function of the optical lattice strength (for a fixed chemical potential, $\mu=3.8 \times 10^{-12}$~eV).
For sufficiently small lattice strength, $V_0<V_{0,cr}$, the symmetric state is stable, while when the lattice strength $V_0$ is increased,
the magnitude of the instability grows and the prediction of
Eq.~(\ref{omega ODE}) deviates from the results obtained from the integration of Eq.~(\ref{1dDelgado}). To explain this discrepancy, it is relevant to recall that the result of the perturbation theory of Sec.~\ref{sec2}.B is valid only if
all perturbation terms are small and of the same order. Nevertheless, for the symmetric one soliton state --- with the soliton located at zero --- the
magnitude of the perturbation term corresponding to the harmonic trap (optical lattice) has its smallest (largest) value; thus,
for large $V_0$, these
perturbations
cannot be of the same order.
As concerns the asymmetric state, the prediction of Eq.~(\ref{omega ODE}) is in fairly good agreement to the results obtained from
the numerical integration of Eq.~(\ref{1dDelgado}). However, Eq.~(\ref{omega ODE}) cannot predict oscillatory instabilities (associated with
the existence of complex eigenfrequencies), which
occur when the anomalous mode collides with a mode of the background condensate. This can readily be understood by the fact that the derivation
of the equation of motion of the dark soliton relies on the assumption that the soliton is decoupled from the background (see, e.g.,
Refs.~\cite{motion2} and \cite{revnonlin} for details), which is not the case for oscillatory instabilities.
\begin{figure}[htbp]
\begin{minipage}[c]{8 cm}
\includegraphics[width=5.5cm,angle=270]{./fig4a.eps}
\end{minipage}
\begin{minipage}[c]{8 cm}
\includegraphics[width=5.5cm,angle=270]{./fig4b.eps}
\end{minipage}
\caption{(Color online) BdG spectra for four different values of the optical lattice strength for the symmetric dark soliton state (with the soliton
located at the trap center):
from left to right, $V_0=0$, $V_0=8.3\times10^{-14}$ eV, $V_0=1.65\times10^{-13}$ eV and $V_0=4.96\times10^{-13}$ eV.
The solid lines correspond to the predictions of Eq.~(\ref{omega ODE}). Symbol $\divideontimes$ (blue) denotes a positive (or zero for a vanishing eigenfrequency), symbol $\times$ (green) a negative and symbol $+$ (red) a vanishing Krein sign mode (for non-zero eigenfrequencies).}
\label{Fig: 3}
\end{figure}
Next, let us study the continuation of both symmetric and asymmetric one-soliton states with increasing chemical potential for various values of
the optical lattice strength.
First, we consider the symmetric one-soliton state and in
Fig.~\ref{Fig: 3} we show the BdG spectra along the corresponding branch.
The first panel shows the case of a pure harmonic trap, i.e., $V_0=0$, and the dashed line is the prediction from Eq.~(\ref{omega ODE})
using $\omega_{\rm ds}=\omega_{z}/\sqrt{2}$. It is clear that the eigenfrequencies are real, indicating the stability of this state for all values
of the chemical potential. The figure shows the constant eigenfrequency of the dipolar mode [see straight (blue) line] at the value of the
trap frequency $\omega_z$ and the anomalous mode bifurcating from the dipolar mode in the linear limit. As mentioned above, the eigenfrequency
of the anomalous mode reaches asymptotically the value $\omega_\text{ds}=30.36$ Hz.
The second panel of Fig.~\ref{Fig: 3} shows the BdG spectrum for $V_0=8.3\times10^{-14}$~eV. All eigenfrequencies are real, so the state is still
stable. The dashed line is the prediction by
Eq.~(\ref{omega ODE}) using $\omega_{ds}=\omega_{z}/\sqrt{2}$, while the
solid line is the prediction with $\omega_\text{ds}=30.36$ Hz.
The dipole mode remains approximately constant for such a small optical lattice strength. However, the degeneracy of the anomalous mode and the
dipole mode in the linear limit disappears.
The third panel shows the spectra for $V_0 =1.65 \times10^{-13}$~eV. In this case, the anomalous mode becomes purely imaginary
for increasing chemical potential, reflecting the fact that the state becomes unstable. Notice that the numerical result and the analytical prediction,
$\omega_{ODE}=i\sqrt{|\omega^2_{\rm ds}-k^2V_{0}|}$, are in a good agreement for large values of the chemical potential, where the TF limit
is reached and Eq.~(\ref{EOM}) is valid.
Since $V_0>{V}_{0,cr}$, this state is expected to be unstable in the TF limit.
The fourth panel shows the spectra for $V_0=4.96\times10^{-13}$~eV. In this
case, the anomalous mode becomes purely imaginary already at small
chemical potentials. Furthermore, the dipole mode eigenfrequency is now
significantly modified, upon the
reported interval of changes of the chemical potential.
The analytical prediction agrees again well with the
numerical BdG result in the TF limit.
\begin{figure}[htbp]
\includegraphics[width=5.5cm,angle=270]{./fig5.eps}
\caption{(Color online) BdG spectra for $V_0 = 1.65\times 10^{-13}$~eV
(left panel) and $V_0=8.3\times10^{-13}$~eV (right panel) for the
asymmetric dark soliton state, bifurcating from the symmetric one.
The notation of eigenstates is the same as in Fig. \ref{Fig: 3}.}
\label{Fig: 5}
\end{figure}
Next, we consider the continuation with increasing chemical potential of the asymmetric states bifurcating from the symmetric one when the latter
becomes unstable.
Fig.~\ref{Fig: 5} shows the BdG spectra along the asymmetric one-soliton state for $V_0=1.65\times10^{-13}$~eV and $V_0=8.3\times10^{-13}$~eV.
The analytical predictions of
Eq.~(\ref{omega ODE}) with $\omega_{\rm ds}=\omega_{z}/\sqrt{2}$ and $\omega_\text{ds}=30.36$Hz, respectively plotted with dashed and solid
lines are
in good agreement
with the respective anomalous mode eigenfrequencies. In the case of $V_0=8.3\times10^{-13}$~eV, the difference between the result of
Eq.~(\ref{omega ODE}) using $\omega_{ds}=\omega_{z}/\sqrt{2}$ and $\omega_\text{ds}=30.36$ Hz is small, showing that the dynamics is
mostly governed by the potential term stemming from the optical lattice.
\subsection{Two dark-soliton states}
We now consider the case of the two-soliton state, for which we need to incorporate the interaction potential in order to describe their dynamics.
Particularly, according to the theoretical framework of Sec.~\ref{sec2}.B,
the effective potential describing the dynamics of two solitons located at $z_1$ and $z_2$ has the following form:
\begin{equation}
V_{\text{eff}}(z_1,z_2)=
\omega_{\rm ds}^2 (z_1^2+z_2^2) + \frac{V_0}{2}(\cos(2kz_1) + \cos(2kz_2)) +\frac{2n_0}{\sinh^2(\sqrt{n_0}(z_2-z_1))}.
\label{eff. pot. two solitons}
\end{equation}
Using $d^2 z_i/dt^2 = -(1/2) dV_\text{eff}(z_1,z_2)/dz_i$ (with $i=1,2$) as per Eq.~(\ref{eqofm}), it is straightforward to derive
the following system of equations of motion,
\begin{eqnarray}
\frac{d^2 z_1}{dt^2} &=& -\omega_{\rm ds}^2 z_1 + \frac{1}{2}k V_0\sin(2kz_1) -2n_0^{3/2}\coth(\sqrt{n_0}(z_2-z_1)){\rm csch}^2(\sqrt{n_0}(z_2-z_1)),
\label{EOM two soliton1} \\
\frac{d^2 z_2}{dt^2} &=& -\omega_{\rm ds}^2 z_2 + \frac{1}{2}k V_0\sin(2kz_2) +2n_0^{3/2}\coth(\sqrt{n_0}(z_2-z_1)){\rm csch}^2(\sqrt{n_0}(z_2-z_1)),
\label{EOM two soliton2}
\end{eqnarray}
which, accordingly, leads
to the following system
for the fixed points ($z^{(1)}_{\rm o}$, $z^{(2)}_{\rm o}$):
\begin{eqnarray}
z^{(1)}_{\rm o} &=&
\frac{V_0 k}{2\omega_{\rm ds}^2}\sin(2kz^{(1)}_{\rm o})-\frac{n_0^{3/2}}{\omega_{\rm ds}^2} \coth(\sqrt{n_0}(z^{(2)}_{\rm o}-z^{(1)}_{\rm o})){\rm csch}^2
(\sqrt{n_0}(z^{(2)}_{\rm o}-z^{(1)}_{\rm o})),
\label{fixpoint two sol1}\\
z^{(2)}_{\rm o} &=&
\frac{V_0 k}{2\omega_{\rm ds}^2}\sin(2kz^{(2)}_{\rm o})+\frac{n_0^{3/2}}{\omega_{\rm ds}^2} \coth(\sqrt{n_0}(z^{(2)}_{\rm o}-z^{(1)}_{\rm o})){\rm csch}^2
(\sqrt{n_0}(z^{(2)}_{\rm o}-z^{(1)}_{\rm o})).
\label{fixpoint two sol2}
\end{eqnarray}
Thus, the fixed points for the two soliton states depend on the density of the background cloud and thereby on the chemical potential.
Due to the presence of the external potentials, the density of the background is inhomogeneous. In our analysis, we replace the inhomogeneous
density, with the density of the background at $z=0$ using the TF approximation; namely, $n_0\rightarrow n_0(z=0)=((\mu-V_0)^2-1)/4$.
Considering small deviations from the fixed points, i.e.,
$z_i=z^{(i)}_{\rm o}+\epsilon \eta_i \exp(i\omega t)$ (with $i=1,2$), leads to the eigenvalue problem for the normal modes:
\begin{equation}
\omega^2\mathbf\eta = \left(
\begin{array}{cc}
A(z^{(1)}_{\rm o})+B(z^{(1)}_{\rm o},z^{(2)}_{\rm o})& -B(z^{(1)}_{\rm o},z^{(2)}_{\rm o})
\\
-B(z^{(1)}_{\rm o},z^{(2)}_{\rm o}) & A(z^{(2)}_{\rm o})+B(z^{(2)}_{\rm o},z^{(1)}_{\rm o})
\\
\end{array} \right)\mathbf{\eta}
\label{2solitonfreqs}
\end{equation}
with
\begin{eqnarray}
A(x)&=&\omega_{\rm ds}^2-V_0 k^2\cos(2kx) \\
B(x,y)&=& 4 n_0^2 \coth^2(\sqrt{n_0} (x - y)) {\rm csch}^2(\sqrt{n_0} (x - y)) +
2 n_0^2 {\rm csch}^4(\sqrt{n_0} (x - y)).
\end{eqnarray}
Diagonalizing the above matrix, we obtain the eigenfrequencies of the two
anomalous modes of the two-soliton states.
Notice that since different two soliton states lead to
different fixed points,
the oscillation frequencies for the two solitons will differ as well.
Negative eigenvalues of this $2 \times 2$ system lead to imaginary frequencies describing unstable states;
however, as mentioned in the previous section,
one cannot describe oscillatory instabilities.
\begin{figure}[htbp]
\begin{center}
\includegraphics[width=8.5cm]{./fig6a.eps}
\end{center}
\vspace{-0.5cm}
\includegraphics[width=6cm,angle=270]{./fig6b.eps}
\caption{(Color online) Top panels: Fixed points (lines) of the system
(\ref{fixpoint two sol1})-(\ref{fixpoint two sol2})
and position of the solitons (circles for the symmetric two-soliton state,
$\times$ for the two-soliton branch with one soliton near the center
and one in one well, and stars for the state with both dark solitons
in the same well)
calculated by the numerical integration of Eq.~(\ref{gerbier})
for $\mu=3.8 \times 10^{-12}$~eV.
Bottom panels: BdG spectra of the two-soliton states for $\mu=3.8 \times 10^{-12}$~eV as a function of $V_0$; left corresponds to the symmetric two-soliton
state, middle to the always unstable state with one soliton near the center
and one in one well, while the right corresponds to the state
with both dark solitons in the same well. Notice also the agreement of
the theoretically predicted modes from Eq.
(\ref{2solitonfreqs}) with the numerical ones.}
\label{Fig: 6}
\end{figure}
In the top panel of Fig.~\ref{Fig: 6}, we show the fixed points of the system (\ref{fixpoint two sol1})-(\ref{fixpoint two sol2}) and the zero points of the solutions with
two nodes of Eq.~(\ref{gerbier}) corresponding to the positions of the two solitons as a function of $V_0$ (the chemical potential is fixed:
$\mu=3.6 \times 10^{-13}$~eV). For small $V_0$ there exists only the
symmetric solution with two nodes.
However, for $V_{0,cr}=3.4 \times 10^{-12}$~eV,
two asymmetric states appear stemming from a saddle-node bifurcation. For all three states, the agreement between the fixed points obtained from
Eqs.~(\ref{fixpoint two sol1})-(\ref{fixpoint two sol2}) and the position of the solitons is excellent.
The bottom panels of Fig.~\ref{Fig: 6} show the BdG spectra for the symmetric and the two asymmetric two soliton states as a function of $V_0$
for fixed $\mu=3.8 \times 10^{-12}$~eV. The solid lines correspond to the normal modes of Eqs.~(\ref{EOM two soliton1})-(\ref{EOM two soliton2})
around the corresponding fixed points. For a small optical lattice strength $V_0$, the two anomalous modes of the symmetric states have a different
magnitude, which indicates that the solitons are coupled. However, for increasing $V_0$, this difference decreases and, thus, the coupling between
the soliton becomes weaker. Notice that for an optical lattice strength $V_0 \simeq 3.6\times 10^{-12}$ the two modes become of the same order, a fact
that indicates that the two solitons become eventually decoupled.
The prediction of Eqs.~(\ref{EOM two soliton1})-(\ref{EOM two soliton2}) agrees very well with the numerical results obtained via the
BdG analysis for the anomalous modes.
The second panel shows the BdG spectrum of the asymmetric state with one soliton located approximately at the center of the barrier. The latter, leads to
a purely imaginary mode with an increasing magnitude with increasing $V_0$. The increase of the magnitude of this mode results from the increase
of the height of the barrier. For large $V_0$ the prediction of Eqs.~(\ref{EOM two soliton1})-(\ref{EOM two soliton2}) deviates from the result
obtained in the framework of Eq.~(\ref{gerbier}): the contributions of the perturbation terms are not of the same order and, thus, perturbation
theory fails in this case. The third panel shows the BdG spectrum for the second asymmetric state with two solitons in one well. The difference
between the modes is always large, a fact indicating the strong coupling between the two solitons. With increasing lattice strength $V_0$, the
eigenfrequencies of the anomalous modes increases too. The predictions of Eqs.~(\ref{EOM two soliton1})-(\ref{EOM two soliton2}) agree once again well with the BdG results; nevertheless, as in previous cases, Eqs.~(\ref{EOM two soliton1})-(\ref{EOM two soliton2}) cannot
predict oscillatory instabilities.
\begin{figure}[htbp]
\begin{minipage}[c]{8 cm}
\includegraphics[width=5.5cm,angle=270]{./fig7a.eps}
\end{minipage}
\begin{minipage}[c]{8 cm}
\includegraphics[width=5.5cm,angle=270]{./fig7b.eps}
\end{minipage}
\caption{(Color online) BdG spectra for different values of the optical lattice strength; from left to right,
$V_0=0$, $V_0=8.3\times10^{-14}$ eV, $V_0=1.65\times10^{-13}$ eV and $V_0=4.96\times10^{-13}$ eV.
The
solid lines are the predictions of Eqs.~(\ref{EOM two soliton1})-(\ref{EOM two soliton2}) with a numerically determined oscillation frequency ($\omega_{\rm ds}$).}
\label{Fig: 7}
\end{figure}
Next, let us consider a continuation with respect to the chemical potential.
Fig. \ref{Fig: 7} shows the BdG spectra, and the predicted oscillation frequencies, for the symmetric two soliton state for different values
of the optical lattice strength:
$V_0=0$, $V_0=8.3\times10^{-14}$ eV, $V_0=1.65\times10^{-13}$ eV and $V_0=4.96\times10^{-13}$ eV.
In the case of a purely harmonic potential, $V_0=0$, one observes the constant eigenfrequency of the dipolar mode at $\omega_z$, as well as the eigenfrequency of one anomalous mode bifurcating from the dipolar mode in the linear limit. The eigenfrequency at $2\omega_z$ is imaginary due to a degeneracy of a negative and a positive Krein sign mode. The former one corresponds to the second anomalous mode and the latter one to the quadrupole mode. This degeneracy is lifted for increasing chemical potential. The predictions of Eqs.~(\ref{EOM two soliton1})-(\ref{EOM two soliton2}) agree very well with the BdG results for the anomalous modes in the case of large chemical potentials.
For $V_0=8.3\times10^{-14}$ eV the degeneracy between the anomalous modes and the dipolar and the quadrupole mode is lifted in the linear limit.
Thus, there exists an energy gap between the anomalous mode and the dipolar mode and, more importantly, an energy gap between the second anomalous
mode and the quadrupole mode. A particularly useful conclusion stemming from the above findings is that the two soliton state can be stabilized in
a harmonic trap by adding a small optical lattice.
For $V_0=1.65\times10^{-13}$ eV the system is stable as well. The anomalous modes decrease with increasing chemical potential and reach similar values
for a large chemical potential. This denotes that the two solitons are decoupled for a large chemical potential. This can be understood at least in the
framework of the effective potential (\ref{eff. pot. two solitons}): the range of the interaction potential scales with the density $n_0$ of the
background atom cloud
and thereby with the chemical potential as well. Thus, for a large chemical potential, i.e., a large density, the interaction between solitons is strong but of short range too.
For $V_0=4.96\times10^{-13}$ eV the anomalous modes have similar
values even in the linear limit. Thus,
the two solitons are already almost decoupled
in the linear limit due to their large separation. The two anomalous modes collide with the dipolar and quadrupole modes for increasing chemical
potential, leading to oscillatory instabilities, and bifurcate from these modes again for a larger chemical potential. Although the predictions of
Eqs.~(\ref{EOM two soliton1})-(\ref{EOM two soliton2}) agree qualitatively with the BdG results, there is no quantitative agreement due to the
crossings and collisions of the soliton modes with the modes of the condensate.
\section{Dark soliton dynamics: numerical results}
\label{section: Dynamics}
So far, we have examined the existence, stability and bifurcations of the
branches of dark-solitonic states in the double-well potential setting.
We now proceed to investigate the dynamics of these states by
direct numerical integration of Eq. (\ref{gerbier}),
using as initial conditions stationary soliton states, perturbed along the
direction of eigenvectors associated with particular eigenfrequencies
(especially ones associated with instabilities).
To better explain the above, we should recall that anomalous modes are
connected to the positions and oscillation frequencies of the solitons;
therefore, perturbations of stationary dark soliton states along the
directions of the eigenvectors of the anomalous modes
result in the displacement of the positions of the solitons, thus giving rise
to interesting soliton dynamics.
Before proceeding with the dynamics of the (excited) soliton states, let us make a remark concerning the ground state of the system.
As mentioned above, the value of the lowest BdG mode of the ground state is given, in the linear limit, by the energy difference
between the ground state and the first excited state. However, this energy gap determines the oscillation frequency of atoms between
the wells of the double-well potential,
which suggests that the lowest BdG mode is connected to an oscillation between the wells.
Let us define the atom numbers $N_{\rm L}$ and $N_{\rm R}$ in the left and right well, respectively, as
$N_{\rm L}=\int_{-\infty}^0 |\psi|^2 dz$ and $N_{\rm R}=\int_{0}^{+\infty} |\psi|^2 dz$
(the total number of atoms is $N=N_\text{L}+N_\text{R}$). In Fig \ref{Fig: 8a}, we show the time evolution of $N_\text{L}$ (solid line)
and $N_\text{R}$ (dashed line) for the ground state, when perturbed by the eigenvector corresponding to the lowest BdG mode (parameter values are
$\mu=3 \times 10^{-12}$ eV and $V_0=1.12 \times 10^{-12}$ eV). One can clearly observe the oscillation of the atom numbers in the different wells, with
a characteristic frequency known as Josephson oscillation
frequency~\cite{smerzi}, determined by the magnitude of the
respective BdG eigenmode.
\begin{figure}[htbp]
\includegraphics[width=5.5cm,angle=270]{./fig8.eps}
\caption{(Color online) Time evolution of the number of atoms $N_\text{L}$ (solid line)
and $N_\text{R}$ (dashed line) in the left and right well of the double-well potential, respectively,
for the ground state of the system perturbed by the lowest BdG mode. Parameter values are
$\mu=3 \times 10^{-12}$ eV and $V_0 =1.12 \times 10^{-12}$ eV.}
\label{Fig: 8a}
\end{figure}
\begin{figure}[htbp]
\includegraphics[width=5.5cm,angle=270]{./fig9.ps}
\caption{(Color online) Contour plot showing the space-time
evolution of the density of the condensate, which carries the symmetric
one-soliton state; dark corresponds to minimal and white to maximal density.
The soliton is unstable and departs from the trap center after a short time period,
performing subsequently large amplitude oscillations.
Parameter values are $\mu=3 \times 10^{-12}$ eV and $V_0=4.96 \times 10^{-13}$.}
\label{Fig: 8}
\end{figure}
Let us now proceed with the one-soliton states.
Fig.~\ref{Fig: 8} shows the time evolution of the symmetric one soliton state perturbed by the eigenvector corresponding to
the unstable mode, for $\mu=3 \times 10^{-12}$ eV and $V_0=4.96 \times 10^{-13}$ eV. The instability manifests itself as a drift of the soliton from the trap center to the rims of the background atom cloud, where
the soliton is reflected back and forth. One clearly observes that the soliton is located at an unstable position, since a small perturbation leads
to a large amplitude oscillation of the soliton.
\begin{figure}[htbp]
\includegraphics[width=5.5cm,angle=270]{./fig10.ps}
\caption{(Color online) Same as in Fig.~\ref{Fig: 8}, but for the asymmetric
one-soliton state,
for $\mu=3 \times 10^{-12}$ eV for $V_0=8.27 \times 10^{-13}$ eV.}
\label{Fig: 9}
\end{figure}
On the other hand, Fig.~\ref{Fig: 9} shows the time evolution of the asymmetric one soliton state perturbed by the eigenvector corresponding to the
unstable mode for $\mu=3 \times 10^{-12}$ eV for
$V_0=8.27 \times 10^{-13}$ eV. The corresponding state is subject to an oscillatory instability, which
results in a growing amplitude of the soliton oscillation. However, in comparison to the previous case, the magnitude of the instability is much smaller
and, thus, the increase of the amplitude is smaller (for the same time scale) as well.
Next, we consider the multi-soliton states and start with the
symmetric two-soliton state which, for
$\mu=3.3 \times 10^{-12}$ eV and $V_0=8.27 \times 10^{-14}$ eV, is found to be linearly stable.
Numerical simulations, using as initial conditions this state perturbed by the eigenvectors corresponding to both anomalous modes,
reveal that the two solitons perform oscillations with a constant amplitude and different frequencies.
Specifically, the mode with smaller amplitude performs an in-phase oscillation (with the two solitons moving towards the same direction),
whereas the mode with larger amplitude performs an out-of-phase oscillation (with the two solitons moving towards opposite directions)
with a larger oscillation frequency. The existence of two different oscillation frequencies indicates the coupling of the dark solitons.
This situation changes for a large optical lattice strength (e.g., $V_0=1.65 \times 10^{-13}$ eV): in this case, the two
dark solitons are decoupled, since the numerical simulations show that both the in-phase and out-of phase oscillations are characterized
by almost the same frequency.
\begin{figure}[htbp]
\includegraphics[width=5.5cm,angle=270]{./fig11a.ps}
\includegraphics[width=5.5cm,angle=270]{./fig11b.ps}
\caption{(Color online) Same as in Fig.~\ref{Fig: 8}, but for the symmetric
two-soliton state with in-phase (left) anomalous
mode and out-of-phase (right) imaginary
mode perturbations, for
$\mu=3.1 \times 10^{-12}$ eV and $V_0=4.96 \times 10^{-13}$ eV.
The former anomalous eigenmode is stable, while the latter
is unstable (as concluded also from the BdG analysis).}
\label{Fig: 12}
\end{figure}
Fig.~\ref{Fig: 12} shows the time evolution of the symmetric
two-soliton state for $\mu=3.1 \times 10^{-12}$ eV and
$V_0=4.96 \times 10^{-13}$~eV perturbed by the eigenvector corresponding to the anomalous mode associated with in-phase (left panel)
and by the eigenvector corresponding to the imaginary mode out-of-phase (right panel) oscillations of the dark solitons.
For these parameters, the BdG analysis predicts that the former mode is
stable while the
latter is unstable. This is confirmed by the direct simulations:
it is clearly observed that the amplitude of the out-of-phase
oscillation increases whereas the amplitude of the in-phase oscillation remains constant.
\begin{figure}[htbp]
\includegraphics[width=5.5cm,angle=270]{./fig12a.ps}
\includegraphics[width=5.5cm,angle=270]{./fig12b.ps}
\caption{(Color online) Same as in Fig.~\ref{Fig: 12}, but for the symmetric
two-soliton state, for
$\mu=2.8 \times 10^{-12}$ eV and $V_0=4.96 \times 10^{-13}$ eV.
Notice the reversal of stability of the two anomalous modes with respect
to Fig.~\ref{Fig: 12}.}
\label{Fig: 13}
\end{figure}
Finally, in Fig. \ref{Fig: 13} we show the time evolution of the symmetric two soliton state for
$\mu=2.8 \times 10^{-12}$ eV and $V_0=4.96 \times 10^{-13}$ perturbed by the eigenvector corresponding to the anomalous mode
with in-phase (left panel) and by the eigenvector corresponding to the imaginary mode out-of-phase (right panel) oscillations.
For these parameters, the simulations show that
the out-of-phase oscillation of the solitons is stable and the in-phase oscillation unstable, in agreement with the predictions of
the BdG analysis.
\section{Conclusions}
In this work, we have performed a detailed study of the statics and
dynamics of matter-wave dark solitons in a cigar-shaped
Bose-Einstein condensate (BEC) confined in a double-well potential.
The latter, was assumed to be formed as a combination of a usual
harmonic trap and a periodic (optical lattice) potential. For our analysis, we
have adopted and used an effectively 1D mean-field model,
namely a Gross-Pitaevskii (GP)-like equation with a non-cubic nonlinearity, which has been used successfully in other works to describe dark solitons
in BECs in the dimensionality crossover regime between 1D and 3D.
As a first step in our analysis, we studied the existence and stability of both the one- and multiple-dark soliton states.
In particular, starting from the respective linear problem, we have used the
continuation with respect to the chemical potential (number of atoms) to reveal
the different branches of purely nonlinear solutions, including the ground
state, as well as excited states, in the form of single or multiple dark
solitons.
We have shown that the presence of the optical lattice is crucial
for the existence of nonlinear states that do not have a linear counterpart:
these are actually asymmetric dark soliton states, with solitons located in one well (rather than in both wells) of the double-well potential, and do not
exist in the non-interacting --- linear --- limit. We have systematically studied the bifurcations of the various branches of solutions and
showed how each particular state emerges or disappears for certain chemical potential thresholds; the latter were determined analytically,
in some cases, by using a Galerkin-type approach, in very good agreement with the numerical results. For each branch, we have also studied the
stability of the pertinent solutions via a Bogoliubov-de Gennes (BdG) analysis. This way, we have discussed the role of the anomalous
(negative energy) modes in the excitation spectra and revealed two types of
instabilities of dark solitons, corresponding to either purely imaginary
or genuinely complex eigenfrequencies; the latter were found to occur due to
the collision of an anomalous mode with a positive energy mode and
are associated with the so-called oscillatory instabilities of dark solitons.
We have also adopted a simple analytical model to study small-amplitude oscillations of the one- and multiple-dark-soliton states. Particularly,
equations of motion for the single and multiple solitons were presented in the weakly-interacting limit (where the model becomes the usual
cubic 1D GP equation), and modified them to take into regard the dimensionality of the system (i.e., the effect of transverse directions). The analysis
of these equations of motion led to the determination of the characteristic soliton frequencies, which were then compared to the eigenfrequencies
found in the framework of the BdG analysis; the agreement between the two
was generally found to be very good. Perhaps even more importantly, the
simple physical (i.e., particle) picture we adopted allowed us to reveal the
role of the optical lattice strength (i.e., the height of the barrier in the
double-well setting)
in the stability of the dark solitons, as well as its role on the coupling
between solitons. Particularly, it was found that sufficiently strong
barriers lead to instability of symmetric and asymmetric soliton states,
while it results in an effective decoupling of multi-solitons: the coupling
between neighboring solitons as described by an effective repulsive
potential, becomes negligible for sufficiently strong barriers.
We should also note that generally the states possessing a dark soliton
at the center of the trap were found to be most strongly unstable,
due to the emergence of a progressively larger (the larger the nonlinearity)
imaginary eigenfrequency.
Finally, we have performed systematic numerical simulations, based on direct
numerical integration of the quasi-one-dimensional model, to investigate the
dynamics of dark solitons in the double-well setup. We studied the manifestation of instabilities, when present, and found in all cases,
a very good agreement between the predictions based on the
BdG analysis and the numerical results. Instabilities have been observed
to manifest themselves through the exponential divergence of the soliton
from its initial location (when associated with an imaginary pair of
eigenfrequencies)
or the oscillatory growth (when associated with a complex quartet of
eigenfrequencies).
An interesting direction for a future study would be a similar analysis in a
higher-dimensional setting: this would refer to vortices confined
in double-well potentials (or more complex ``energy surfaces'' such
as quadruple-well potentials \cite{chenyu2d})
in higher-dimensional BECs. Such studies are
presently in progress.
\section*{Acknowledgments.}
The work of D.J.F. was partially supported by the Special Account for Research Grants of the University of Athens. P.G.K. gratefully acknowledges
support from NSF-DMS-0349023 (CAREER) and NSF-DMS-0806762, as well as from
the Alexander von Humboldt Foundation.
|
\section{Introduction}
The nuclear shell model was proposed almost sixty years ago \cite{Gop49,Hax49}. Soon afterwards, the interacting shell model (ISM) was developed by Lane \cite{Lan55}, Kurath \cite{Kur56}, and others (see \cite{Cau05} for a recent review). The ISM-based description of an evolution of the nucleonic coupling scheme from the $LS$ to $jj$ coupling with increasing mass number \cite{Lan55,Kur56,Wil57}, provided foundations of modern nuclear structure theory and helped to understand and categorize a wealth of data on nuclear levels, moments, collective excitations, electromagnetic and $\beta$ decays, and various particle decays \cite{Boh75}.
In its traditional form, ISM describes the nucleus as a closed quantum system: nucleons occupy bound, hence well localized, single-particle orbits of an infinite (e.g., harmonic oscillator) potential and are isolated from the environment of unbound scattering states that are not square integrable. Since the scattering continuum is not considered explicitly, the presence of branch points (decay thresholds) and double-poles of the scattering matrix ($S$-matrix) is neglected. The divide between the discrete states and the scattering continuum, i.e., the focus on one or another, has unfortunately become a kind of paradigm. In the long term, this has led to an artificial separation of nuclear structure from nuclear reactions, and hindered a deeper understanding of nuclear properties. Indeed, many structural properties of the nucleus are determined by means of nuclear collisions. Hence, the knowledge of nuclear structure depends on nuclear reactions and {\em vice versa}, and this cries out for a unified theoretical framework.
The first attempt in this direction came from Feshbach \cite{Fes58,Fes62} who expressed the collision matrix of the optical model using matrix elements of the Hamiltonian. This development gave a strong push to the ISM approach to nuclear reactions \cite{Bre59,Rod61,Mac64,Fan61} (see introduction in Ref. \cite{Bar73} for a detailed historical account) and led to various formulations of the continuum shell model (CSM) \cite{Mah69,Bar77,Phi77,Rot78,Ben99,Vol06}. A modern version of CSM in the Hilbert space, the Shell Model Embedded in the Continuum (SMEC) \cite{Ben99,Rot06}, provides a unified description of the structure and reactions with up to two nucleons in the scattering continuum using realistic ISM Hamiltonians. Nevertheless, the fully symmetric treatment of bound and scattering states in the multiparticle wave function is still too ambitious a goal \cite{Oko03}.
A different attempt to formulate the ISM for open quantum systems has been proposed recently \cite{Mic02,Idb02,Mic03,Mic04} within the Berggren ensemble \cite{Ber68}. The resulting complex-energy open quantum system extension of the ISM, the Gamow Shell Model (GSM), can be conveniently formulated in the Rigged Hilbert Space (Gel'fand triple) \cite{Gel61,Mau68}, which encompasses Gamow states \cite{Gam28,Gur29}, and is suitable for extending the quantum mechanics into the domain of time-asymmetric processes (e.g. decays). The GSM offers a fully symmetric treatment of bound, resonance, and scattering single-particle states but, until now, has been primarily used in the context of nuclear structure. (For a recent review of the complex-energy shell model, see Ref. \cite{Mic09}.)
In this paper, we shall draw attention to salient threshold effects related to generic properties of branch points and $S$-matrix poles that are expected to impact structural properties of nuclei. In Sec.~\ref{avoided}, we shall use the phase rigidity indicator to demonstrate effects of the branch point at the particle emission threshold and the exceptional point (the double-pole of the $S$-matrix) on the configuration mixing in the continuum. Section~\ref{instability} discusses the stability of ISM eigenfunctions in the neighborhood of the reaction threshold. Finally, in Sec.~\ref{unification} we outline necessary developments to achieve a unified theory of structure and reactions offering a symmetric treatment of bound, resonance, and scattering states.
\section{Salient near-threshold phenomena}
At low excitation energies, well-bound nuclei can be considered as closed quantum systems, well described by the standard ISM or its modern versions such as the no-core shell model \cite{Nav96,Nav00,Rot07}. Moving towards drip lines, or higher in excitation energy, the continuum coupling becomes gradually more important, changing the nature of weakly bound states. (Properties of unbound states are impacted by couplings to reaction channels.) In this regime, the chemical potential has a magnitude similar to the pairing gap; hence, the system is dominated by many-nucleon correlations which no longer can be considered as small perturbations atop the average potential \cite{Dob07}. Many-body states in neighboring nuclear systems with different proton and neutron numbers become interconnected via continuum, forming correlated domains (clusters) of quantum states. At present, little is known about the basic features of such clusters, e.g. a typical `cluster size' and its dependence on binding properties of participating states.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.8\linewidth]{Figure1.eps}
\caption{At low excitation energy, well-bound nuclei close to the stability valley can be described as closed quantum systems. When moving towards the particle drip lines, nuclei with different proton and neutron numbers become interconnected via continuum space, i.e., reaction channels, forming correlated domains of quantum states.}
\label{network}
\end{center}
\end{figure}
In standard network problems, like the percolation problem \cite{Sta85,Bot02}, different phases of correlated domains are formed depending on the mean network activity with respect to some threshold. Analogously, one expects that the key to understanding the formation of correlated domains of quantum states is a near-threshold behavior of a single network member. In the correlation-dominated regime of a nuclear system, the percolation threshold(s) should be identified with reaction threshold(s).
What can be said about properties of many-body systems in the narrow range of energies around the reaction threshold? Are those properties independent of any particular realization of the Hamiltonian? Below, we shall try to address these questions in the framework of SMEC, but essentials of this discussion are valid in the GSM approach as well.
\subsection{The main ingredients of the continuum shell model}
To facilitate further discussion, we shall now recall fundamentals of SMEC.
The Fock space of an $A-$particle system consists of the set of square-integrable functions
${\cal Q}\equiv \{\Phi_i^{(A)}\}$, internal space described by the standard ISM, and the space of embedding scattering states ${\cal P}\equiv \{\zeta_E^{c}\}$. These two orthogonal and complementary subspaces of the Fock space are obtained by the projection formalism. In one-particle continuum problems, reaction channels `$c$' correspond to the motion of an unbound nucleon in a state $(lj)$ relative to the daughter nucleus in some ISM eigenstate $\Phi_k^{(A-1)}$. An open quantum system description of `internal' dynamics, i.e., in ${\cal Q}$, includes couplings to the
`environment' of scattering states and decay channels and is given by the energy-dependent effective Hamiltonian:
\begin{eqnarray}\label{HQQ}
{\cal H}_{QQ}(E)=H_{QQ}+W(E),
\end{eqnarray}
where $H_{QQ}$ is the ISM Hamiltonian (the closed quantum system Hamiltonian) in the internal space ${\cal Q}$ and
\begin{eqnarray}\label{coupl}
W(E)=H_{QP}G_P^{(+)}(E)H_{PQ}
\end{eqnarray}
is the energy-dependent term describing both virtual particle excitations and irreversible decays to the environment of reaction channels. In the latter expression, $G_P^{(+)}(E)$ is the Green's function for the motion of a single nucleon in the continuum, and $E$ is the nucleon's energy (i.e., the scattering energy). The external mixing of two ISM states $i$ and $j$ due to $W(E)$ consists of the Hermitian principal value integral $W^R(E)$ describing virtual excitations to the continuum and the anti-Hermitian residuum $W^I(E)=-(i/2){\bf VV}^T$ in a dyadic form that represents decay out of the internal space ${\cal Q}$. The $M\times\Lambda$ matrix ${\bf V}=\{V^c_i\}$ denotes the amplitudes connecting the ISM state $\Phi_i$ ($i=1,\cdots,M$) to the reaction channel $c$ ($c=1,\cdots,\Lambda$) \cite{Oko03}.
At energies below the lowest reaction threshold, the effective Hamiltonian is Hermitian, i.e., $W^I=0$. Above the first threshold, the non-Hermitian part of ${\cal H}_{QQ}(E)$ describes the irreversible decay from the internal space. The effective Hamiltonian in this case becomes complex symmetric. Each eigenstate of ${\cal H}_{QQ}(E)$ is coupled to states in neighboring nuclei via a network of reaction channels, either closed or open. The contribution of different reaction channels to the total continuum coupling is highly non-uniform and spans over a considerable range of excitation energies \cite{Oko07}.
Since the Hamiltonian (\ref{HQQ}) depends explicitly on the energy, it is highly non-linear. Moreover, the continuum-coupling term generates effective many-body interactions in the internal space, even if it is originally two-body in the full space.
Effects of the resulting many-body interactions have been studied by considering the continuum-coupling correction to the binding energy for the chains of oxygen and fluorine isotopes \cite{Oko07,Luo02}. It has been found that the induced many-body interactions explain a significant shift in the neutron drip line for fluorine isotopes as compared to the oxygens. Another explanation, explicitly invoking effective three-body interaction, has recently been suggested in Ref.~\cite{Ots09}. These two scenarios could be difficult to distinguish.
Eigenstates of the open quantum system Hamiltonian ${\cal H}_{QQ}(E)$ are biorthogonal. The left $|\Psi_j\rangle$ and right $|\Psi_{\tilde j}\rangle$ eigenstates have complex conjugate wave functions. The Hermitian conjugation of ${\cal H}_{QQ}$ switches the left and right vectors. The orthonormality condition in the biorthogonal basis can be written as $\langle\Psi_{\tilde j}|\Psi_k\rangle=\delta_{jk}$.
The diagonalization of the effective SMEC Hamiltonian can be achieved by means of an orthogonal transformation. The resulting eigenvalues are real if all reaction channels are closed. Above the threshold, the transformation becomes non-unitary:
\begin{eqnarray}
\label{transf}
{\Phi}_i \longrightarrow \Psi_j = {\sum}_{i}^{} b_{ji}{\Phi}_i
\end{eqnarray}
and it yields complex eigenvalues. Physical resonances can be identified with narrow poles of the
$S$-matrix
\cite{Sie39,Ber68,Mic09}, or using the Breit-Wigner approach, which leads to a fixed-point condition \cite{Vol06,Oko03,Mad05}.
\subsection{How to monitor effects of branch points and avoided level crossings in the complex energy plane?}\label{avoided}
The energy-dependent term (\ref{coupl}) describing the coupling of closed quantum system eigenfunctions $\Phi_i$ to the environment of scattering states does not act uniformly on all ISM states. A useful indicator of effects caused by an anti-Hermitian coupling term $W^I(E)$ on biorthogonal eigenfunctions $\{\Psi_k\}$ of the effective Hamiltonian is the phase rigidity \cite{Lan97,Bro03,Bul06}:
\begin{eqnarray}
r_j=\frac{\langle\Psi_{\tilde j}|\Psi_j\rangle}{\langle\Psi_j|\Psi_j\rangle}~ \ ,
\label{rig}
\end{eqnarray}
given by the ratio of biorthogonal and Hermitian norms of an eigenfunction. The indicator (\ref{rig}) can also be written in a form \cite{Bul06}:
\begin{eqnarray}
r_j=e^{2i\theta_j}\frac{\int dr \left(|{\rm Re}{\bar \Psi}_j(r)|^2-|{\rm Im}{\bar \Psi}_j(r)|^2\right)}{\int dr \left(|{\rm Re}{\bar \Psi}_j(r)|^2+|{\rm Im}{\bar \Psi}_j(r)|^2\right)} ~ \ ,
\label{rig1}
\end{eqnarray}
which better illustrates its physical meaning. In this expression, the angle $\theta_j$ arises from a transformation of $\Psi_j$ so that ${\rm Re}{\bar \Psi}_j$ and ${\rm Im}{\bar \Psi}_j$ are orthogonal and it characterizes the degree to which the eigenfunction $\Psi_j$ is complex.
Phase rigidity varies between 1 and 0. It equals 1 for bound state eigenfunctions. For unbound states, the condition $r_j$=1 means that the non-Hermitian coupling $W^I(E)$ exerts a negligible effect on the structure of an eigenfunction, i.e., its biorthogonal and Hermitian norms are identical. The energy variation of $r_j$ is an internal property of a considered open quantum system and measures a mutual influence of the neighboring resonances.
\begin{figure}[htb]
\begin{center}
\includegraphics[scale=0.42]{Figure2.eps}
\caption{The energy dependence of the phase rigidity for the wave functions of the $0^+_1$ and $0^+_2$ SMEC states of $^{16}$Ne. The dashed curve shows the difference between energies of these states in the complex energy plane.}
\label{rigid}
\end{center}
\end{figure}
Figure~\ref{rigid} shows the phase rigidity as a function of the scattering energy $E$ for the two lowest-energy $0^+$ levels of the effective Hamiltonian of $^{16}$Ne. These SMEC calculations are performed in a $p_{1/2}, d_{5/2}, s_{1/2}$ model space. For $H_{QQ}$ we use the ZBM Hamiltonian \cite{Zuk68}. The residual coupling $H_{QP}$ between the internal space and the surrounding continuum is generated by the contact force: $H_{QP}=H_{PQ}=V_0\delta(r_1-r_2)$ with $V_0=-1100$ MeV fm$^3$. The $0^+$ ISM eigenstates are coupled through
the common {\em physical} one-proton emission channels $[^{15}{\rm F}(I^{\pi})\otimes {^1{\rm H}}(lj)]_{E'}^{J^{\pi}}$ with $I^{\pi}=1/2^+, 5/2^+$, and $1/2^-$ which have thresholds at $E=0$ (elastic channel), 0.67 MeV, and 2.26 MeV, respectively. (Each physical channel, specified by a states $I^{\pi}$ and $J^{\pi}$ in daughter $^{A-1}X$ and parent $^AY$ nuclei, respectively, may correspond to a number of different mathematical channels characterized by different $(lj)$-values of a nucleon which are allowed by the coupling of $I^{\pi}$ and $J^{\pi}$.)
These are all possible one-proton emission channels in $^{16}$Ne, described in ZBM space.
One can notice two local variations of the phase rigidity for $0_1^+$ and $0_2^+$ eigenfunctions. The first one appears close to the elastic channel threshold whereas the second one, at higher energies, indicates the avoided level crossing between $0_1^+$ and $0_2^+$ eigenvalues in the complex energy plane. Avoided level crossings are traces of exceptional points \cite{Kat95,Zir83,Hei91} in the parameter space of the effective Hamiltonian \cite{Her03,Oko09}.
The dashed curve in Fig. \ref{rigid} shows a distance in the complex energy plane between the two $0^+$ eigenvalues. The minimum in $\Delta\varepsilon=|\varepsilon_1-\varepsilon_2|(E)$ coincides with the minimum in $r(E)$ at a position of the avoided crossing. However, neither maximum nor minimum of $\Delta\varepsilon$ coincides with the minimum of the phase rigidity around the elastic reaction threshold.
A striking example of the configuration mixing caused by the coupling term $W(E)$ is shown in Fig. \ref{rigid3}. The phase rigidity
for $2_i^+$ ($i=1,\cdots,4$) states in $^{16}$Ne is plotted here for $V_0=-1357$ MeV fm$^3$. The $2^+$ states are interconnected via the coupling to the inelastic channel: $[^{15}{\rm F}(5/2^{\pi})\otimes {^1{\rm H}}(lj)]_{E'}^{2^{+}}$. One can clearly see that at this value of the continuum coupling strength there appears the degeneracy (exceptional point) of $2_3^+$ and $2_4^+$ levels at which the phase rigidity drops abruptly to 0. The remaining $2^+$ eigenfunctions are mere spectators of this violent wave function rearrangement.
\begin{figure}[htb]
\begin{center}
\includegraphics[scale=0.42]{Figure3.eps}
\caption{The energy dependence of the phase rigidity for lowest $2_i^+$ ($i=1,\cdots,4)$ eigenfunctions of the SMEC Hamiltonian in $^{16}$Ne. The $2_3^+$ and $2_4^+$ eigenvalues form an exceptional point at $E \simeq 4.8$ MeV.}
\label{rigid3}
\end{center}
\end{figure}
At low energies, close to the branch point at the channel threshold ($E$=0.67 MeV), one can see a strong mixing of $2_1^+$ and $2_2^+$ eigenfunctions.
\subsection{Near-threshold instability of the wave function}\label{instability}
Both the branch points associated with reaction thresholds and the avoided level crossings are essential elements of the configuration mixing in open quantum systems. Abrupt variations of the phase rigidity as a function of the energy for certain eigenstates of the effective Hamiltonian are indicative of wave function instability. Another indicator of such instabilities could be the continuum-coupling energy correction, $E_{\rm corr}(E)=\langle\Phi_i|W(E)|\Phi_i\rangle$, to the ISM eigenvalue $\langle\Phi_i|H_{QQ}|\Phi_i\rangle$. Figure~\ref{ccorr} (left) shows the continuum-coupling correction
for the ground states of oxygen isotopes calculated for a single channel corresponding to a neutron coupled to the ground state in the daughter nucleus $^{A-1}$O.
\begin{figure}[htb]
\begin{center}
\includegraphics[scale=0.6]{Figure4.eps}
\caption{The continuum-coupling energy correction $E_{\rm corr}(E)$. The value $E$=0 corresponds to the first (one-neutron emission) threshold. For $H_{QQ}$, we take USD Hamiltonian for the $sd$ shell \cite{Wil84} and KB' interaction for the $pf$ shell \cite{Pov81}. The cross-shell interaction is the G-matrix \cite{Kah69}. The coupling to continuum is given by the Wigner-Bartlett interaction with the strength $V_0=-414$ MeV fm$^3$.
Left: ground state configurations of $^{23-29}$O. Only a contribution from
a single channel corresponding to a neutron coupled to the ground state in the daughter $^{A-1}$O nucleus is included.
Right: the ground state configuration of $^{27}$O. Contributions from couplings to various positive parity states in $^{26}$O are indicated.
}
\label{ccorr}
\end{center}
\end{figure}
One can see that the continuum-coupling energy correction does depend on wave functions involved in the coupling matrix element, i.e., the wave function $\Psi_i$ in the parent nucleus $[N,Z]$ and the channel wave function $[_ZX_{N-1}(I^{\pi})\otimes{^1{\rm n}}(lj)]^{J^{\pi}}_{E'}$. The correction
$E_{\rm corr}$ varies strongly in a broad interval of energies around the neutron emission threshold:
from $\pm4$ MeV in $^{28}$O to $\pm2$ MeV in $^{26}$O and $^{24}$O.
As a result of the centrifugal barrier, a maximum of $|E_{\rm corr}|$ is shifted above the threshold if the angular momentum $l$ of a neutron is different from zero. For $l$=0, the first derivative of $|E_{\rm corr}(E)|$ is discontinuous at the neutron emission threshold. A similar shift can be caused by a Coulomb barrier if the continuum coupling involves charged particle emission channels.
Another interesting observation is the odd-even staggering of $E_{\rm corr}$ \cite{Luo02}: a blocking of the virtual scattering to the particle continuum by an odd nucleon diminishes the continuum correction to the binding energy in odd-$N$ nuclei.
Figure \ref{ccorr} (right) shows individual contributions to the continuum-coupling energy correction in
$^{27}$O coming from couplings to various reaction channels $[_8{\rm O}_{18}(I^{\pi})\otimes{^1{\rm n}}(lj)]^{J^{\pi}}_{E'}$. The largest contribution to $E_{\rm corr}$ comes from the coupling to the lowest physical inelastic channel: $[_8{\rm O}_{18}(2_1^+)\otimes {^1{\rm n}}(lj)]^{3/2^+}$. The same wave appears in the elastic channel
$[_8{\rm O}_{19}(3/2_1^+)\otimes {^1{\rm n}}(lj)]^{0^+}$ contribution to $E_{\rm corr}$ in the neighboring nucleus $^{28}$O (see Fig.~\ref{ccorr} on the left).
The angular momentum dependence and charge dependence of the continuum-coupling correction around the particle-emission threshold
can be explained in the same way as
a characteristic threshold behavior of scattering and reaction cross sections \cite{Wig48} (Wigner cusp). Due to the unitarity of the $S$-matrix and the resulting flux conservation, Wigner's threshold effect has a broad impact on various channel wave functions due to channel coupling \cite{Baz57,New59,Mey63,Baz66,Hat78}. Hence, the branch point singularity at the opening of a given reaction channel can induce non-local correlations between wave functions in distant channels. This effect has been found experimentally, for example, in the coupling of the analogous channels in $(d,p)$ and $(d,n)$ reactions \cite{Moo66}. These channel-channel correlations are yet another manifestation of the eigenfunction correlations induced by the presence of branch points or exceptional points.
Recently, it has been found that a threshold law, similar to the Wigner's law for scattering and reaction cross-sections, holds for one-nucleon overlap amplitudes \cite{Mic07}, which probe shell occupancies; hence, they characterize many-body correlations. An immediate consequence of this finding is that the influence of the branch point on the configuration mixing in weakly bound mirror states will be different, even for identical separation energies in mirror states \cite{Mic09a}.
The size of the energy correction is a measure of the mixing of unperturbed ISM wave functions due to the coupling to the reaction channel(s), either open or closed. For ISM wave functions below all reaction thresholds, only the Hermitian part of the coupling term $W(E)$ acts. For unbound ISM wave functions, the competition of Hermitian $W^R$ and anti-Hermitian $W^I$ parts is an essential ingredient of the formation mechanism of exceptional points in the spectra of resonances. As shown above, $|E_{\rm corr}|$ is both state- and system-dependent, i.e., the continuum coupling does not induce a global instability of all ISM eigenstates close to the particle emission threshold. On the contrary, for those ISM wave functions which are susceptible to the continuum coupling, the near-threshold instability is seen predominantly in a single pair of unperturbed ISM states. Other ISM states with the same quantum numbers remain spectators even though all of them are interconnected by the coupling to the same reaction channel(s). Both avoided level crossings (or exceptional points) and branch points associated with reaction thresholds induce the configuration mixing of continuum states. For weakly bound states, only branch point(s) can contribute to the configuration mixing.
In general, the mixing mechanism associated with the coupling of ISM eigenstates to the reaction channel(s) enhances the similarity of a bound, near-threshold ISM state $\Psi(J^{\pi})$ with the decay channel $[\Psi^{(d)}(I^{\pi})\otimes({\tau_z}lj)]^{J^{\pi}}$,
i.e., increases the corresponding spectroscopic factor $\langle {\tilde \Psi(J^{\pi}})||a_{lj}^+||\Psi^{(d)}(I^{\pi})\rangle^2$, where
$\Psi^{(d)}$ is the wave function of the daughter
system. (The tilde symbol above bra vector signifies that the complex conjugation in the dual space affects the angular part and leaves the radial part unchanged.)
This `alignment' of a near-threshold eigenstate of the effective Hamiltonian with the corresponding reaction channel has been demonstrated in the approximation of a one-nucleon continuum \cite{Cha06}. The phenomenon is based on three general principles: (i) the branch point nature of the reaction channel threshold, (ii) the unitarity of the $S$-matrix, and (iii) the $2\times2$ matrix structure in the eigenfunction mixing. These principles neither invoke any particular interaction in the internal space nor any special structure of eigenfunctions involved in the mutual coupling. Hence, the alignment of near-threshold eigenfunctions of the effective Hamiltonian with the reaction channel should be present in any open quantum many-body system which has complex reaction channels. In particular, the entanglement of ISM wave functions due to their coupling to a common particle emission channel should be responsible for the omnipresence of clustering phenomena in states close to their cluster decay thresholds. Indeed, as illustrated in Fig.~\ref{clusters}, $\alpha$-cluster states have been found among the ISM wave functions close to the $\alpha$-particle emission thresholds in, for example, $^{12}$C, whereas correlated two-neutron pair has been seen in the halo ground state of $^{11}$Li close to the two-neutron emission threshold.
\begin{figure}[htb]
\begin{center}
\includegraphics[scale=0.75]{Figure5.eps}
\caption{Cluster states close to decay thresholds.
Left: Excited 0$^+$ state in $^{12}$C (Hoyle state) just above the
$^{8}$Be+$\alpha$ threshold. Right: Subthreshold halo ground state of $^{11}$Li.
}
\label{clusters}
\end{center}
\end{figure}
The threshold features are generic: they are neither accidental nor arise from particular properties of the nuclear Hamiltonian.
This universality forms a foundation of Ikeda systematics \cite{Ike68} which can be extended naturally to halo phenomena, see Fig.~\ref{clusters}. In this formulation, Borromean systems correspond to an example of clustering where the cluster itself is not a self-bound system.
We are still far away from a quantitative understanding of the organization of ISM eigenfunctions near the particle emission threshold. The structure of the ground state wave function of $^{11}$Li is a particularly interesting case-study due to an intricate sequence of low-lying two- and one-neutron emission thresholds which are separated by less than 30 keV (!). This approximate degeneracy of two different emission thresholds may have important consequences on the structure of halo two-neutron pair. The relative weights of configurations, such as
$[^9{\rm Li}(3/2^-)]\otimes[{\rm n}^2(0^+)]$, corresponding to the two-neutron continuum of $^9$Li, and
$\{[^9{\rm Li}(1/2^-)]\otimes[{\rm n}(1/2^-)]\}^{1^+}\otimes[{\rm n}(1/2^-)]$,
$\{[^9{\rm Li}(1/2^-)]\otimes[{\rm n}(1/2^+)]\}^{1^-}\otimes[{\rm n}(1/2^+)]$, or
$\{[^9{\rm Li}(3/2^-)]\otimes[{\rm n}(1/2^-)]\}^{2^+}\otimes[{\rm n}(1/2^-)]$, associated with an unbound $^{10}$Li, reflects the influence of these two almost degenerate branch points on the structure of the ground state wave function of $^{11}$Li.
The exceptional points, whose dramatic effects on the phase rigidity can be seen in Fig. \ref{rigid3}, are also generic in Hamiltonian systems. They arise as a result of an interplay between Hermitian and non-Hermitian parts of the continuum coupling term $W(E)$ for energies above the elastic channel threshold. The number of exceptional points depends only on some basic features of the system, such as its dimensionality and quantum integrability \cite{Duk09}. On the contrary, the precise position of any exceptional point in the spectrum of eigenvalues of the energy-dependent effective Hamiltonian depends on the choice of the ISM Hamiltonian $H_{QQ}$, the strength of the coupling term $W(E)$, and the energy $E$ of the system. Hence, its finding provides the test of both the effective interaction and the configuration mixing in any OQS.
It has been found recently that low-energy exceptional points appear for realistic values of coupling to the continuum and hence could be accessible experimentally \cite{Oko09}. At low excitation energies, they could be seen, for example, as individual peaks associated with a jump by $2\pi$ of the elastic scattering phase shift. Also in the neighborhood of an exceptional point for avoided crossing of resonances, this characteristic imprint of the scattering phase shift remains, which gives a real chance that traces of the exceptional could actually be searched for in reaction studies.
Complex and biorthogonal eigenstates of ${\cal H}_{QQ}(E)$ provide a natural basis in which one can express the resonant part of any scattering wave function:
\begin{eqnarray}
\Upsilon^c_E=\zeta^c_E+\sum_ja_j{\tilde \Psi}_j ~ \ ,
\label{ups}
\end{eqnarray}
where $a_j=\frac{\langle\Psi_{\tilde j}|H_{QP}|\zeta^c_E\rangle}{(E-E_j)}$
and
${\tilde \Psi}_j\equiv(1+G_P^{(+)}H_{PQ})\Psi_j$.
Dominant contributions to $\Upsilon^c_E$ inside an interaction region are given by eigenfunctions $\Psi_j$ of ${\cal H}_{QQ}$, i.e.,
\begin{eqnarray}
\Upsilon^c_E\sim\sum_ja_j\Psi_j ~ \ .
\end{eqnarray}
Both branch points (reaction thresholds) and double poles of the $S$ matrix (exceptional points) lead to a non-separable entanglement of two eigenstates of the effective Hamiltonian which manifests itself in a singular behavior of matrix elements of any operator which does not commute with ${\cal H}_{QQ}$ \cite{Oko09,Duk09,Oko09b}. These unusual features are yet another facet of the profound change of the nature of ISM states close to the reaction thresholds and avoided resonance crossings which are responsible for clustering and strong mixing in OQSs.
\subsection{Future developments: happy marriage of nuclear structure and nuclear reactions}\label{unification}
To understand the formation of a network of correlated states, e.g., the chain of isotopes from a well-bound $^9$Li to an unbound $^{12}$Li, a structure of the Hoyle resonance, and the radiative capture reactions, such $^{12}$C($\alpha,\gamma$)$^{16}$O, a theoretical framework is required that would provide a unified approach to nuclear structure and reactions. Such a theoretical framework could be developed based on the GSM \cite{Mic02,Idb02,Mic03,Mic04}. This model, formulated in the Rigged Hilbert Space \cite{Gel61,Mau68} and using a complete Berggren ensemble \cite{Ber68}, is a generalization of the nuclear ISM for a description of bound states, resonances, and many-body scattering continuum. In a nuclear structure application, solutions of the GSM can be found by diagonalizing a complex-symmetric Hamiltonian matrix. The `dimensional catastrophe' in GSM when increasing the number of active particles is much more serious than in the standard ISM because each single-particle continuum state in the Berggren ensemble becomes a new shell in the many-body GSM formulation. This acute problem has been alleviated by recent progress in the generalization of the Density Matrix Renormalization Group \cite{Whi92,Whi93,Car99}
method to non-Hermitian, complex-symmetric matrix problems \cite{Rot06a,Rot09}.
Significant progress has also been made in applications of realistic interactions in GSM \cite{Hag06}. Finally, powerful techniques for a selection of physical resonances, based on the overlap method, have been developed \cite{Mic02,Mic03}. Still, much work is needed to develop appropriate effective interactions which would allow a systematic study of the structure of nuclear states in long isotopic or isotonic chains: from the valley of stability towards particle drip lines.
So far, most applications of the GSM have addressed nuclear structure phenomena. Further progress can only be achieved if the method could be fully extended to reaction problems. This can be achieved by the coupled-channel formulation of the scattering problem:
\begin{eqnarray}
\int dr'\sum_{c'}\sum_{m,n} \langle c;r|\Psi_n\rangle\left[\langle \Psi_{\tilde n}|H-E\delta_{nm}|\Psi_m\rangle\right]\langle\Psi_{\tilde m}|c';r'\rangle \Omega_{c'}(r')=0 ~ \ .
\label{cceq}
\end{eqnarray}
A similar approach, based on the Resonating Group Method \cite{Wil58,Wil59,Tan78},
has been developed in the no-core shell-model approach to nuclear scattering problems \cite{Qua08,Qua09}. In the above equation,
$H_{nm}\equiv\langle\Psi_n|H|\Psi_m\rangle$ is the GSM Hamiltonian matrix and $O_{c;n}\equiv\langle c;r|\Psi_n\rangle$ are overlaps of the reaction channel and the GSM many-body state. Formally, Eq.~(\ref{cceq}) takes the same form independently of the choice of reaction channels; the whole complication is hidden in the overlaps $O_{c;n}$. Their calculation is a formidable task for the GSM approach. It requires a complicated change of the coordinate system which makes the calculation difficult for heavier projectiles, even for reactions with two-body asymptotic conditions. Any future progress in the application of GSM to nuclear reactions is ultimately related to the progress in the development of new algorithms for fast calculation of overlaps.
\section{Acknowledgments}
We wish to thank J. Rotureau for useful discussions.
This work was supported in part by the U.S.
Department of Energy under Contract No. DE-FG02-96ER40963 (University
of Tennessee) and by the CICYT-IN2P3
cooperation.
\section*{References}
|
\section{}
In recent years different approaches, originating from the physics
community, have shed new light on sports events, e.g. by studying
the behavior of spectators \cite{laola}, by elucidating the
statistical vs. systematic features behind league tables
\cite{ben1,ben2,buch}, by studying the temporal sequence of ball
movements \cite{mendes} or using extreme value statistics
\cite{Suter,Greenhough02} known, e.g., from finance analysis
\cite{Stanley}. For the specific case of soccer matches different
models have been introduced on phenomenological grounds
\cite{Lee97,Dixon97,Dixon98,Rue00,Koning00,Dobson03}. However,
very basic questions related, e.g., to the relevance of systematic
vs. statistical contributions or the temporal fitness evolution
are still open. It is known that the distribution of soccer goals
is broader than a Poissonian distribution
\cite{Greenhough02,janke,janke09}. This observation has been
attributed to the presence of self-affirmative effects during a
soccer match\cite{janke,janke09}, {i.e. an increased
probability to score a goal depending on the number of goals
already scored by that team}.
In this work we introduce a general model-free approach which allows us to elucidate the outcome of
sports events. Combining strict mathematical
reasoning, appropriate finite-size scaling and comparison with actual data
all ingredients of this framework can be quantified for the specific example of soccer. A unique relation can be derived to calculate
the expected outcome of a soccer match and three hierarchical levels of statistical influence can be identified.
As one application we show that the skewness of the distribution of soccer goals \cite{Greenhough02,janke,janke09} can be fully
related to fitness variations among different teams and does not require the presence of self-affirmative effects.
As data basis we take all matches in the German Bundesliga (www.bundesliga-statistik.de) between seasons 1987/88 and 2007/08 except for the year 1991/92 (in that year the league contained 20 teams). Every team plays 34 matches per season.
Earlier seasons are not taken into account because the underlying statistical properties (in particular number of goals per match) are somewhat different.
Conceptually, our analysis relies on recent observations in describing
soccer leagues \cite{HR09}: (i) The home advantage is characterized by a team-independent but season-dependent increase of the home team goal
difference $c_{home}>0$. (ii) An appropriate observable to characterize the fitness of a team $i$ in a given season is the average goal
difference (normalized per match) $\Delta G_i(N)$, i.e. the difference of
the goals scored and conceded during $N$ matches. In particular it contains more information about the team fitness than, e.g., the number of points.
\begin{figure}[ht]
\includegraphics[width=0.95\linewidth]{fig1.eps}
\caption{ The correlation function $h(t)$. The average value of $h(t)$ is included (excluding the value
for $t=17$) yielding approx. $0.22$ \cite{HR09}.}
\label{fig.1}
\end{figure}
Straightforward information about the team behavior during a season can be extracted from correlating its match
results from different match days. Formally, this is expressed by
the correlation function $h(t) = \langle \Delta g_{ij}(t_0) \Delta g_{ik}(t_0+t) \rangle$. Here
$\Delta g_{ij} := g_i - g_j$ denotes the goal difference of a match of team $i$
vs. team $j$ with the final result $g_i:g_j$. $j$ and $k$ are the opponents of
team $i$ at match days $t_0$ and $t_0 + t$. The home-away asymmetry
can be taken into account by the transformation $\Delta g_{ij} \rightarrow \Delta g_{ij} \mp c_{home}$
where the sign depends on whether team $i$ plays at home or away.
The resulting function $h(t)$ is shown in Fig.1. Apart from the data point
for $t=17$ one observes a time-independent positive plateau value. The absolute value of this constant corresponds
to the variance $\sigma_{\Delta G}^2$ of $\Delta G_i$ and is thus a measure for the fitness variation in a league \cite{HR09}.
Furthermore, the lack of any decay shows that the fitness of a team is constant during the whole season.
This result is fully consistent with the finite-size scaling analysis in
Ref.\cite{HR09} where additionally the fitness change between two seasons was
quantified. The exception for $t=17$
just reflects the fact that team $i$ is playing against the same
team at days $t_0$ and $t_0 + 17$, yielding additional
correlations between the outcome of both matches (see also below).
As an immediate consequence, the limit of $\Delta G_i(N)$
for large $N$, corresponding to the true fitness $\Delta G_i$, is well-defined. A consistent estimator for $\Delta G_i$, based on the information
from a finite number of matches, reads
\begin{equation}
\label{eqgest}
\Delta G_i = a_N \Delta G_i(N).
\end{equation}
with $a_N \approx 1/[1+3/(N \sigma^2_{\Delta G})]$ \cite{HR09} .
For large $N$ the factor $a_N$ approaches unity and the estimation
becomes error-free, i.e. $\Delta G_i(N) \rightarrow \Delta G_i$.
For $N=33$ one has $a_N = 0.71$ and the variance of the estimation error is
given by $ \sigma_{e,N}^2 = (N/3+1/\sigma_{\Delta G}^2)^{-1} \approx 0.06$ \cite{HR09}. This statistical framework
is known as regression toward the mean \cite{Stigler}.
Analogously, introducing
$\Sigma G_i(N)$ as the average sum of goals scored and conceded by team $i$ in $N$ matches
its long-time limit is estimated via $\Sigma G_i -\lambda = b_N (\Sigma G_i(N)-\lambda)$ where $\lambda$ is the average number of goals
per match in the respective season. Using $\sigma^2_{\Sigma G} \approx 0.035$ one correspondingly obtains $b_{N=33} = 0.28$ \cite{HR09}.
{Our key goal is to find a sound characterization of the match result when team $i$ is playing vs. team $j$, i.e. $\Delta g_{ij}$
or even $g_i$ and $g_j$ individually. The final outcome $\Delta g_{ij}$ has three conceptually different and uncorrelated contributions
\begin{equation}
\Delta g_{ij} = q_{ij} + f_{ij} + r_{ij}.
\end{equation}
Averaging over all matches one can define the respective variances $\sigma_q^2, \sigma_f^2$ and $\sigma_r^2$.
(1) $q_{ij}$ expresses the average outcome which can be expected based on knowledge of the team fitness values $\Delta G_i$ and $\Delta G_j$, respectively.
Conceptually this can be determined by averaging over all matches when teams with these fitness values play against each other. The task is
to determine the dependence of $q_{ij} \equiv q(\Delta G_i, \Delta G_j) $ on $\Delta G_i$ and $\Delta G_j$.
(2) For a specific match, however, the outcome can be {\it systematically}
influenced by different factors beyond the general fitness values using the variable $f_{ij}$ with a mean of zero:
(a) External effects such as several players which are injured or tired,
weather conditions (helping one team more than the other), or red cards. As a consequence the effective fitness of a team relevant for this match may differ from the
estimation $\Delta G_{i}$ (or $\Delta G_{j}$). (b) Intra-match effects depending on the actual course of a match. One example is the suggested presence
of self-affirmative effects, i.e. an increased probability to score a goal (equivalently an increased fitness) depending
on the number of goals already scored by that team \cite{janke,janke09}. Naturally, $f_{ij}$ is much harder to predict if possible at all. Here we restrict
ourselves to the estimation of its relevance via determination of $\sigma_f^2$.
(3) Finally, one has to understand the emergence of the actual goal distribution based on expectation values as expressed by the random
variable $r_{ij}$ with average zero. This
problem is similar to the physical problem when a decay rate (here corresponding to $q_{ij} + f_{ij}$) has to be translated into the actual
number of decay processes. }
{\it Determination of $q_{ij}$}:
$q_{ij}$ has to fulfill the two basic conditions (taking into account the home advantage):
$q_{ij}-c_{home} = - (q_{ji}-c_{home})$ (symmetry condition) and
$\langle q_{ij}\rangle_j -c_{home} = \Delta G_i$ (consistency condition) where the average is over
all teams $j \ne i$ (in the second condition a minor correction due to the finite number of teams in a league is neglected).
The most general dependence on $\Delta G_{i,j}$ up to third order, which is compatible
with both conditions, is given by
\begin{equation}
\label{eqg1}
q_{ij} = c_{home} + (\Delta
G_i - \Delta G_j) \cdot [ 1 - c_3(\sigma^2_{\Delta G} + \Delta G_i
\Delta G_j ) ].
\end{equation}
Qualitatively, the $c_3$-term takes into account the possible
effect that in case of very different team strengths (e.g. $\Delta
G_i \gg 0$ and $\Delta G_j \ll 0$) the expected goal difference is even more
pronounced ($c_3 > 0$: too much respect of the weaker team) or
reduced ($c_3 < 0$: tendency of presumption of the better team).
On a phenomenological level this effect is already considered in
the model of, e.g., Ref.\cite{Rue00}. The task is to determine the
adjustable parameter $c_3$ from comparison with actual data. We
first rewrite Eq.\ref{eqg1} as $q_{ij} - (\Delta G_i - \Delta G_j)
- c_{home} = - c_3(\Delta G_i - \Delta G_j)(\sigma^2_{\Delta G} +
\Delta G_i \Delta G_j )$. In case that $\Delta G_{i,j}$ is known
this would correspond to a straightforward regression problem of
$\Delta g_{ij} - (\Delta G_i - \Delta G_j) - c_{home}$ vs.
$-(\Delta G_i - \Delta G_j)(\sigma^2_{\Delta G} + \Delta G_i
\Delta G_j )$. An optimum estimation of the fitness values for a
specific match via Eq.\ref{eqgest} is based on $\Delta
G_{i,j}(N)$, calculated from the remaining $N=33$ matches of both
teams in that season . Of course, the resulting value of
$c_3(N=33)$ is still hampered by finite-size effects, in analogy
to the regression towards the mean. This problem can be solved by
estimating $c_3(N)$ for different values of $N$ and subsequent
extrapolation to infinite $N$ in an $1/N$-representation. Then our estimation
of $c_3$ is not hampered by the uncertainty in the determination of $\Delta
G_{i,j}$. For a
fixed $N \le 30$ the regression analysis is based on 50 different
choices of $\Delta G_{i,j}(N)$ by choosing different subsets of
$N$ matches to improve the statistics. The result is shown in
Fig.2. The estimated error results from performing this analysis
individually for each season. Due to the strong correlations for
different $N$-values the final error is much larger than suggested
by the fluctuations among different data points. The data are
compatible with $c_3 = 0$. {Thus, we have shown that the simple
choice
\begin{equation}
\label{eqfinal}
q_{ij} = \Delta G_i - \Delta G_j + c_{home}
\end{equation}
is the uniquely defined relation (neglecting irrelevant terms of
5th order) to characterize the average outcome of a soccer match.}
In practice the right side can be estimated via Eq.\ref{eqgest}.
{This result implies that $h(t) = \langle (\Delta G_i -
\Delta G_j)(\Delta G_i - \Delta G_k) \rangle = \sigma_{\Delta G}^2
+ \langle \Delta G_j \Delta G_k \rangle$, i.e. $h(t \ne 17) =
\sigma_{\Delta G}^2$ and $h(t = 17) = 2\sigma_{\Delta G}^2$. This
agrees very well with the data.} Furthermore, the variance of the
$q_{ij}$ distribution, i.e. $\sigma_q^2$, is by definition given
by $2 \sigma_{\Delta G}^2 \approx 0.44$.
\begin{figure}[ht]
\includegraphics[width=0.95\linewidth]{fig2.eps}
\caption{Determination of $c_3$ by finite-size
scaling.}
\label{fig.2}
\end{figure}
{\it Determination of $\sigma_f^2$}: This above analysis does not contain any information about the
match-specific fitness relative to $\Delta G_i - \Delta G_j$. For example $f_{ij} > 0$
during a specific match implies that team $i$ plays better than
expected from $q_{ij}$. The conceptual problem
is to disentangle the possible influence of these fitness fluctuations
from the random aspects of a soccer match. The key idea is based on the observation
that, e.g., for $f_{ij} > 0$ team $i$ will play better than expected in both
the first and the second half of the match. In contrast, the random features of a match
do not show this correlation. For the identification of $\sigma_f^2$ one defines $A =
\langle ((\Delta g^{(1)}_{ij}/b_1-c_{home})\cdot ((\Delta g^{(2)}_{ij}/b_2-c_{home})\rangle_{ij}$
where $\Delta g^{(1),(2)}_{ij}$ is the goal
difference in the first and second half in the specific match,
respectively and $b_{1,2}$ the fraction of goals scored during the first
and the second half, respectively ($b_1 =
0.45; b_2 = 0.55$). Based on Eq.\ref{eqfinal} one has $\sigma_f^2 = A - 2\sigma_{\Delta G}^2$. Actually,
to improve the statistics we have additionally used different partitions of the match (e.g. first and third
quarter vs. second and fourth quarter).
Numerical evaluation yields $\sigma_f^2 = -0.04 \pm 0.06$ where the error bar is estimated
from individual averaging over the different seasons. Thus one obtains in particular
$\sigma_f^2 \ll \sigma^2_q$ which renders match-specific fitness fluctuations irrelevant.
Actually, as shown in \cite{HR09}, one can observe a tendency that teams which have lost 4 times
in a row tend to play worse in the near future than expected by their fitness. Strictly speaking
these strikes indeed reflect minor temporary fitness variations. However, the number of strikes
is very small (less than 10 per season) and, furthermore, mostly of statistical nature. The same holds
for red cards which naturally influence the fitness but fortunately are quite rate. Thus, these extreme
events are interesting in their own right but are not relevant for the overall statistical description.
The negative value of $\sigma_f^2$ points towards anti-correlations between both partitions of the match.
A possible reason is the observed tendency towards a draw, as outlined below.
\begin{figure}[ht]
\includegraphics[width=0.95\linewidth]{fig3.eps}
\caption{({\bf a})Distribution of goals per team and match and the Poisson prediction if the different fitness values
are taken into account (solid line). Furthermore a Poisson estimation is included where only the home-away asymmetry is included
(broken line). The quality of the predicted distribution is highlighted in ({\bf b}) where the ratio of the estimated and the actual probability is shown.}
\label{fig.3}
\end{figure}
{\it Determination of $r_{ij}$}: The actual number of
goals $g_{i,j}$ per team and match is shown in Fig.3. The error
bars are estimated based on binomial statistics. As discussed
before the distribution is significantly broader than a Poisson
distribution, even if separately taken for the home and away goals
\cite{Greenhough02,janke,janke09}. Here we show that this
distribution can be generated by assuming that scoring goals are
independent Poissonian processes. We proceed in two steps. First,
we use Eq.\ref{eqfinal} to estimate the average goal difference
for a specific match with fitness values estimated from the
remaining 33 matches of each team. Second, we supplement
Eq.\ref{eqfinal} by the corresponding estimator for the sum of the
goals $g_i + g_j$ given by $ \Sigma G_i + \Sigma G_j-
\lambda$. Together with Eq.\ref{eqfinal} this allows us to
calculate the expected number of goals for both teams individually.
Third, we generate for both teams a Poissonian
distribution based on the corresponding expectation values. The
resulting distribution is also shown in Fig.1 and perfectly agrees
with the actual data up to 8 (!) goals. In contrast, if the
distribution of fitness values is not taken into account
significant deviations are present. { Two conclusions can be drawn. First,
scoring goals is a highly random process. Second, the good
agreement again reflects the fact that $\sigma_f^2$ is small because
otherwise an additionally broadening of the actual data would be expected.
Thus there is no indication of a possible influence of
self-affirmative effects during a soccer match \cite{janke,janke09}}.
Because of the underlying Poissonian process the value of $\sigma_r^2$ is just given by the average number of
goals per match $(\approx 3)$.
\begin{figure}[ht]
\includegraphics[width=0.95\linewidth]{fig4n.eps}
\caption{({\bf a}) The probability distribution of the goal difference per match together
with its estimation based on independent Poisson processes of both teams. In
({\bf b}) it is shown for different scores how the ratio of the estimated and the
actual number of draws differ from unity.}
\label{fig.4}
\end{figure}
As already discussed in literature the number of draws is
somewhat larger than expected on the basis of independent Poisson distributions; see, e.g., Refs. \cite{Dixon97,Rue00}.
As an application of the present results
we quantify this statement. In Fig.4 we compare the calculated distribution of $\Delta g_{ij}$
with the actual values. The agreement is very good except for
$\Delta g_{ij} = -1,0,1$. Thus, the simple picture of independent goals of the home and the away
team is slightly invalidated. The larger number of draws is balanced by a
reduction of the number of matches with exactly one goal
difference. More specifically, we have calculated the relative
increase of draws for the different results. The main effect is due to the
strong increase of more than 20\% of the 0:0 draws. Note that the present analysis has
already taken into account the fitness distribution for the estimation of this number. Starting from 3:3 the
simple picture of independent home and away goals holds again.
The three major contributions to the final soccer result
display a clear hierarchy, i.e. $\sigma_r^2 : \sigma_q^2:
\sigma_f^2 \approx 10^2:10^1:10^0$.
$\sigma_f^2$, albeit well defined and quantifiable, can be
neglected for two reasons. First, it is small as compared to the
fitness variation among different teams. Second, the uncertainty
in the prediction of $q_{ij}$ is, even at the end of the season,
significantly larger (variance of the uncertainty: $2 \cdot
\sigma_{e,N=33}^2 = 0.12$, see above). {Thus, the limit of predictability of a soccer
match is, beyond the random effects, mainly related to the uncertainty in the
fitness determination rather than to match specific effects}. Thus, the hypothesis of a
strictly constant team fitness during a season, even on a
single-match level cannot be refuted even for a data set
comprising more than 20 years. In disagreement with this
observation soccer reports in media often stress that a team
played particularly good or bad. Our results suggest that there
exists a strong tendency to relate the assessment too much to the
final result thereby ignoring the large amount of random aspects
of a match.
In summary, apart from the minor correlations with
respect to the number of draws soccer is a surprisingly simple
match in statistical terms. Neglecting the minor differences
between a Poissonian and binomial distribution and the slight
tendency towards a draw a soccer match is equivalent to two teams
throwing a dice. The number 6 means goal and the number of
attempts of both teams is fixed already at the beginning of the
match, reflecting their respective fitness in that season.
More generally speaking, our approach may serve as a general
framework to classify different types of sports in a
three-dimensional parameter space, expressed by $\sigma_r^2,
\sigma_q^2, \sigma_f^2$. This set of numbers, e.g., determines the
degree of competitiveness \cite{ben2}. For example for matches
between just two persons (e.g. tennis) one would expect that
fitness fluctuations ($\sigma_f^2$) play a much a bigger role and
that for sports events with many goals or points (e.g.
basketball) the random effects ($\sigma_r^2$) are much less
pronounced, i.e. it is more likely that the stronger team indeed wins.
Hopefully, the present work stimulates activities to
characterize different types of sports along these lines.
We greatly acknowledge helpful discussions with B. Strauss, M. Trede, and M. Tolan about this topic.
|
\section{Introduction}
Looking back at the nineteenth century, we should not be surprised that the formulation of thermodynamics coincided with the industrial revolution.
Thermodynamics provides a deep and indispensable understanding of the principles by which engines and refrigerators operate, and the fundamental limits they must obey.
The development of thermodynamics in turn gave rise to statistical mechanics, as Boltzmann, Maxwell, Gibbs and others sought the link between concepts such as heat and temperature, and the Newtonian motion of atoms and molecules.
From these developments there emerged our modern understanding of thermodynamics as an effective theory for macroscopic systems, arising from a statistical treatment of their microscopic constituents.
The success of this approach rests the law of large numbers, which dictates
that the greater the number of constituents of the system, the smaller the relative size of deviations (fluctuations) from the average
behaviour. For macroscopic systems, significant deviations from the typical behaviour are virtually impossible, hence
computed averages coincide with measured values of the observables.
We are now in an era where the microscopic world is increasingly accessible, as evidenced by
dramatic progress in areas of research such as single-molecule manipulation and nanotechnology.
Advances in experimental techniques, theoretical modeling and numerical simulation have produced an increasingly detailed understanding of biomolecular machines such as myosin, kinesin, and the complex apparatus responsible for the replication, transcription and translation of the genetic code in living cells \cite{how01}.
Moreover, laboratories around the world are using the techniques of supramolecular chemistry to synthesize molecular complexes with moving parts \cite{kay07}.
For such microscopic systems thermal fluctuations are centrally important, and it is no longer sufficient to focus our attention exclusively on average behavior.
This, together with natural scientific curiosity, has led to increased interest in asking what the laws of thermodynamics ``look like'', when applied to microscopic systems \cite{bus05}.
In this text, we address this issue, focusing in particular on the behavior of systems driven away from an initial state of thermal equilibrium.
We will discuss several results that pertain to such systems, and that involve the relationship between {\it work} and {\it free energy}.
While these results are in principle valid quite generally, in practice they are relevant mostly to microscopic systems, for which fluctuations are substantial.
As this text is a pedagogical discussion, we will start with a brief review of the relevant macroscopic thermodynamic relations,
using the familiar example of a stretched rubber band.
We will then consider how these relations might apply to a microscopic analogue: a single RNA molecule stretched with optical tweezers.
Here the laws of thermodynamics, particularly the second law, must be interpreted statistically, in terms of averages over many measurements.
However, it is the fluctuations around these averages that will be the object of central interest in this text:
as we will describe in detail in Section \ref{sec4}, these fluctuations -- which we might be tempted to dismiss as mere ``noise'' -- in fact satisfy a number of strong and unexpected relations, which remain valid even far from thermal equilibrium.
This text was prepared in large part by two of the authors (E.B. and B.W.), based on lectures given by the third author (C.J.) at the International Summer School on Fundamental Problems in Statistical Physics XII (FPSPXII), held in Leuven, Belgium in September, 2009.
It is meant to provide an introduction to the topics it covers, in a style that reflects the informal and stimulating atmosphere of the summer school.
\section{Brief review of thermodynamic processes}
\subsection{An experiment with a rubber band}
\begin{figure}[ht]
\begin{center}
\includegraphics[width = 7cm]{rubber.eps}
\caption{Stretching a rubber band}
\label{rubber}
\end{center}
\end{figure}
Consider the system shown in Figure \ref{rubber}.
One end of a rubber band is attached to an immovable wall, the other end to a spring on which a force can be exerted. The length of the rubber band will be denoted by $z$ and the combined length of the rubber band and spring will be denoted by $\lambda$.
The system interacts with the surrounding air (thermal environment) which is at temperature $T$.
The parameter $\lambda$ will be controlled externally and will be used to exert work on the system. We assume that for every fixed value of
$\lambda$ and $T$ there is a unique equilibrium state, to which the system relaxes if undisturbed.
We will use the term {\it thermodynamic process} to denote a sequence of events during which a system evolves from one equilibrium state to another.
During such a process, work ($W$) is performed on the system,
heat ($Q$) flows into the system from the environment, and
the internal energy of the system may change ($\Delta U$).
The first law then reads
$$
\Delta U = W + Q
$$
There is some subtlety even in this seemingly straightforward statement of energy balance.
Namely, the definitions of $W$, $Q$ and $\Delta U$ depend on how we define our system of interest.
For example, in Figure \ref{rubber} we could take the system of interest to be either the rubber band itself, or the rubber band together with the spring.
The corresponding definitions of work are then, respectively,
\begin{itemize}
\item system = rubber band alone : $W = \int F_{\textrm{spring}} dz$
\item system = rubber band + spring : $W = \int F_{\textrm{spring}} d\lambda$
\end{itemize}
where $F_{\textrm{spring}}$ is the instantaneous tension of the spring.
The difference between these two values is simply the net change in the energy of the spring itself.
Either choice is valid, but in this text we will use the second definition.
The second law of thermodynamics asserts that there exists a state function, {\it entropy} ($S$), that obeys the Clausius inequality,
$$
\int_{A}^{B} \frac{{\mathchar'26\mkern-12mu {\rm d}} Q}{T} \leq \Delta S = S_B - S_A
$$
for any process that begins with the system in equilibrium state $A$ and ends in equilibrium state $B$.
Here the inexact differential $\,{\mathchar'26\mkern-12mu {\rm d}} Q$ denotes an amount of heat absorbed from the thermal surroundings, at temperature $T$.
The equality is attained if and only if the process is reversible.
Let us now imagine the following process: initially the rubber band is at equilibrium
with the surroundings at temperature $T$ and the initial value $\lambda=A$ of the work parameter. Next we quickly stretch
the band by changing $\lambda$ to a new value $B$. During this process the rubber band will heat up a bit, so once the stretching is complete we let it cool down
until it is again at equilibrium with the surroundings. For such a process, by combining the first and the second law we find :
\begin{eqnarray*}
W & = & \Delta U - Q \\
& = & \Delta U - T \int_{A}^{B} \frac{{\mathchar'26\mkern-12mu {\rm d}} Q}{T} \\
& \geq & \Delta U - T \Delta S
\end{eqnarray*}
Note that $T$ here denotes the (constant) temperature of the bath, rather than that of the system itself.
In terms of the Helmholtz free energy, $F = U - TS$, we have
\begin{equation}
\label{seclaw}
W \geq \Delta F
\end{equation}
This inequality is the general statement of the second law for processes in which the system of interest is in contact with (at most) a single heat bath.
Since we will never consider the situation of two or more heat baths in this text, equation (\ref{seclaw}) is the form of the second law that we will use.
\subsection{Cyclic processes}
We will occasionally consider {\it cyclic} processes, that is, processes during which the value of the external parameter $\lambda$ is the same at the beginning as at the end.
For the rubber band example, one can imagine that after the stretching procedure described above, one brings the band back to its initial state, by varying $\lambda$ from
$B$ to $A$ and again allowing the system to equilibrate with its surroundings.
For cyclic processes we have $\Delta F_{\textit{cyc}} = 0$ (since free energy is a state function)
thus equation (\ref{seclaw}) becomes
$$
W_{\textit{cyc}} \geq 0
$$
This is Thomson's formulation of the second law. \cite{tho1851}
Since the net change in the internal energy of the system is also zero during a cyclic process ($\Delta U_{\textit{cyc}}=0$) we must also have $Q_{\textit{cyc}}
\leq 0$.
The opposite case ($W_{\textit{cyc}}<0<Q_{\textit{cyc}}$), illustrated schematically in Figure \ref{free_lunch}, is not possible for a cyclic process.
If it were, we could construct a so-called perpetual motion machine of the second kind, a device that would convert thermal fluctuations into work.
Thus Thomson's formulation is essentially a ``no free lunch'' statement: we cannot (for example) harvest the energy contained in the thermal motions of air molecules, and deliver that energy as work, simply by repeatedly stretching and contracting a rubber band.
(We stress that this conclusion is valid in the presence of a single thermal reservoir.
When two or more heat baths are present, we {\it can} obtain work simply by stretching and contracting of a rubber band, as illustrated beautifully in Ref. \cite{fey66}.)
\begin{figure}[ht]
\begin{center}
\includegraphics[width = 7cm]{free_lunch.eps}
\caption{The second law prohibits this situation, for cyclic processes.}
\label{free_lunch}
\end{center}
\end{figure}
Now let us split the cyclic process that takes $\lambda$ from $A$ to $B$ and back to $A$ into
a {\it forward} and a {\it reverse} process.
\begin{itemize}
\item Forward process : $\lambda$ : A $\rightarrow$ B
\item Reverse process : $\lambda$ : B $\rightarrow$ A
\end{itemize}
More precisely, let us assume the reverse process uses the time-reversed protocol for varying the work parameter: $\lambda_{t}^{R} = \lambda_{\tau-t}^{F}$,
where $\tau$ is the duration of either process.
If we now apply (\ref{seclaw}) to both the forward and the reverse process we find
\begin{eqnarray*}
W_F \geq \Delta F & = & F_B - F_A \\
W_R \geq - \Delta F & = & F_A - F_B ,
\end{eqnarray*}
equivalently
\begin{equation}
\label{eq:ineq_chain}
-W_R \leq \Delta F \leq W_F
\end{equation}
Here too, the equalities are attained if and only if the process is reversible.
Equation (\ref{eq:ineq_chain}) immediately implies
\begin{equation}
W_F + W_R \geq 0
\end{equation}
which is simply Thomson's formulation of the second law.
\section{How might these results apply to microscopic systems?} \label{sec3}
\subsection{Example: stretching a strand of RNA}
\label{sec:RNA}
To illustrate how the thermodynamic results of the last section translate to
the microscopic world, it is again useful to consider a specific example.
Figure \ref{molecule} schematically depicts an experimental setup for stretching a single molecule, a microscopic analogue of the rubber band discussed in the previous section.
The system of interest here is a strand of RNA, which is chemically attached to a polystyrene bead, held fixed by a micropipette (analogous to the wall in
the rubber band example). The other end of the RNA is also attached to a bead, which is confined by
a laser trap (analogous to the spring in the previous example). The whole system is surrounded by
water at room temperature, providing the thermal environment. By varying the position of the laser
trap ($\lambda$) work is performed on the system.
For actual experimental implementations of such single-molecule manipulation experiments, within the context of the theoretical predictions of Section \ref{sec4}, see Liphardt et al. \cite{lip02}, Collin et al \cite{col05}, or (using atomic force microscopy instead of a laser trap) Harris and Song \cite{har07}.
\begin{figure}[ht]
\includegraphics[width = 10cm]{picture1.eps}
\caption{Stretching a single molecule.
Note that this figure is not to scale: in reality the beads and DNA handles are much larger than the strand of RNA.}
\label{molecule}
\end{figure}
Now imagine the following irreversible process: start at $\lambda=A$, with the molecule in equilibrium with the thermal environment, then quickly stretch the molecule by varying $\lambda$ to a new value $B$, then let the system relax to a new equilibrium state, at $\lambda=B$.
Once this sequence of events is complete, change $\lambda$ back to
$A$, following the reverse protocol, as in the previous section, and let the system relax again to equilibrium.
As above, let $W_F$ and $W_R$ denote the work performed during the forward (stretching) and reverse (contraction) stages, and $W_{\textit{cyc}} = W_F + W_R$.
If we repeat this cyclic process many times, we will not be surprised to measure different values of work, from one repetition to the next, as thermal fluctuations of the RNA strand and the surrounding water molecules play a significant role at this microscopic scale. We might even encounter rare occasions for which $W_{\textit{cyc}}<0$; these are in effect fortuitous events, during which the thermal motions
of the water molecules happen to facilitate the stretching and contraction of the RNA strand.
Because fluctuations are not negligible at this microscopic scale, we should interpret the second law as a statistical law, pertaining to expectation values, rather than a prediction about any particular realization of an experiment.
Thus equation (\ref{eq:ineq_chain}) is replaced by the following statement about average work values:
\begin{equation}
\label{eq:avgineqchain}
\left<-W_R\right> \leq \Delta F \leq \left<W_F\right>
\end{equation}
Here and throughout this text, angular brackets denote an average over many realizations (repetitions) of the process.
For the full cycle we have
\[ \left<W_{cyc}\right> = \left<W_F+W_R\right> \geq 0 \]
These considerations are illustrated in Figure \ref{workdis}, where
work distributions $\rho_F(W)$ and $\rho_R(-W)$ for the forward and reverse processes are depicted.
We see that the average work values satisfy equation (\ref{eq:avgineqchain}), but each distribution shows a substantial spread of possible work values.
Note that the distributions have been drawn to intersect precisely at $W=\Delta F$; this is not a coincidence, but rather the consequence of equation (\ref{res2}) discussed in Section \ref{sec4} below.
\begin{figure}[ht]
\includegraphics[width = 10cm]{picture2.eps}
\caption{Work distributions of the forward and reverse processes.}
\label{workdis}
\end{figure}
\subsection{The second law from microscopic arguments}
The discussion in Section \ref{sec:RNA} can be summarized as follows: while the Clausius inequality might be violated during individual realizations of the process, $W < \Delta F$, it will be satisfied {\it on average}, $\left< W \right> \ge \Delta F$.
This conclusion is an intuitively natural way to extend the second law to microscopic systems, where fluctuations are non-negligible.
It is interesting to ask whether we can actually justify this intuitive conclusion using microscopic principles.
This is a deep question that has led to both insight and controversy, and we will not attempt to frame a satisfactory answer here.
Nevertheless, it is instructive to go through some simple analysis of an isolated Hamiltonian system, evolving under a cyclic process.
(In the discussion below, it is useful to keep in mind that this ``isolated system'' might be a proxy for the combination of a system of interest and a heat bath, if we consider the two together as a large, isolated system.)
Let the microstate
$x = (q,p) = (\vec{q}_1,\ldots,\vec{q}_m,\vec{p}_1 \ldots,\vec{p}_m )$ describe the positions and
momenta of all the particles in the system. The $6m$-dimensional space that $x$ belongs to is the
phase space $\Gamma$. The Hamiltonian $H_{\lambda}(x)$ is a function of this microstate and one or more parameters $\lambda$,
and governs the dynamics by Hamilton equations:
\[ \dot{q} = \frac{\partial H}{\partial p}, \ \ \ \ \ \ \ \dot{p} = -\frac{\partial H}{\partial q} \]
This describes a fully deterministic dynamics, i.e. when $x(0)$ is given, $x(t)$ is known for any later
time $t$. This $x(t)$ (or $x_t$) defines what we call a trajectory or path through the phase space.
Hamiltonian dynamics satisfy Liouville's theorem: take a subset $C$ of the phase space, with volume $|C|$.
If we let all the points $x\in C$ evolve under Hamiltonian dynamics to a later time $t$, this defines
a new subset $C_t = \{ x(t)| x(0)\in C \}$ with volume $|C_t|$. Liouville's theorem then states that $|C_t| = |C|$,
or equivalently:
\[ \left| \frac{\partial x_t}{\partial x_0}\right| \equiv \det\left( \frac{\partial x_i(t)}{\partial x_j(0)} \right) = 1 \]
For an isolated system, the first law of thermodynamics dictates that
\[ W = \Delta U = H_B(x_t) - H_A(x_0) \]
where again $A$ and $B$ denote the values of the external parameter $\lambda$ at the beginning and the end of the evolution.
For a cyclic evolution of the parameter, $A=B$.
We can now pose the question:
does the inequality $W_{\textit cyc} \ge 0$ hold for every initial condition $x(0)$?
In other words, for this isolated system is Thomson's law satisfied for every trajectory?
Without attempting a rigorous analysis, the following simple argument suggests the answer is ``no''.
\begin{figure}[ht]
\includegraphics[width = 10cm]{picture3.eps}
\caption{Hypothetical phase diagram illustrating the scenario $C \subset C_\tau$.}
\label{liouville}
\end{figure}
For simplicity, imagine a Hamiltonian $H_A(x)$ with a single minimum in phase space, and consider the set of initial conditions $C$ corresponding to all microstates with an energy lower than
some value $U_0$, see Figure \ref{liouville} (on the left).
Let $C_\tau$ denote the set of final conditions reached after Hamiltonian evolution during a cyclic process.
There are three scenarios for the relationship between these two subsets of phase space: $C \not\subseteq C_\tau$, $C \subset C_\tau$, and $C = C_\tau$.
The first implies that there exists some point $\bar x$ that belongs to $C$ but not to $C_\tau$.
If we now consider a trajectory that ends at this point, then by construction the initial and final conditions for this trajectory satisfy:
\begin{equation}
x_0 \notin C \quad,\quad x_\tau = \bar x \in C
\end{equation}
Hence $W_{\textit cyc} < 0$ for this trajectory.
The second scenario ($C \subset C_\tau$) is depicted on the right side of figure \ref{liouville}, and implies $|C_\tau|\geq|C|$, a clear contradiction with Liouville's theorem; therefore we can rule out this scenario.
Finally, the third scenario, namely $C = C_\tau$, corresponds to the limiting case of a reversible process, for which we expect $W_{\textit cyc}=0$.
Thus we conclude that (unless the process is performed reversibly) there will exist initial conditions for which $W_{cyc} < 0$.
The above considerations suggest that the second law ought to be interpreted statistically:
\[ \left<W_{cyc}\right> = \int dx_0 f_A(x_0)\left[ H_A(x_t(x_0)) - H_A(x_0) \right] \geq 0 \]
where $f_A(x_0)$ is a normalized probability distribution that represents the statistical state of equilibrium at $\lambda=A$.
Whether or not this inequality is satisfied depends, of course, on the choice of distribution $f_A$.
In the following paragraph we show below that the inequality is satisfied for a canonical distribution of initial conditions, $f_A(x) \propto e^{-\beta H_A(x)}$.
It is also straightforward to establish that the inequality is satisfied if $f_A(x) \propto \theta(U_0-H_A(x))$,
where $U_0$ is a constant and $\theta(\cdot)$ is the unit step function.
(We leave this as an exercise for the reader.)
Somewhat surprisingly, however, for a microcanonical distribution of initial conditions,
$f_A(x) \propto \delta(U_0-H_A(x))$, it is possible to construct explicit examples for which $\langle W_{cyc}\rangle < 0$ \cite{mar09,sat02}.
Consider a map $M: \Gamma \to \Gamma: x \mapsto y$, for which $\left|{\partial y}/{\partial x}\right| = 1$
and for which the inverse $M^{-1}$ exists.
An example of such a map is Hamiltonian evolution for some fixed time ($x = x_0$, $y = x_\tau$).
Define furthermore a positive constant $\beta$ and the work $W(x) := H_A(y(x))-H_A(x)$.
We compute the following average:
\begin{eqnarray}
\left<e^{-\beta W}\right> &=& \int dx \frac{1}{Z_A}e^{-\beta H_A(x)}e^{-\beta W(x)}\nonumber\\
&=& \frac{1}{Z_A}\int dx \,e^{-\beta H_A(y(x))}\nonumber\\
&=& \frac{1}{Z_A}\int dy \left|\frac{\partial x}{\partial y}\right| e^{-\beta H_A(y)}\nonumber\\
&=& 1 \label{thomson}
\end{eqnarray}
where $Z_A$ is the usual partition function.
From Jensen's inequality (see Appendix A for a brief derivation) it follows that
\[ 1 = \left<e^{-\beta W}\right> \geq e^{-\beta\left<W\right>} \]
so that indeed $\left<W\right>\geq0$.
A more general version of this derivation can be found in e.g. \cite{jar02,obe05}.
These very simple considerations should not be viewed as first-principles derivations of the second law.
Rather, they provide some quantitative justification for the intuition stated at the beginning of this section: the second law (which in this case takes the form $W_{cyc}\ge 0$) will be violated for individual trajectories, but will be satisfied for an average over an ensemble of trajectories, at least for certain choices for the distribution of initial conditions.
\section{Nonequilibrium work and fluctuation relations}\label{sec4}
The central point of Section \ref{sec3} above is that -- for microscopic systems -- the second law ought to be interpreted statistically, as a statement about averages.
We now turn to the main focus of this contribution, namely the {\it fluctuations} around these averages.
We will discuss three prediction, given by equations (\ref{res1}), (\ref{res2}) and (\ref{res3}) below, that pertain to these fluctuations.
After introducing these results, we will derive each of them, using three different schemes for modeling the microscopic dynamics of a system driven away from equilibrium.
In each case, we imagine (as with the single-molecule stretching in Section \ref{sec:RNA} above) that the system of interest is prepared in equilibrium with a heat bath, then driven away from equilibrium by varying the work parameter over a time interval $0<t<\tau$, from $\lambda_0=A$ to $\lambda_\tau=B$.
\begin{enumerate}
\item The first prediction is the {\it nonequilibrium work relation} \cite{jar97a},
\begin{equation}\label{res1}
\left< e^{-\beta W} \right> = e^{-\beta \Delta F}
\end{equation}
As before, $W$ is the work performed during a single realization, $\Delta F = F_B - F_A$ is the free energy difference between the initial and final equilibrium states, $\beta = (k_BT)^{-1}$ is the inverse temperature of the heat bath, and angular brackets denote an average over an ensemble of realizations of the process.
In Section \ref{sec:hamiltonian} below, we will derive equation (\ref{res1}) under Hamiltonian dynamics, and we will show that this equality immediately implies two inequalities related to the second law:
\begin{itemize}
\item $\left< W \right> \geq \Delta F$
\item ${\rm Prob}(W \leq \Delta F - n\, k_BT) \leq e^{-n} $
\end{itemize}
for any real, dimensionless $n > 0$.
\item The second prediction, a {\it fluctuation theorem} due to Crooks \cite{cro98,cro99},
\begin{equation}\label{res2}
\frac{\rho_F(W)}{\rho_R(-W)} = e^{\beta (W - \Delta F)}
\end{equation}
will be proven in Section \ref{sec:stochastic} for stochastic jump processes.
Note that equation (\ref{res1}) follows directly from equation (\ref{res2}).
Moreover, this result implies that the
work distributions for the forward and reverse processes cross one another at $W=\Delta F$, as depicted in Figure \ref{workdis}.
\item
Equations (\ref{res1}) and (\ref{res2}) in principle allow us to determine an equilibrium free energy difference from measurements of nonequilibrium work values.
The third prediction goes further, giving a prescription for constructing the entire equilibrium distribution \cite{hum01,hum05}:
\begin{equation}\label{res3}
\left< \delta(x-x_t) e^{-\beta W_t} \right> = \frac{1}{Z_A} e^{-\beta H(x, \lambda_t)}
\end{equation}
where $W_t$ is the work done up to time t for a trajectory $x_t$ and $Z_A$ is the partition function at time $t=0$.
(Equivalent formulations of this result appear in Refs. \cite{jar97b,cro00}.)
We will derive this result for the case of continuous time Markov processes.
\end{enumerate}
Before proceeding to the derivations of these results, some general remarks are in order.
First, these results remain valid even if the system is driven far from equilibrium during the process.
They place strong and rather unexpected constraints on the work fluctuations: one would not have naturally guessed the validity of equations (\ref{res1}) - (\ref{res3}) simply by extrapolating from macroscopic, thermodynamic intuition.
Next, these results reveal that information about equilibrium properties (such as $\Delta F$) is in effect encoded in the fluctuations of the system when driven out of equilibrium.
Finally, equations (\ref{res1}) - (\ref{res3}) are just three results among a considerably larger set of predictions related to fluctuations in far-from-equilibrium processes.
In particular, the early work of Bochkov and Kuzovlev \cite{boc77,boc81a,boc81b} includes related predictions, using a somewhat different definition of work.
Furthermore, while we consider only fluctuations in {\it work}, there has been extensive investigation of fluctuations in {\it entropy production}, which satisfy similar relations; see e.g.\ Refs. \cite{eva93,eva94,gal95,kur98,leb99,mae99,mae03,har07b,sev08}.
\subsection{Hamiltonian dynamics}
\label{sec:hamiltonian}
Here we derive equation (\ref{res1}), using Hamiltonian dynamics to model the microscopic evolution of the system.
Specifically, we imagine that after the system has been prepared in equilibrium, at $\lambda=A$, the heat bath is removed and the now-isolated system evolves under Hamilton's equations as the work parameter is varied from $A$ to $B$.
At the end of the process the system is again brought into thermal contact with the heat bath and allowed to reach equilibrium.
This scenario does not realistically describe many physical situations, such as the RNA stretching experiment, therefore the derivation of equation (\ref{res1}) that we present here is not the most general one.
However, it is a particularly simple derivation, which avoids some of the technical complications that arise when treating a system that remains in contact with a heat bath during the entire process;
see Ref.~\cite{jar04} for a Hamiltonian analysis that explicitly includes the degrees of freedom of the system and the heat bath.
Our system is described by a Hamiltonian $H(x,\lambda)$, where $x = (\textbf{q}\: , \textbf{p})$.
Because the system is isolated, the work performed as the external parameter is varied from $\lambda_0=A$ to $\lambda_\tau=B$ is
\begin{equation*}
W = H(x_{\tau}, B) - H(x_0, A)
\end{equation*}
Since Hamiltonian dynamics are deterministic, we can view the final conditions as a function of the initial conditions, thus writing
\begin{equation}
W = H(x_{\tau}, B) - H(x_0, A) = H(x_{\tau}(x_0), B) - H(x_0, A) = W(x_0)
\end{equation}
The only stochasticity is in the assignment of initial conditions for a given realization, therefore to compute the left side of equation (\ref{res1}) we must average over the distribution of these conditions:
\[ \left< e^{-\beta W} \right> = \int dx_0 f^{\textit{eq}}(x_0, A)e^{-\beta W(x_0)} \]
Here $f^{\textit{eq}}$ is the equilibrium distribution, given by the canonical expression
\begin{equation}
f^{\textit{eq}}(x, \lambda) = \frac{1}{Z_{\lambda}} e^{-\beta H (x, \lambda)}
\end{equation}
We now perform the averaging as in the previous section:
\begin{eqnarray*}
\left< e^{-\beta W} \right> & = & \int dx_0 \frac{1}{Z_A} e^{- \beta H(x_0, A)} e^{- \beta W(x_0)}\\
& = & \frac{1}{Z_A} \int dx_0 e^{-\beta H(x_{\tau}(x_0), B)} \\
& = & \frac{Z_B}{Z_A} = e^{- \beta \Delta F}
\end{eqnarray*}
where the free energy is given by the familiar expression,
\begin{equation}
F(\lambda) = - \frac{1}{\beta} \ln Z_{\lambda}\ \ \ \ \ \ \ \ \Delta F = F(B) - F(A)
\end{equation}
One consequence of (\ref{res1}) is that the second law (in the form of the Clausius inequality) appears as an application of Jensen's inequality:
since the exponential function is convex, we have that $\left< e^{-\beta W} \right> \geq e^{-\beta \left<W\right>}$,
and therefore $\left< W\right> \geq \Delta F$.
We can also use equation (\ref{res1}) to establish a bound on the probability to observe work values that ``violate'' the second law, $W < \Delta F$, as follows.
If $\rho(W)$ represents the probability distribution of work values, then the probability that the work is smaller than $\Delta F - nk_BT$, where $n>0$, is
\begin{eqnarray*}
{\rm Prob}(W\leq \Delta F - nk_BT) &=& \int_{-\infty}^{\Delta F - nk_BT} dW \rho(W)\\
&\leq& \int_{-\infty}^{\Delta F - nk_BT} dW \rho(W)e^{-\beta(W-\Delta F + nk_BT)}\\
&\leq& e^{\beta(\Delta F - nk_BT)}\int_{-\infty}^{+\infty}dW\rho(W)e^{-\beta W}\\
&=& e^{-n}
\end{eqnarray*}
In other words the probability to observe work values smaller than $\Delta F$ decays exponentially, or faster, in the size of the violation.
This is consistent with our experience that macroscopic violations of the second law are never observed.
\subsection{Stochastic jump processes}
\label{sec:stochastic}
Here we derive equation (\ref{res2}) for a discrete-time Markov dynamics.
Instead of representing the evolution of the system as a continuous trajectory, we imagine a chronologically ordered sequence of configurations,
\begin{eqnarray*}
x_0 \rightarrow x_1 \rightarrow x_2 \rightarrow \cdots \rightarrow x_{\tau - 1} \rightarrow x_{\tau}
\end{eqnarray*}
which might represent ``snapshots'' of the system at equally spaced intervals of time.
We imagine that the system remains in contact with the heat bath throughout this process, therefore the transitions from one state to the next are stochastic.
These dynamics are specified by the transition probability densities $P_{\lambda} (x \rightarrow x')$, which depend on time through the parameter $\lambda$, and which satisfy the normalization condition
\begin{equation}
\int dx' P_{\lambda} (x \rightarrow x') = 1
\end{equation}
We assume that the process is Markovian, in other words each transition is statistically independent of the previous transitions.
We also assume that each transition satisfies a detailed balance condition
\begin{equation}\label{detbal}
\frac{P_{\lambda} (x \rightarrow x')}{P_{\lambda} (x \leftarrow x')} = e^{-\beta [H_{\lambda}(x') - H_{\lambda}(x)]}
\end{equation}
where the direction of the arrow indicates the transition.
We can interpret this condition by imagining for a moment that $\lambda$ is held fixed, allowing the system to relax to a state of equilibrium represented statistically by the Boltzmann distribution.
Equation (\ref{detbal}) then tells us that in this state the net rate of transitions from $x$ to $x'$ is the same as that from $x'$ to $x$:
\begin{equation}
\frac{1}{Z_{\lambda}} e^{-\beta H_{\lambda}(x)} P_{\lambda} (x \rightarrow x') = \frac{1}{Z_{\lambda}} e^{-\beta H_{\lambda}(x')} P_{\lambda} (x \leftarrow x')
\end{equation}
Of course, we are interested in a process during which the value of $\lambda$ is not held fixed, and therefore the system does not (in general) remain in equilibrium.
To describe this situation, let us imagine that $\lambda$ changes in between each transition.
We represent this situation schematically as follows:
\begin{eqnarray*}
\lambda_0 = A & x_0 \xrightarrow{\lambda_1} x_1 \xrightarrow{\lambda_2} x_2 \xrightarrow{\lambda_3} \cdots \xrightarrow{\lambda_{\tau -1}} x_{\tau -1}
\xrightarrow{\lambda_{\tau}} & \lambda_{\tau} = B
\end{eqnarray*}
Following Crooks \cite{cro98}, now establish a link with thermophysics by specifying what we mean by heat and work in this model.
We define heat as
\begin{equation}\label{heat}
Q = \sum\limits_{t=1}^{\tau} \left[ H(x_t, \lambda_t) - H(x_{t-1}, \lambda_t) \right]
\end{equation}
which is the sum of energy changes due to Markov transitions, while the work is defined as
\begin{equation}
W = \sum\limits_{t=1}^{\tau} \left[ H(x_{t-1}, \lambda_t) - H(x_{t-1}, \lambda_{t-1}) \right]
\end{equation}
which is the sum of energy changes due to the changing of $\lambda$.
Taking the sum of these two contributions, we arrive at a statement of the first law of thermodynamics:
\begin{equation}
Q + W = H(x_{\tau}, B) - H(x_0, A)
\end{equation}
Consider now a forward protocol and its reversed protocol,
\begin{eqnarray*}
A = \lambda_0^F \rightarrow \lambda_1^F \rightarrow \cdots \rightarrow \lambda_{\tau}^F = B \\
A = \lambda_{\tau}^R \leftarrow \lambda_{\tau-1}^R \leftarrow \cdots \leftarrow \lambda_{0}^R = B
\end{eqnarray*}
related by $\lambda_{t}^{F} = \lambda_{\tau - t}^{R}$. Likewise we define a forward and backward trajectory,
$X = (x_0 \rightarrow x_1 \rightarrow \cdots \rightarrow x_{\tau})$
and
$X^{\dagger} = (x_0 \leftarrow x_1 \leftarrow \cdots \leftarrow x_{\tau})$.
If $P^F(X)$ denotes the probability density for generating the trajectory $X$ during the forward process, and $P^R(X^\dagger)$ is defined analogously, then their ratio is
\begin{equation}
\frac{P^{F}(X)}{P^R(X^{\dagger})} = \frac{f_{A}^{\textit{eq}}(x_0) P_{\lambda_1^F} (x_0 \rightarrow x_1) \cdots P_{\lambda^F_{\tau}} (x_{\tau -1} \rightarrow x_{\tau})}{f_{B}^{\textit{eq}}(x_{\tau}) P_{\lambda^R_0} (x_{\tau} \rightarrow x_{\tau-1}) \cdots P_{\lambda^R_{\tau-1}} (x_{1} \rightarrow x_{0})}
\end{equation}
with initial conditions for either process sampled from the corresponding Boltzmann distribution.
Combining the assumption of detailed balance (\ref{detbal}) with the definition of heat (\ref{heat}), we find that
\[ \frac{P^{F}(X)}{P^R(X^{\dagger})} = \frac{f_{A}^{\textit{eq}}(x_0)}{f_{B}^{\textit{eq}}(x_{\tau})}e^{-\beta Q^F}\]
where $Q^F$ is the heat dissipated during the forward path ($Q^F = -Q^R$). Using the explicit expressions for the canonical distributions,
we arrive at \cite{cro98}
\begin{equation}\label{tussenres} \frac{P^{F}(X)}{P^R(X^{\dagger})} = e^{-\beta \Delta F + \beta \Delta U - \beta Q^F} = e^{\beta W^F-\beta \Delta F} \end{equation}
where $W^F = W^F (X)$ is the work done during the forward process.
With this expression, deriving equation (\ref{res2}) is simple:
\begin{eqnarray*}
\rho_F (W) & = & \int d X P^{F} (X) \delta (W - W^F (X)) \\
& = & \int d X P^R (X^{\dagger}) e^{\beta (W^F (X) - \Delta F)} \delta (W - W^F (X)) \\
& = & e^{\beta (W - \Delta F)} \int d X^{\dagger} P^{R}(X^{\dagger}) \delta (W + W^R (X^{\dagger}))\\
& = & e^{\beta (W - \Delta F)} \rho_R (-W)
\end{eqnarray*}
where $dX = dx_0 dx_1 \cdots dx_{\tau} = dX^{\dagger}$, and we have used $W^F(X) = -W^R(X^{\dagger})$.
From (\ref{res2}), we see that $\rho_F(\Delta F) = \rho_R(-\Delta F)$, meaning that observing a work value equal to the free energy differences is equally (un)likely during either process.
Furthermore, rewriting equation (\ref{res2}) as
\[ \rho_F (W)e^{-\beta W} = \rho_R(-W)e^{\beta \Delta F} \]
then integrating over $W$, we recover equation (\ref{res1}).
\subsection{Continuous time Markov dynamics}
\label{sec:feynmankac}
Equation (\ref{res3}) is essentially an application of the Feynmann-Kac theorem, as emphasized by Hummer and Szabo \cite{hum01,hum05}.
Let us therefore first sketch a (non-rigorous) derivation of this theorem in the context of continuous time Markov processes.
For a more mathematically rigorous approach, see Ge and Jiang~\cite{ge09}.
\paragraph*{Preliminaries:}
We define $P\left[X|x_0,t_0\right]$ to be the probability density
to observe the trajectory $X$, given that the system was in state $x_0$ at time $t_0$.
When we integrate this over all possible trajectories that end in state $x$ at time $t$,
we get the probability density of the system being in state $x$ at time $t$ given that it was
in state $x_0$ at time $t_0$:
\[ K(x,t|x_0,t_0) = \int dX P\left[X|x_0,t_0\right] \]
and $K(x,t_0|x_0,t_0) = \delta(x-x_0)$.
(The notation $\int dX$ here denotes integration over all intermediate states of the trajectory, with the initial and final points fixed.)
Note that $\int dx K(x,t_0|x_0,t_0)$ should be equal to one.
As we are working with a Markov process, one can use the Chapman-Kolmogorov
relations to write a differential equation for $K$:
\[
\frac{\partial}{\partial t}K(x,t|x_0,t_0) = \int dx' R_t(x'\to x)K(x',t|x_0,t_0)
\]
where $R_t$ is an instantaneous transition rate, given by
\[ R_t(x'\to x) = \lim_{\Delta t\to 0}\frac{K(x,t+\Delta t|x',t) - \delta(x-x')}{\Delta t} \]
and as a consequence $\int dx R_t(x'\to x) = 0$.
We schematically write the differential equation as
\begin{equation}\label{diffK}
\frac{\partial}{\partial t}K = L_t K
\end{equation}
where the operator $L_t$ is called the generator of the Markov process.
\paragraph*{Feynman-Kac:}
Now let us consider giving each trajectory a time-dependent statistical weight.
Specifically, we define an arbitrary function $\omega(x,t)$,
and the trajectory-dependent quantity
\begin{equation}
\label{eq:Omega}
\Omega(X) := \int_{t_0}^{t}\omega(x_s,s)ds
\end{equation}
With this, we define a quantity analogous to $K$:
\[ G(x,t|x_0,t_0) := \int dX P\left[X|x_0,t_0\right]e^{\Omega(X)} \]
where the difference lies in the exponential weight assigned to each path.
The goal now is to write down a differential equation for $G$ analogous to (\ref{diffK}).
For a short time interval $dt$ we have
\begin{eqnarray*}
G(x,t+dt|x',t) &=& \int dX P\left[X|x',t\right]e^{\int_{t}^{t+dt}ds\omega(x_s,s)}\\
&=& e^{\omega(x_t,t)dt}K(x,t+dt|x',t) + o(dt)\\
&=& e^{\omega(x_t,t)dt}[\delta(x-x') +R_t(x'\to x)dt] +o(dt)\\
\end{eqnarray*}
so that
\begin{eqnarray*}
G(x,t+dt|x_0,t_0) &=& \int dx' G(x,t+dt|x',t)G(x',t|x_0,t_0)\\
&=& (1+\omega(x,t)dt + dtL_t)G(x,t|x_0,t_0)+o(dt)
\end{eqnarray*}
This then finally leads us to:
\begin{equation}\label{diffG}
\frac{\partial G}{\partial t} = (\omega+L_t)G
\end{equation}
This result is the Feynman-Kac theorem in the present context:
when exponential weights given by equation (\ref{eq:Omega}) are assigned to the trajectories of a process with transition probabilities $K$ satisfying (\ref{diffK}), then the weighted transition probabilities $G$ satisfy (\ref{diffG}).
\paragraph*{Proof of equation (\ref{res3}):}
First note that the probability that the process
is in state $x$ at time $t$ can be written as
\[ f(x,t) = \left< \delta(x-x_t) \right> \]
As this is in our physical systems equal to $\int dx_0 f^{eq}_A(x_0)K(x,t|x_0,t_0)$ for any $t_0<t$, we see that $f$ itself satisfies
\[ \frac{\partial f}{\partial t} = L_t f \]
Recalling that the time-dependence comes from the external work parameter,
we write $L_t = L_{\lambda_t}$.
We further make the assumption of detailed balance, as above:
\[ L_{\lambda}e^{-\beta H(x,\lambda)}=0 \]
for every $\lambda$.
The work done during a particular process, up to time $t$, is equal to $w_t(X) = \int_0^t ds \dot{\lambda}\,{\partial H}/{\partial \lambda}$.
Here this work will enter into an exponential weight:
\begin{eqnarray*}
g(x,t) &:=& \left< \delta(x-x_t)e^{-\beta w_t} \right>\\
&=& \int dx_0 f^{eq}_A(x_0)\int dX P[X|x_0,0]e^{-\beta w_t(X)}\\
&=& \int dx_0 f^{eq}_A(x_0)G(x,t|x_0,0)
\end{eqnarray*}
By the Feynman-Kac theorem we know the differential equation for $G$, and therefore also the one for $g$:
\[ \frac{\partial g}{\partial t} = \Bigl(-\beta\dot{\lambda}\frac{\partial H}{\partial \lambda}+L_t \Bigr)g \]
It can be easily checked that $g(x,t) \propto e^{-\beta H(x,\lambda_t)} $ solves this differential equation,
and because the initial conditions are $g(x,0) = f^{eq}_A(x)$, we finally arrive at the desired result:
\[ g(x,t) = \left< \delta(x-x_t)e^{-\beta w_t} \right> = \frac{1}{Z_A}e^{-\beta H(x,\lambda_t)} \]
\section{Guessing the direction of the arrow of time}
We conclude with a brief discussion of the nature of irreversibility at the microscopic scale,
framed in the context of a thought experiment that involves guessing the chronology of a sequence of events.
Suppose you are shown a movie of a tennis ball falling freely in the absence of significant frictional forces, for instance near the surface of the moon.
If asked whether the movie is shown as it actually happened, or whether it is being run backward in time, you will not be able to answer; the dynamics of the ball is symmetric in time.
On the other hand, when you are shown a movie of a glass of wine falling
and shattering on the ground, or of the reverse -- shards of broken glass spontaneously assembling themselves and then collecting a stream of wine -- there is no difficulty determining which version depicts the actual sequence of events.
Between these two limiting situations, there are cases like the RNA pulling
experiment, where frictional forces are certainly non-negligible, but thermal fluctuations complicate the picture.
Let us analyze this situation using the results of the previous section.
As with the examples of the tennis ball and glass of wine, suppose you are shown a microscopic ``movie'' depicting an RNA molecule being stretched, from which you are able to reconstruct the trajectory $X$, as defined earlier.
Note that the corresponding time-reversed trajectory $X^{\dagger}$ depicts a realization of the time-reversed process, involving the contraction of the RNA molecule.
From the data -- that is, the movie -- you are asked to guess the direction of time's arrow:
does the movie show the sequence of events as they actually occurred (e.g. stretching) or are you instead observing a movie that was filmed during the reverse process (contraction), now run backward in time?
Let $F$, for ``forward'', and $R$, for ``reverse'', denote these two possibilities, respectively.
Since both $X$ and $X^\dagger$ represent perfectly valid microscopic trajectories, there is no definitive answer to the above question.
Rather, the problem becomes an exercise in likelihood estimation: given the data ($X$, equivalently $X^\dagger$) which of the two hypotheses, $F$ or $R$, is more likely?
The answer is provided by Bayes' rule in probability theory:
assuming we have no {\it a priori} reason to favor one hypothesis over the other, the likelihood $L(F|X)$ of the forward hypothesis, given the observed data $X$, is proportional to the probability to generate that data when performing the forward process:
$L(F|X) \propto P^F(X)$.
The analogous statement holds in the reverse case: $L(R|X^\dagger) \propto P^R(X^\dagger)$.
Combining these with a statement of normalization, namely $L(F|X) + L(R|X^\dagger) = 1$, we conclude that
\begin{eqnarray*}
L(F|X) = \frac{P^F(X)}{P^F(X)+P^R(X^{\dagger})}
\end{eqnarray*}
We now invoke equation (\ref{tussenres}) to arrive at the result,
\begin{equation}
L(F|X) = \frac{1}{1+ e^{-\beta(W-\Delta F)}}
\end{equation}
This result has previously been derived, with a somewhat different intepretation, in the context of free energy estimation \cite{mar07,shi03}.
Recall that $W$ is the work done during the process $X$. In Figure \ref{timesarrow} we see
a graphical depiction of $L(F|X)$.
\begin{figure}[ht]
\includegraphics[width = 10cm]{picture4.eps}
\caption{A graphical depiction of $L(F|X)$.}
\label{timesarrow}
\end{figure}
When $W>\Delta F$, it is more likely that we saw the movie in the right order -- that is, the molecule was actually stretched --
while for $W<\Delta F$, it is more likely that we saw the movie in reverse.
The transition from $L \sim 0$ to $L \sim 1$ occurs over an interval of $W$ of width $\sim k_BT$.
Thus unless the work is within a few $k_BT$ of the free energy difference, we can determine the direction of the arrow of time with near certainty.
This is entirely consistent with our observation that we have no problem distinguishing the direction of time's arrow for macroscopically irreversible processes like that shattering of a glass of wine.
\section*{Acknowledgments}
We are grateful to the organizers of FPSPXII for arranging a highly enjoyable and successful summer school.
C.J. acknowledges financial support from the National Science Foundation (USA), under CHE-0841557.
B.W acknowledges support as Aspirant of FWO, Flanders.
|
\section{High energy QCD and the Color Glass Condensate}
Only 5\% of the mass of the universe is ``bright matter'', but 99\% of
this visible matter is described by QCD, the theory of the strong
interactions. QCD has been described as a nearly perfect theory, with
the only parameters being the masses of the current
quarks~\cite{Wilcz1}. The vast bulk of visible matter is therefore
``emergent'' phenomena arising from the rich dynamics of the QCD
vacuum and the interactions of the fundamental quark and gluon
constituents of QCD. Although we have enough information from a wide
range of experimental data and from numerical lattice computations
that QCD is the right theory of the strong interactions, outstanding
questions remain about how a variety of striking phenomena arise in
the theory. This is because the complexity of the theory also makes
it very difficult to solve.
Our focus in this review is on high energy scattering in QCD. A
traditional approach to these phenomena is to divide them into ``hard''
or ``soft'' scattering, corresponding respectively to large or small
momentum exchanges in the scattering. In the former case, because of
the ``asymptotic freedom'' of QCD, phenomena such as jets can be
computed in a perturbative framework. In the latter case, because the
momentum transfer is small, the ``infrared slavery'' of QCD suggests
that the coupling is large; the scattering therefore is intrinsically
non-per\-tur\-ba\-ti\-ve and therefore not amenable to first principles
analysis. This is problematic because ``soft'' dynamics comprises the
bulk of QCD cross-sections. In contrast, the perturbatively calculable
hard cross-sections are rare processes\footnote{Lattice QCD is not of
much help here because it is best suited to compute static
properties of the theory.}.
We shall argue here that the traditional separation of hard versus
soft QCD dynamics is oversimplified because novel semi-hard scales
generated dynamically at high energies allow one to understand highly
non-perturbative phenomena in QCD using weak coupling methods. To
clarify what we mean, consider inclusive cross-sections in deeply
inelastic scattering (DIS) experiments of electrons off hadrons.
\begin{figure*}[htb!]
\begin{center}
\resizebox*{5cm}{!}{\includegraphics{DIS_kin.ps}}
\hskip 1cm
\raise 16mm\hbox to 5cm{
$\displaystyle{\begin{matrix}
& s&\equiv& (P+k)^2\cr
&Q^2&\equiv& -q^2\cr
&x&\equiv&Q^2 / 2P\cdot q\cr
&xy&\equiv& Q^2/s\cr
\end{matrix}}$\hfil
}
\end{center}
\caption{\baselineskip 15pt \label{fig:DISkinematics}\sl Kinematics of Deep Inelastic
Scattering.}
\end{figure*}
Inclusive cross-sections in DIS (see figure
\ref{fig:DISkinematics}) can be expressed in terms of the Lorentz
invariants i) $x$ which corresponds, at lowest order in perturbation
theory, to the longitudinal momentum fraction carried by a parton in
the hadron, ii) the virtual photon four-momentum squared $q^2 = -Q^2 <0$
exchanged between the electron and the hadron, iii) the inelasticity
$y$, the ratio of the photon energy to the electron energy in the
hadron rest frame, and iv) the center of mass energy squared $s$.
These satisfy the relation $x = Q^2/s y$. Instead of the dichotomy of
``hard'' versus ``soft'', it is useful, for fixed $y$, to consider two asymptotic
limits in DIS that better illustrate the QCD dynamics of high energy
hadron wavefunctions. The first, called the Bjorken limit, corresponds
to fixed $x$ with $Q^2, s\rightarrow \infty$. The second is the
Regge-Gribov limit of fixed $Q^2$, $x\rightarrow 0$ and $s\rightarrow
\infty$.
In the Bjorken limit of QCD, one obtains the parton model wherein the
hadron is viewed, in the infinite momentum frame (IMF), as a dilute
collection of valence quarks and ``wee'' (in the terminology coined
by Feynman) partons--small $x$ gluons and sea quark pairs. Albeit the number of
wee partons is large at high energies, the hadron is dilute (as illustrated in
figure~\ref{fig:HERA-gluon}) because the phase space density is very
small for $Q^2\rightarrow \infty$. In this limit, to leading order in
the coupling constant, the interaction of a probe with the hadron, on
the characteristic time scale $1/Q$, can be expressed as a hard
interaction with an individual parton. In this ``impulse
approximation'', the interaction of the struck parton in the hadron
with co--moving partons is suppressed due to time dilation. The
separation of hard and soft scattering alluded to previously is valid
here. Powerful tools such as the Operator Product Expansion (OPE) and
factorization theorems extend, to high orders in the coupling constant
expansion, the hard-soft separation between process-dependent physics
at the scale $1/Q$ and universal, long distance, non-perturbative Parton
Distribution Functions (PDFs). The evolution of the separation of hard
and soft scales is given by renormalization group (RG) equations
called the DGLAP (Dokshitzer-Gribov-Lipatov-Altarelli-Parisi)
equations~\cite{DGLAP}.
\begin{figure*}[htb!]
\begin{center}
\includegraphics[width=0.43\textwidth]{PDFS_HERA2009.eps}
\hskip 3mm
\includegraphics[width=0.45\textwidth]{phase2_Eng.eps}
\end{center}
\caption{\baselineskip 15pt \label{fig:HERA-gluon}\sl Left: the
$x$-evolution of the gluon, sea quark, and valence quark
distributions for $Q^2=10$ GeV$^2$ measured at HERA~\cite{Aarona1}.
Right: the ``phase--diagram'' for QCD evolution; each colored dot
represents a parton with transverse area $\delta S_\perp \sim 1/Q^2$
and longitudinal momentum $k^+=x P^+$. }
\end{figure*}
The study of strong interactions in the Regge-Gribov limit predates
QCD, and underlies concepts such as soft Pomeron and Reggeon
exchanges, whose dynamics, described by ``Reggeon Field Theory''
(RFT), was believed to encompass much of the phenomena of
multi-particle production. Albeit phenomenologically very
suggestive~\cite{DonnaL1}, the dynamics of RFT is intrinsically
non-perturbative and therefore not easily amenable to systematic
computations. With the advent of high energy colliders, hadron
structure in the Regge-Gribov limit can be explored with $Q^2\geq 1$
GeV$^2$. As strikingly demonstrated by the HERA DIS data shown in
figure~\ref{fig:HERA-gluon}, the gluon distribution $x G(x,Q^2)$ in a
proton rises very fast with decreasing $x$ at large, fixed $Q^2$ --
roughly, as a power $x^{-\lambda}$ with $\lambda\simeq 0.3$. In the
IMF frame of the parton model, $x G(x,Q^2)$ is the number of gluons
with a transverse area $\delta S_\perp \ge 1/Q^2$ and a fraction
$k^+/P^+ \sim x$ of the proton longitudinal momentum\footnote{The
light cone co-ordinates are defined as $k^\pm = (k^0 \pm
k^3)/\sqrt{2}$.}. In the Regge-Gribov limit, the rapid rise of the
gluon distribution at small $x$ is given by the BFKL
(Balitsky-Fadin-Kuraev-Lipatov) equation~\cite{BFKL}, which we will
discuss at some length later.
The stability of the theory formulated in the IMF requires that gluons
have a maximal occupation number of order $1/\alpha_s$. This bound is
saturated for gluon modes with transverse momenta $k_\perp \leq Q_s$,
where $Q_s(x)$ is a semi-hard scale, the ``saturation scale'', that
grows as $x$ decreases. In this novel ``saturation'' regime of
QCD~\cite{saturation}, illustrated in figure~\ref{fig:HERA-gluon}
(right panel), the proton becomes a dense many body system of gluons.
In addition to the strong $x$ dependence, the saturation scale $Q_s$
has an $A$ dependence because of the Lorentz contraction of the
nuclear parton density in the probe rest frame. The dynamics of
gluons in the saturation regime is non-perturbative as is typical of
strongly correlated systems. However, in a fundamental departure from
RFT, this dynamics can be computed using weak coupling methods as a
consequence of the large saturation scale dynamically generated by
gluon interactions. Thus instead of the ``hard'' plus ``soft''
paradigm of the Bjorken limit, one has a powerful new paradigm in the
Regge-Gribov limit to compute the bulk of previously considered
intractable scattering dynamics in hadrons and nuclei.
The Color Glass Condensate (CGC) is the description of the properties
of saturated gluons in the IMF in the Regge-Gribov limit. The
effective degrees of freedom in this framework are color sources
$\rho$ at large $x$ and gauge fields $A^\mu$ at small $x$. At high
energies, because of time dilation, the former are frozen color
configurations on the natural time scales of the strong interactions
and are distributed randomly from event to event. The latter are
dynamical fields coupled to the static color sources. It is the
stochastic nature of the sources, combined with the separation of time
scales, that justify the ``glass'' appellation. The ``condensate''
designation comes from the fact that in the IMF, saturated gluons have
large occupation numbers ${\cal O}(1/\alpha_s)$, with typical momenta
peaked about a characteristic value $k_\perp \sim Q_s$. The dynamical
features of the CGC are captured by the JIMWLK\footnote{JIMWLK
$\equiv$ Jalilian-Marian, Iancu, McLerran, Weigert, Leonidov,
Kovner.} renormalization group (RG) equation that describes how the
statistical distribution $W[\rho]$ of the fast sources at a given $x$
scale evolves with decreasing $x$. The JIMWLK RG is Wilsonian in
nature because weakly coupled fields integrated out at a given step in
the evolution are interpreted as ``induced color charges'' that modify
the statistical weight distribution.
The CGC framework is quite powerful because, given an initial
non-per\-tur\-ba\-ti\-ve distribution of sources at an initial scale
$x_0$, it allows one to compute systematically $n$-point gluon correlation functions
and their evolution with $x$ order by order in perturbation theory. In analogy to parton distribution
functions in the Bjorken limit, the distribution $W[\rho]$ captures
the properties of saturated gluons. Unlike PDFs however, which in the
IMF are parton densities, $W[\rho]$ carries much more information.
Further, in contrast to the ``twist'' expansion of the Bjorken limit
which becomes extremely cumbersome beyond the leading order in $1/Q$,
the CGC framework includes all twist correlations. Like the
PDFs, the $W$'s are universal: the same distributions appear in
computations of inclusive quantities in both lepton-nucleus and
hadron-nucleus collisions.
The CGC is interesting in its own right in what it reveals about the
collective dynamics of QCD at high parton densities. The CGC RG
equations indicate that --at fixed impact parameter-- a proton and a
heavy nucleus become indistinguishable at high energy; the physics of
saturated gluons is universal and independent of the details of the
fragmentation region. This universal dynamics has a correspondence
with reaction--diffusion processes in statistical physics. In
particular, it may lie in a ``spin glass'' universality class.
Understanding the nature of color screening and ``long range order'' in
this universal dynamics offers possibilities for progress in resolving
fundamental QCD questions regarding properties such as confinement and
chiral symmetry. A specific area of progress is in the mapping of the
CGC degrees of freedom to the traditional language of Pomerons and the
consequent prospects for understanding soft QCD dynamics. In nuclear
collisions, CGC dynamics produces ``Glasma'' field configurations at
early times: strong longitudinal chromo-electric and chromo-magnetic
fields color screened on transverse distance scales $1/Q_s$. These generate
long range rapidity correlations, ``sphaleron-like'' topological
fluctuations characterized by large Chern-Simons charge, and
instabilities analogous to those seen in QED plasmas.
The CGC framework is applicable to a wide variety of processes in
e+p/A, p+A and A+A collisions. It provides an {\it ab initio} approach
to study thermalization in heavy ion collisions and the initial
conditions for the evolution of a thermalized Quark Gluon Plasma
(QGP). The interaction of hard probes with the Glasma is little
understood and is important for quantifying the transport properties
of the QGP precursor. A further phenomenological application of the
CGC is to the physics of ultra high energy cosmic rays \cite{cosmic}.
There are several reviews that discuss various aspects of the CGC and
its applications in depth~\cite{reviews1,reviews2,GelisLV2}. Our aim
here is to provide a broad pedagogical overview of the current status
of theory and phenomenology for non-experts. The next section will
focus on the theoretical status, while section 3 discusses the CGC in
the context of DIS, hadronic and nuclear collisions. Section 4
presents the experimental results that probe the dynamics of saturated
gluons. We will end this review with a brief outlook.
\section{Color Glass Condensate: theoretical status}
We will begin this section with an elementary discussion of QCD bremsstrahlung in the
Regge-Gribov limit and use this discussion to motivate the BFKL
equation, how it leads to gluon saturation and describe key
features of saturation. The theoretical status of the CGC effective theory is then
presented. The section ends with a discussion of some advanced
topics that highlight open issues.
\subsection{Parton evolution at high energy}
\label{sec:BFKL}
\begin{figure*}[htb!]
\begin{center}
\includegraphics[width=0.70\textwidth]{saturation.ps}
\end{center}
\caption{\baselineskip 15pt \label{fig:cascade}\sl Left: elementary bremsstrahlung
radiation. Right: high energy scattering with evolution and
recombination. }
\end{figure*}
The structure of a hadron depends upon the scales resolved by an
external probe. Quantum fields bound within the hadron can radiate
other virtual quanta which, by the uncertainty principle, live only
for a short period of time. Therefore, what looks like an isolated
quantum at a given resolution scale reveals a complicated substructure
when probed with finer space--time resolution. The structure of the
hadron in the IMF is specified by the longitudinal ($k^+$) and
transverse ($k_\perp$) momentum distributions of its quanta and by
their correlations--the latter become increasingly important with
increasing density.
In perturbative QCD, parton evolution proceeds via bremsstrahlung,
which favors the emission of {\em soft} and {\em collinear} gluons.
The left part of figure~\ref{fig:cascade} illustrates this elementary
radiation process. When $x\ll 1$, to lowest order in $\alpha_s$, the
differential probability for this emission is given by
\begin{eqnarray}\label{brem} \rmd P_{\rm
Brem}\,\simeq\,\frac{\alpha_s
C_R}{\pi^2}\,\frac{\rmd^2k_\perp}{k_\perp^2}\,\frac{\rmd
x}{x}\; ,
\end{eqnarray}
where $C_R$ is the SU$(N_c)$ Casimir in the color representation of
the emitter-- $N_c$ for a gluon and $(N_c^2-1)/2N_c$ for a quark.
This formula demonstrates the collinear ($k_\perp\to 0$) and soft
($x\to 0$) singularities mentioned above, which produce a logarithmic
enhancement of gluon emission at small $k_\perp$ and/or $x$. If the
emitter at small $x$ were a quark instead a gluon, there would be no
small $x$ enhancement; while the collinear enhancement is present for
either emitter. This asymmetry is due to the spin $1$ nature of the
gluon.
A high energy scattering such as the one illustrated in the right part
of figure~\ref{fig:cascade} probes the number of gluons with a given
$x$ and transverse momenta $k_\perp\le Q$. From eq.~(\ref{brem}), the
gluon number is
\begin{eqnarray}\label{xGxp}
x \frac{\rmd N_g}{\rmd x}(Q^2)\,=\,\frac{\alpha_s
C_R}{\pi}\,\int_{\Lambda_{_{\rm QCD}}^2}^{Q^2}\frac{\rmd k_\perp^2}{k_\perp^2}
\,=\,\frac{\alpha_s C_R}{\pi}\,\ln\left(\frac{Q^2}{\Lambda_{_{\rm QCD}}^2}\right)\; ,
\label{eq:Ng-1}
\end{eqnarray}
where the cutoff $\Lambda_{_{\rm QCD}}$ has been introduced as a crude way to account
for confinement: when confined inside a hadron, a parton has a minimum
virtuality of $\order{\Lambda_{_{\rm QCD}}^2}$.
In QCD, gluons can radiate softer gluons and thus rapidly multiply as
illustrated in figure~\ref{fig:cascade}. Each subsequent emission is
$\alpha_s$ suppressed however when the final value of $x$ is small, these
corrections become large. For a cascade with $n$ intermediate gluons
strongly ordered in $x$, one obtains
\begin{eqnarray}
\alpha_s^n\,\int_{x}^1\frac{\rmd x_n}{x_n}
\int_{x_n}^1\frac{\rmd x_{n-1}}{x_{n-1}}\cdots
\int_{x_2}^1\frac{\rmd x_1}{x_1}
\,=\,\frac{1}{n!}\left(\alpha_s\ln\frac{1}{x}\right)^n
\; .
\end{eqnarray}
When $\alpha_s\ln(1/x)\gtrsim 1$, the correct result for the gluon
distribution is obtained by summing contributions from all such
ladders. The sum {\em exponentiates}, modifying eq.~(\ref{eq:Ng-1})
into
\begin{eqnarray}\label{eq:unintp}
x \frac{\rmd N_g}{\rmd x \rmd k_\perp^2}\,\sim\,\frac{\alpha_s
C_R}{\pi}\,\frac{1}{k_\perp^2}\,{\rm e}^{\omega\alpha_s Y}\; ,\qquad
Y\equiv\ln \frac{1}{x}\;,
\end{eqnarray}
where $\omega$ is a number of order unity which is not fixed by this
rough estimate. The variable $Y$ is known as the {\em rapidity}.
To go beyond this simple power counting argument, one must treat more
accurately the kinematics of the ladder diagrams and include the
associated virtual corrections. The result is the previously
mentioned {\em BFKL equation} for the $Y$-evolution of the
unintegrated gluon distribution. The solution of this equation, which
resums perturbative corrections $(\alpha_s Y)^n$ to all orders,
confirms the exponential increase in eq.~(\ref{eq:unintp}), albeit
with a $k_\perp$-dependent exponent and modifications to the
$k_\perp^{-2}$-spectrum of the emitted gluons.
An important property of the BFKL ladder is {\em coherence in time}.
Because the lifetime of a parton in the IMF, $\Delta x^+\sim
k^+/k_\perp^2\propto x$, the ``slow'' gluons at the lower end of the
cascade have a much shorter lifetime than the preceding ``fast''
gluons. Therefore, for the purposes of small $x$ dynamics, fast
gluons with $x'\gg x$ act as {\em frozen color sources emitting
gluons at the scale $x$}. Because these sources may overlap in the
transverse plane, their color charges add coherently, giving rise to
a large color charge density. The {\em average} color charge density
is zero by gauge symmetry but fluctuations in the color charge
density are nonzero and increase rapidly with $1/x$.
These considerations are at the heart of the CGC reformulation of
BFKL evolution. However, in contrast to the original BFKL
formulation, the CGC formalism \cite{MV,Balit1,Kovch3,CGC1,CGC2}
includes {\em non--linear} effects which appear when
the gluon density becomes large.
The quantity which controls gluon interactions in the IMF is their
{\em occupation number}--the number of gluons of a given color per
unit transverse phase--space and per unit rapidity,
\begin{eqnarray}\label{phidef}
n(Y,k_\perp,b_\perp)\,\equiv\,\frac{(2\pi)^3}{2(N_c^2-1)}\, \,\frac{{\rm
d} N_g}{\rmd Y\, {\rm d}^2{\boldsymbol k}_\perp\,{\rm d}^2{\boldsymbol b}_\perp}\;,
\end{eqnarray}
where the impact parameter ${\boldsymbol b}_\perp$ is the gluon position in the
transverse plane. If $n\ll 1$, the system is dilute and gluon
interactions are negligible. When $n\sim\order{1}$ gluons start
overlapping, but their interactions are suppressed by $\alpha_s\ll 1$.
The interaction strength becomes of order one when $n\sim
\order{1/\alpha_s}$. It is then that non--linear effects become
important leading to gluon saturation.
Gluon occupancy is further amplified if instead of a proton we
consider a large nucleus with atomic number $A\gg 1$. The transverse
area scales like $A^{2/3}$, and the gluon occupation number scales as
$A^{1/3}$. Thus, for a large nucleus, saturation effects become
important at larger values of $x$ than for a proton.
To be more specific, let us discuss a simplified version of the
non--linear evolution equation for the occupation number. Consider an
elementary increment in rapidity: $Y\to Y+\rmd Y$. Each preexisting
gluon in the hadron has a probability $\alpha_s\rmd Y$ to emit an
additional soft gluon -- the average increase is $\rmd n\sim \alpha_s
n\rmd Y$. Further, the emission vertex is non-local in $k_\perp$
because the transverse momentum of the parent gluon is shared among
the two daughter gluons. At high-energies, this non-locality is well
approximated as a {\em diffusion} in the logarithmic variable $t\equiv
\ln(k_\perp^2)$. Finally, two preexisting gluons can merge and produce
a single final gluon with rapidity $Y+\rmd Y$. This process is
quadratic in $\alpha_s n$ and leads to a negative term in the evolution
equation. Adding up these three effects, one obtains the evolution
equation
\begin{equation}
\frac{\partial n}{\partial Y}
\simeq
\omega\alpha_s n
\,+\,\chi\alpha_s\partial_t^2 n\,-\,\beta \alpha_s^2 n^2\; ,
\label{eq:GLR}
\end{equation}
where $\omega$, $\chi$, and $\beta$ are numbers of order unity. This
equation mimics the BFKL equation \cite{BFKL} if one drops the term
quadratic in $\alpha_s n$ in the r.h.s., and mimics its non-linear extensions
that include saturation if one keeps the quadratic term.
\subsection{Generic features of gluon saturation}
Although a toy model, \eqnum{eq:GLR} captures essential features of
saturation\footnote{Our discussion oversimplifies the mechanism for
gluon saturation. In the saturation regime, gluons form
configurations that screen their color charge over transverse scales
$\sim 1/Q_s$ \cite{IancuM1,screening}. Thus, gluons with momentum
$k_\perp\simle Q_s$ are emitted from a quasi-neutral patch of color
sources, and their occupation number grows only {\em linearly} in
$Y$ \cite{Muell3,IancuM1} -- much slower than the exponential growth
in the region $k_\perp\gg Q_s$.}. If $\alpha_s n\ll 1$, one can neglect the quadratic term, and eq.~(\ref{eq:GLR})
predicts an exponential growth in $Y$ of the gluon occupation number.
But when $\alpha_s n\sim 1$, the negative non-linear term turns on and
tames the growth. In fact, eq.~(\ref{eq:GLR}) has a fixed point at
$n=\omega/(\beta\alpha_s)$ where its r.h.s. vanishes--the evolution stops
when this value of $n$ is reached, resulting in gluon saturation in
the spirit of early works~\cite{saturation}.
Eq.~(\ref{eq:GLR}) also reveals the emergence of a transverse momentum
scale $Q_s(Y)$ that characterizes saturation. This scale is the
$k_\perp$ where gluon occupancy becomes of $\order{1/\alpha_s}$. As a
function of transverse momentum, the occupation number $n(Y,k_\perp)$
is $\order{1/\alpha_s}$ if $k_\perp \lesssim Q_s(Y)$, and decreases
rapidly above $Q_s(Y)$ ($n\propto 1/k_\perp^2$ for $k_\perp\gg
Q_s(Y)$). The shape of $n(Y,k_\perp)$ as a function of $k_\perp$ is
known as the {\sl saturation front}. Note that $Q_s$, the typical
gluon transverse momentum at the rapidity $Y$, increases with
energy and becomes a semi-hard scale ($Q_s(Y)\gg\Lambda_{_{\rm QCD}}$) at sufficiently high energy.
Eq.~(\ref{eq:GLR}) belongs to the generic class of {\em
reaction-diffusion processes}~\cite{Saarl1}. These are processes
where an entity can hop to neighboring locations (diffusion term
$\chi\alpha_s\partial_t^2 n$), can split into two identical entities
(the term $\omega\alpha_s n$), and where two entities can merge into a single
one (the term $\beta \alpha_s^2 n^2$). In the limit of large occupation
numbers, these processes admit the mean field description of
eq.~(\ref{eq:GLR}), which in the context of statistical physics is
known as the {\em Fisher-Kolmogorov-Petrovsky-Piscounov (FKPP)
equation}~\cite{Saarl1}. In QCD, the closest equation of this type
is the Balitsky-Kovchegov (BK) equation~\cite{Kovch3}. The
correspondence between the BK and FKPP equations, originally noticed
in \cite{FKPP}, clarifies the properties of saturation fronts in QCD
in analogy with known properties of reaction--diffusion processes.
A crucial property is the emergence of {\em traveling waves}. The
saturation front generated by this equation propagates without
distortion at constant speed; one has $n(Y,t)=n(t-\lambda_s Y)$ with
$\lambda_s$ a constant. This property has been verified in numerical
studies of the BK equation~\cite{BKnum,RummuW1} and in analytic
studies of the BFKL equation in the presence of a saturation
boundary~\cite{BFKLboundary,Trian1}. It provides a natural explanation
of the {\em geometric scaling} phenomenon observed in the HERA
data~\cite{StastGK1,GelisPSS1} (see section \ref{sec:coll-DIS}). In
QCD, a front moving with constant speed $\lambda_s$ is equivalent to
the saturation momentum increasing exponentially with $Y$,
\begin{eqnarray}\label{Qsat}
Q_s^2(Y)\simeq Q_0^2\,{\rm e}^{\lambda_s Y}\quad{\rm with}\quad
\lambda_s \approx 4.9\,\alpha_s\; ,
\end{eqnarray}
where $Q_0$ is some non-perturbative initial scale. For a large
nucleus, $Q_0^2$ scales like $A^{1/3}$ as does $Q_s^2(Y)$ for any $Y$.
This form of the saturation momentum is modified to $Q_s^2=Q_0^2\,
e^{\sqrt{\lambda(Y+Y_0)}}$ when the running of the strong coupling is
taken into account; see ~\cite{Trian1} for a detailed study of higher
order effects on the energy dependence of $Q_s$.
\subsection{The Color Glass Condensate}
The CGC is an {\em effective field theory (EFT)} based on the
separation of the degrees of freedom into fast frozen color sources
and slow dynamical color fields~\cite{MV}. A {\em renormalization
group equation} --the JIMWLK equation \cite{CGC1,CGC2}-- ensures the
independence of physical quantities with respect to the cutoff that
separates the two kinds of degrees of freedom.
The fast gluons with longitudinal momentum $k^+>\Lambda^+$ are frozen
by Lorentz time dilation in configurations specified by a color
current $J^\mu_a \equiv \delta^{\mu +}\rho^a$, where
$\rho^a(x^-,x_\perp)$ is the corresponding color charge density. On
the other hand, slow gluons with $k^+<\Lambda^+$ are described by the
usual gauge fields $A^\mu$ of QCD. Because of the hierarchy in $k^+$
between these two types of degrees of freedom, they are coupled
eikonaly by a term $J_\mu A^\mu$. The fast gluons thus act as
sources for the fields that represent the slow gluons. Although it is
frozen for the duration of a given collision, the color source density
$\rho^a$ varies randomly event by event. The CGC provides a gauge
invariant distribution $W_{\Lambda^+}[\rho]$, which gives the
probability of a configuration $\rho$. This functional encodes all the
correlations of the color charge density at the cutoff scale
$\Lambda^+$ separating the fast and slow degrees of freedom. Given
this statistical distribution, the expectation value of an operator
at the scale $\Lambda^+$ is given by
\begin{equation}
\left<{\cal O}\right>_{\Lambda^+}\equiv
\int\big[D\rho\big]\;W_{\Lambda^+}\big[\rho\big]\;{\cal O}\big[\rho\big]\; ,
\end{equation}
where ${\cal O}[\rho]$ is the expectation value of the operator for a
particular configuration $\rho$ of the color sources.
The power counting of the CGC EFT is such that in the saturated regime
the sources $\rho$ are of order $g^{-1}$. Attaching an additional
source to a given Feynman graph does not alter its order in $g$; the
vertex where this new source attaches to the graph is compensated by
the $g^{-1}$ of the source. Thus, computing an observable at a certain
order in $g^2$ requires the resummation of all the contributions
obtained by adding extra sources to the relevant graphs. The leading
order in $g^2$ is given by a sum of tree diagrams, which can be
expressed in terms of classical solutions of the Yang-Mills equations.
Moreover, for inclusive observables~\cite{classical}, these classical
fields obey a simple boundary condition: they vanish when $t\to
-\infty$.
Next-to-leading order (NLO) computations in the CGC EFT involve a sum
of one-loop diagrams embedded in the above classical field. To prevent
double counting, momenta in loops are required to be below the cutoff
$\Lambda^+$. This leads to a logarithmic dependence in $\Lambda^+$ of
these loop corrections. These logarithms are large if $\Lambda^+$
is well above the typical longitudinal momentum scale of the
observable considered, and must be resummed.
For gluon correlations inside the the hadron wavefunction and also for
sufficiently inclusive observables in a collision, the leading
logarithms are universal and can be absorbed into a redefinition of
the distribution $W_{\Lambda^+}[\rho]$ of the hard sources. The
evolution of $W_{\Lambda^+}[\rho]$ with $\Lambda^+$ is governed by the
functional JIMWLK equation
\begin{eqnarray}
\frac{\partial\, W_{\Lambda^+}[\rho] }{{\partial \ln(\Lambda^+)}}=
-{\cal H}\left[\rho,\frac{\delta}{ {\delta \rho}}\right]\,W_{\Lambda^+}[\rho]\;,
\end{eqnarray}
where ${\cal H}$ is known as the JIMWLK Hamiltonian. This operator
contains up to two derivatives $\partial/\partial\rho$, and arbitrary powers
in $\rho$. Its explicit expression can be found in
refs.~\cite{CGC1,CGC2,reviews1}. The derivation of the JIMWLK
equation will be sketched in the section \ref{sec:coll-DIS}.
Numerical studies of JIMWLK evolution were performed in
\cite{RummuW1,KovchKRW1}. An analytic, albeit formal, solution to the
JIMWLK equation was constructed in \cite{BlaizIW1} in the form of a
path integral. Alternatively, the evolution can can be expressed as
an infinite hierarchy of coupled non-linear equations for $n$-point
Wilson line correlators--often called the Balitsky
hierarchy~\cite{Balit1}. In this framework, the BK equation is a mean
field approximation of the JIMWLK evolution, valid in the limit of a
large number of colors $N_c\to\infty$. Numerical studies of the JIMWLK
equation~\cite{RummuW1,KovchKRW1} have found only small differences
with the BK equation.
Let us finally comment on the initial condition for the JIMWLK
equation which is also important in understanding its derivation. The
evolution should start at some cutoff value in the longitudinal
momentum scale $\Lambda^+_0$ at which the saturation scale is already
a (semi)hard scale, say $Q_{s0}\gtrsim 1$~GeV, for perturbation theory
to be applicable. The gluon distribution at the starting scale is in
general non--perturbative and requires a model. A physically
motivated model for the gluon distribution in a large nucleus is the
McLerran-Venugopalan model~\cite{MV}. In a large nucleus, there is a
window in rapidity where evolution effects are not large but $x$ is
still sufficiently small for a probe not to resolve the longitudinal
extent of the nucleus. In this case, the probe ``sees'' a large number
of color charges, proportional to $A^{1/3}$. These charges add up to
form a higher dimensional representation of the gauge group, and can
therefore be treated as classical color
distributions~\cite{MV,JeonV1}. Further, the color charge
distribution $W_{\Lambda_0^+}[\rho]$ is a Gaussian
distribution\footnote{There is a additional term, corresponding to the
cubic Casimir; which is parametrically suppressed for large
nuclei~\cite{JeonV2}. This term generates Odderon excitations in the
JIMWLK/BK evolution~\cite{odderon}.} in $\rho$. The variance of this
distribution --the color charge squared per unit area-- is
proportional to $A^{1/3}$ and provides a semi-hard scale that makes
weak coupling computations feasible. In addition to its role in
motivating the EFT and serving as the initial condition in JIMWLK
evolution, the MV model allows for direct phenomenological studies in
p+A and A+A collisions in regimes where the values of $x$ are not so
small as to require evolution.
The dynamics of small x gluons in QCD may be universal in more than
one sense~\cite{McLer2}. A weak form of this universality is that
their dynamics in both hadrons and nuclei is controlled only by the
saturation scale with its particular dependence on energy and nuclear
size. A stronger form of the universality is noticed in particular for
the solution of the BK equation with running coupling effects; the
saturation scale, for both hadrons and nuclei, at fixed impact
parameter, becomes the same asymptotically with increasing
energy~\cite{Muell7}. The strongest form of the universality is that
the RG flows in the saturation regime have a fixed point corresponding
to universal ``critical'' exponents describing the behavior of
multi-parton correlation functions. As discussed further below, the RG
equations for high energy QCD lie in a wide class of
reaction-diffusion processes which have universal properties
remarkably close to those of spin glasses~\cite{FKPP1}.
\subsection{Advanced Theory topics}
Reaction--diffusion processes exhibit an extreme sensitivity to
{\em particle number fluctuations} \cite{sFKPP,MuellS2,IancuMM1},
generated by gluon splittings, which produce correlations among pairs
of gluons \cite{ploop}. This effect is of higher order in $\alpha_s$, and
is linear in $n$ since it results from the splitting of a single
gluon. Conversely, producing two gluons without this splitting leads
to a term that has one less power of $\alpha_s$, but of order $n^2$.
Thus, the splitting contribution is important in the {\em dilute}
regime where $n\lesssim \alpha_s$. Since, as mentioned previously, the
dynamics of the saturation front is driven by the BFKL growth of its
dilute tail, these fluctuations play an important role in the 2--gluon
density $\langle nn\rangle$ at high energy.
As manifest in \eqnum{eq:GLR}, this 2-gluon density enters
in the non--linear term leading to saturation of the single gluon
density. Thus, gluon number fluctuations in the dilute regime can
strongly influence the approach towards saturation. For instance,
as argued in \cite{sFKPP} for generic reaction--diffusion processes, and
independently in the QCD context\cite{MuellS2}, these
fluctuations reduce the (average) speed of the saturation front.
Besides making the value of $Q_s$ a fluctuating quantity, they tend to
wash out the geometric scaling property of the individual fronts
\cite{IancuMM1,ploop}. Both effects are quantitatively
important, as shown by explicit numerical simulations within various
reaction--diffusion models (including those inspired by the QCD
dynamics at high energy \cite{IancuAST1}).
There is presently no general theory that includes BFKL ladders,
saturation and fluctuations. (In terms of Feynman graphs, this
corresponds to resumming ``Pomeron loops'' diagrams to all orders
\cite{ploop}.) Attempts to construct such a theory
\cite{ploop,MuellSW3,ploop1} have led to incomplete formalisms that
are difficult to exploit in phenomenological applications.
Fortunately, the effect of these fluctuations is considerably reduced
by the running of the strong coupling \cite{DumitIPST1}, which tends
to postpone their importance to unrealistically large rapidities.
Thus, for practical applications at least up to LHC energies, a
sufficient theory for the approach towards saturation is the
leading--order mean-field evolution extended with running coupling
corrections. This theory has developed significantly as the
running--coupling version of the BK equation has been constructed in
\cite{BKrunning} and successfully applied to studies of the
phenomenology at HERA \cite{AlbacAMS1}.
\section{Collisions in the CGC framework}
The CGC is an effective theory for the wavefunction of a high-energy
hadron or nucleus. In this section, we apply it, with particular
emphasis on factorization, to deeply inelastic scattering and hadronic
collisions. In A+A collisions, the formation of the Glasma and its key
features are emphasized.
\subsection{The CGC and DIS at small $x$}
\label{sec:coll-DIS}
At small $x$, corresponding to large Ioffe times~\cite{ioffe}, DIS is
characterized by the fluctuation of the virtual photon into a
quark--antiquark pair which then scatters off the hadronic or nuclear
target. The inclusive DIS cross-section can be expressed
as~\cite{NikolZ1}
\begin{equation}
\sigma_{\gamma^*T}
=\int_0^1 \rmd z\int \rmd^2{\boldsymbol r}_\perp
{
\left|\psi(z,{\boldsymbol r}_\perp)\right|^2
}\;
\sigma_{\rm dipole}(x,{\boldsymbol r}_\perp)
\; ,
\label{eq:X-DIS}
\end{equation}
where $\psi(z,{\boldsymbol r}_\perp)$ is the $q\bar{q}$ component of the
wave-function of the virtual photon (known from QED) and $\sigma_{\rm
dipole}(x,{\boldsymbol r}_\perp)$ is the QCD ``dipole'' cross-section for the
quark-antiquark pair to scatter off the target. This process is shown
in figure~\ref{fig:DIS-LO}, where we have assumed that the target
moves in the $-z$ direction. In the leading order (LO) CGC description
of DIS, the target is described, as illustrated in
figure~\ref{fig:CGC-1}, as static sources with $k^->\Lambda_0^-$. The
field modes do not contribute at this order.
\begin{figure*}[htb!]
\begin{center}
\includegraphics[width=0.25\textwidth]{DIS-LO.eps}
\hskip 20mm
\includegraphics[width=0.25\textwidth]{DIS-NLO.eps}
\end{center}
\caption{\baselineskip 15pt \label{fig:DIS-LO}\sl Left: leading Order (LO) contribution
to DIS off the CGC. Right: NLO contribution.}
\end{figure*}
Employing the optical theorem, $\sigma_{\rm dipole}(x,{\boldsymbol r}_\perp)$ can
be expressed in terms of the forward scattering amplitude ${\boldsymbol
T({\boldsymbol x}_\perp,{\boldsymbol y}_\perp)}$ of the $q\bar{q}$ pair at LO as
\begin{equation}
\sigma_{\rm dipole}^{_{\rm LO}}(x,{\boldsymbol r}_\perp)
=
2 \int \; d^2{\bf b}\;\int \; [D\rho] W_{\Lambda_0^-}[\rho]\;
{\boldsymbol T}_{_{\rm LO}}({\boldsymbol b}+\frac{{\boldsymbol r}_\perp}{2}, {\boldsymbol b} - \frac{{\boldsymbol r}_\perp}{2})\; ,
\label{eq:opt-thm-LO}
\end{equation}
where, for a fixed configuration of the target color sources
\cite{McLerV4,Venug1}
\begin{equation}
{ {\boldsymbol T}_{_{\rm LO}}({\boldsymbol x}_\perp,{\boldsymbol y}_\perp)}
=
1-\frac{1}{N_c}\,{\rm tr}\,( U({\boldsymbol x}_\perp)U^\dagger({\boldsymbol y}_\perp))\; ,
\label{eq:Wilson-amp}
\end{equation}
with $U({\boldsymbol x}_\perp)$ a Wilson line representing the interaction between
a quark and the color fields of the target, defined to be
\begin{equation}
U({\boldsymbol x}_\perp)
=
{\rm T}\,\exp i{g}\int^{1/x P^-}
dz^+\,{{\cal A}^-(z^+,{\boldsymbol x}_\perp)}\; .
\end{equation}
In this formula, ${\cal A}^-$ is the minus component of the gauge
field generated (in Lorenz gauge) by the sources of the target; it is
obtained by solving classical Yang-Mills equations with these sources.
The upper bound $xP^-$ (where $P^-$ is the target longitudinal
momentum and $x$ the kinematic variable defined in
figure~\ref{fig:DISkinematics}) indicates that source modes with
$k^-<xP^-$ do not contribute to this scattering amplitude. Thus if the
cutoff $\Lambda_0^-$ of the CGC EFT is lower than $xP^-$, ${\boldsymbol
T}_{_{\rm LO}}$ is independent of $\Lambda_0^-$.
However, when $\Lambda_0^-$ is larger than $xP^-$, the dipole
cross-section is in fact independent of $x$ (since the CGC EFT does
not have source modes near the upper bound $xP^-$) and depends on the
unphysical parameter $\Lambda_0^-$. As we shall see now, this is
related to the fact that eq.~(\ref{eq:opt-thm-LO}) is incomplete and
receives large corrections from higher order diagrams. Consider now
the NLO contributions (one of them is shown in the right panel in
figure \ref{fig:DIS-LO}) with gauge field modes in the slice
$\Lambda_1^-\le k^- \le \Lambda_0^-$ (see figure~\ref{fig:CGC-1}).
\begin{figure*}[htb!]
\begin{center}
\centerline{\epsfig{file=CGC-0.eps,width=7cm}}
\vskip 2mm
\centerline{\epsfig{file=CGC-1.eps,width=7cm}}
\caption{\baselineskip 15pt \label{fig:CGC-1}\sl Top: sources and fields in the CGC
effective theory. Bottom: NLO correction from a layer of field
modes just below the cutoff.}
\end{center}
\end{figure*}
An explicit computation of the contribution of field modes in this
slice gives
\begin{equation}
{\delta {\boldsymbol T}_{_{\rm NLO}}({\boldsymbol x}_\perp,{\boldsymbol y}_\perp)}
=
\ln\left(\frac{\Lambda_0^-}{\Lambda_1^-}\right)\;
{{\cal H}}\;
{{\boldsymbol T}_{_{\rm LO}}({\boldsymbol x}_\perp,{\boldsymbol y}_\perp)}\; ,
\label{eq:DIS-NLO}
\end{equation}
where ${\cal H}$ is the JIMWLK Hamiltonian. All dependence on the
cutoff scales is in the logarithmic prefactor alone. This Hamiltonian
has two derivatives with respect to the classical field ${\cal
A}\sim{\cal O}(1/g)$; ${\cal H}\,{\boldsymbol T}_{_{\rm LO}}$ is of order
$\alpha_s {\boldsymbol T}_{_{\rm LO}}$ and therefore clearly an NLO
contribution. However, if the new scale $\Lambda_1^-$ is such that
$\alpha_s\ln(\Lambda_0^-/\Lambda_1^-)\sim 1$, this NLO term becomes
comparable in magnitude to the LO contribution. Averaging the sum of
the LO and NLO contributions over the distribution of sources at the
scale $\Lambda_0^-$, one obtains
\begin{eqnarray}
\int[D\rho]\;W_{\Lambda_0^-}[\rho]\;
\left({\boldsymbol T}_{_{\rm LO}}+\delta{\boldsymbol T}_{_{\rm NLO}}\right)=
\int[D\rho]\;W_{\Lambda_1^-}[\rho]\;{\boldsymbol T}_{_{\rm LO}}\; ,
\label{eq:int-by-parts}
\end{eqnarray}
where $W_{\Lambda_1^-}\equiv (1+\ln(\Lambda_0^-/\Lambda_1^-)\,{\cal
H})\,W_{\Lambda_0^-}$. We have shown here that the NLO correction
from quantum modes in the slice $\Lambda_1^-\le k^- \le \Lambda_0^-$
can be absorbed in the LO term, provided we now use a CGC effective
theory at $\Lambda_1^-$ with the modified distribution of sources
shown in eq.~(\ref{eq:int-by-parts}). In differential form, the
evolution equation of the source distribution,
\begin{equation}
\frac{\partial}{\partial\ln(\Lambda^-)}W_{\Lambda^-}
=-{\cal H}\,W_{\Lambda^-}\; ,
\label{eq:JIMWLK-diff}
\end{equation}
is the JIMWLK equation.
Repeating this elementary step, one progressively resums quantum
fluctuations down to the scale $k^-\sim xP^-$. Thanks to
eq.~(\ref{eq:int-by-parts}), the result of this resummation for the
dipole cross-section is formally identical to
eq.~(\ref{eq:opt-thm-LO}), except that the source distribution is
$W_{xP^-}$ instead of $W_{\Lambda_0^-}$. Note that if one further
lowers the cutoff below $xP^-$, the dipole cross-section remains
unchanged.
\subsection{The CGC in p+A collisions}
\begin{figure}[htb!]
\begin{center}
\includegraphics[width=0.36\textwidth]{pA.eps}
\hskip 5mm
\includegraphics[width=0.54\textwidth]{AA-example.eps}
\end{center}
\caption{\baselineskip 15pt \label{fig:pA}\sl Left: sketch of a proton-nucleus
collision. Right: example of leading order contribution in a
nucleus-nucleus collision.}
\end{figure}
Collisions between a dilute hadron projectile and a dense hadron
target can be studied semi-analytically in the CGC framework. The
archetype of such collisions is a proton-nucleus collision. However,
the dilute-dense treatment also applies to proton-proton collisions
for measurements at forward rapidities where the wavefunction of one
of the projectiles is probed at large $x$ and that of the other at
small $x$. These asymmetrical collisions can be treated using
conventional parton distributions for the proton and the CGC for the
nucleus.
The simplest quantity to compute in this context is the single
particle inclusive spectrum. In the CGC framework, this process is due
to the scattering of a parton from the proton off the color field of
the target nucleus, as illustrated in figure~\ref{fig:pA} (left). When
the parton is a quark, the amplitude for the scattering is
proportional to the Fourier transform of a Wilson line,
\begin{equation}
{\cal M}_{_{\rm LO}}\propto \int d^2{\boldsymbol x}_\perp\;
e^{i{\boldsymbol p}_\perp\cdot{\boldsymbol x}_\perp}\; U({\boldsymbol x}_\perp)\; .
\label{eq:pA-1incl-q}
\end{equation}
Squaring this amplitude, summing over the color of the quark in the
final state and averaging over the color of the incoming quark, we get
\begin{equation}
\left|{\cal M}\right|^2_{_{\rm LO}}
\propto
\int d^2{\boldsymbol b} \; d^2{\boldsymbol r}_\perp\;
e^{i{\boldsymbol p}_\perp\cdot{\boldsymbol r}_\perp}\;
{\boldsymbol T}_{_{\rm LO}}\left({\boldsymbol b}+\frac{{\boldsymbol r}_\perp}{2},{\boldsymbol b}-\frac{{\boldsymbol r}_\perp}{2}\right)\; ,
\label{eq:tmp1}
\end{equation}
where ${\boldsymbol T}_{_{\rm LO}}$ is the dipole scattering amplitude (see
eq.~(\ref{eq:Wilson-amp})) that already appeared in our discussion of
DIS. Because the same quantity appears here, the treatment of the NLO
corrections we described in the DIS case is similarly applicable; one
integrates out softer modes by lowering the cutoff of the CGC EFT
letting $W_{\Lambda^-}[\rho]$ evolve according to the JIMWLK equation.
The scale to which one evolves the cutoff is $\Lambda^-=xP^-$ with
$x=(p_\perp/\sqrt{s})\exp(-y)$ where $p_\perp$ is the transverse
momentum of the scattered parton and $y$ its rapidity. Therefore, in
the CGC framework, the cross-section for this process is simply the
Fourier transform of the dipole cross-section $\sigma_{\rm
dipole}(x,{\boldsymbol r}_\perp)$. In p+A collisions, final states containing a
photon or a lepton pair can be similarly expressed in terms of the
same Fourier transform \cite{photon}.
An additional contribution to the single inclusive particle spectrum
in a p+A collision is due to an incoming gluon instead of a quark. The
treatment is nearly identical to the incoming quark case, except that
in eq.~(\ref{eq:pA-1incl-q}) one must replace the Wilson line in the
fundamental representation by a Wilson line in the adjoint
representation.
Similar calculations can be performed for processes with more
complicated final states, such as the production of a quark-antiquark pair.
Although this observable has a more complicated expression (containing terms that are the product of four Wilson lines) its NLO
corrections still comply with eq.~(\ref{eq:DIS-NLO}), which ensures
their factorization into the distribution of sources
$W_{\Lambda^-}[\rho]$. The crucial ingredient for factorization to
work is to consider an observable that is sufficiently inclusive to
allow the corresponding final state to be accompanied by an arbitrary
number of gluons. Any restriction on associated gluon radiation will
not permit NLO corrections to factorize simply in the distribution of
sources.
A much more limited but widely used form of factorization is $k_\perp$
factorization~\cite{KTfact} in terms of $k_\perp$ dependent
unintegrated quark and gluon distributions of the projectile and
target. Within the CGC framework, these results can be reproduced for
gluon~\cite{KovchM3,DumitM1} and heavy quark~\cite{GelisV1}
distributions at large transverse momenta $k_\perp \geq Q_s$; however,
at smaller transverse momenta, $k_\perp$ factorization is broken even
at leading order~\cite{KrasnV2,BlaizGV2,NikolSZ1}.
\subsection{Shattering CGCs in A+A collisions}
\label{sec:AA}
Collisions between two nuclei (``dense-dense'' scattering) are
complicated to handle on the surface. However, in the CGC framework,
because the wave functions of the two nuclei are saturated, the
collision can be treated as the collision of classical fields. This insight significantly simplifies the treatment of A+A scattering.
The classical fields are coupled to fast partons of each nucleus respectively described by the external
current $J^\mu=\delta^{\mu+}\rho_1+\delta^{\mu-}\rho_2$. The source
densities of fast partons $\rho_{1,2}$ are both parametrically of
order $1/g$, which implies that graphs involving multiple sources from
both projectiles must be resummed. (See the right panel of
figure~\ref{fig:pA} for an illustration.)
At leading order, inclusive observables\footnote{Exclusive observables
may also be expressed in terms of solutions of the same Yang-Mills
equations, but with more complicated boundary conditions than for
inclusive observables.} depends on the retarded classical color
field ${\cal A}^\mu$, which solves the Yang-Mills equations $[{\cal
D}_\mu,{\cal F}^{\mu\nu}]=J^\nu$ with the boundary condition
$\lim_{x^0\to -\infty}{\cal A}^\mu =0$. Among the observables to which
this result applies is the expectation value of the energy-momentum
tensor at early times after the collision. At leading order,
\begin{equation}
T^{\mu\nu}_{_{\rm LO}}
=
\frac{1}{4}g^{\mu\nu}\,{{\cal F}^{\lambda\sigma}{\cal F}_{\lambda\sigma}}
-{{\cal F}^{\mu\lambda}{\cal F}^\nu{}_\lambda}\; ,
\end{equation}
where ${\cal F}^{\mu\nu}$ is the field strength of the classical field
${\cal A}^\mu$.
Although A+A collisions are more complicated than e+A or p+A
collisions, one can still factorize the leading higher order
corrections into the evolved distributions $W_{\Lambda^-}[\rho_1]$ and
$W_{\Lambda^+}[\rho_2]$. At the heart of this factorization is a
generalization of eq.~(\ref{eq:DIS-NLO}) to the case where the two
projectiles are described in the CGC framework \cite{factorization}.
When one integrates out the field modes in the slices
$\Lambda_1^\pm\le k^\pm\le \Lambda_0^\pm$, the leading correction to
the energy momentum tensor is
\begin{equation}
{\delta T^{\mu\nu}_{_{\rm NLO}}}
=
\Big[
\ln\left(\frac{\Lambda_0^-}{\Lambda_1^-}\right)\,{{\cal H}_1}
+
\ln\left(\frac{\Lambda_0^+}{\Lambda_1^+}\right)\,{{\cal H}_2}
\Big]\;{T^{\mu\nu}_{_{\rm LO}}}\; ,
\label{eq:AA-NLO}
\end{equation}
where ${\cal H}_{1,2}$ are the JIMWLK Hamiltonians of the two nuclei
respectively. What is crucial here is the absence of mixing between
the coefficients ${\cal H}_{1,2}$ of the logarithms of the two
projectiles; they depend only on $\rho_{1,2}$ respectively and not on
the sources of the other projectile. Although the proof of this
expression is somewhat involved, the absence of mixing is deeply
rooted in causality. The central point is that because the duration
of the collision (which scales as the inverse of the energy) is so
brief, soft radiation must occur before the two nuclei are in causal
contact. Thus logarithms associated with this radiation must have
coefficients that do not mix the sources of the two projectiles.
Following the same procedure for eq.~(\ref{eq:AA-NLO}), as for the e+A
and p+A cases, one obtains for the energy-momentum tensor in an A+A
collision the expression
\begin{equation}
\left<T^{\mu\nu}\right>_{_{\rm LLog}}
=
\int
\big[D{\rho_{_1}}\,D{\rho_{_2}}\big]
\;
{ W_1\,[\rho_{_1}\big]}\;
{ W_2\big[\rho_{_2}\big]}
\;
T^{\mu\nu}_{_{\rm LO}}\; .
\label{eq:Tmunu}
\end{equation}
This result can be generalized to multi-point correlations of the
energy-mo\-men\-tum tensor,
\begin{eqnarray}
\left<T^{\mu_1\nu_1}(x_1)\cdots T^{\mu_n\nu_n}(x_n)\right>_{_{\rm LLog}}
&=&
\int
\big[D{\rho_{_1}}\,D{\rho_{_2}}\big]
\;
{ W_1\,[\rho_{_1}\big]}\;
{ W_2\big[\rho_{_2}\big]}
\nonumber\\
&&\qquad\times
T^{\mu_1\nu_1}_{_{\rm LO}}(x_1)\cdots T^{\mu_n\nu_n}_{_{\rm LO}}(x_n)\; .
\label{eq:Tmunu2}
\end{eqnarray}
In this expression, all the correlations between the energy-momentum
tensor at different points are from the distributions
$W_{1,2}[\rho_{1,2}]$. Thus, the leading correlations are already
built into the wavefunctions of the projectiles prior to the
collision.
The expressions in eqs.~(\ref{eq:Tmunu}) and (\ref{eq:Tmunu2}) are
valid for proper times $\tau\sim 1/Q_s$ after the heavy ion collision.
The energy-momentum tensor, for each configuration of sources
$\rho_{1,2}$ is determined by solving classical Yang--Mills equations
to compute the gauge fields ${\cal A}_\mu^{\rm cl.}[\rho_1,\rho_2]$ in
the forward light cone with initial conditions determined by the
classical CGC fields of each of the nuclei at
$\tau=0$~\cite{class0,KrasnV2,KrasnNV1,Lappi,class1}.
The corresponding non-equilibrium matter, with high occupation numbers
$\sim 1/\alpha_s$ is called the Glasma~\cite{LappiM1}. The Glasma
fields at early times are longitudinal chromo-electric and
chromo-magnetic fields that are screened at distances $~1/Q_s$ in the
transverse plane of the collision. As a consequence, the matter
produced can be visualized (see figure~\ref{fig:flux-tubes})
\begin{figure}[htb!]
\begin{center}
\resizebox*{5.8cm}{!}{\includegraphics{flux_tubes.eps}}
\hskip 1mm
\resizebox*{5.9cm}{!}{\includegraphics{phobos.eps}}
\end{center}
\caption{\baselineskip 15pt \label{fig:flux-tubes}\sl Left: Gauge
field configurations in the form of ``flux tubes'' of longitudinal
chromo-electric and chromo-magnetic fields screened on transverse
scales $1/Q_s$. Right: Model comparison~\cite{DusliGLV1} to long
range rapidity correlations measured by the PHOBOS
collaboration~\cite{Alvera1}.}
\end{figure}
as comprising $R_A^2 Q_s^2$ color flux tubes of size $1/Q_s$, each
producing $1/\alpha_s$ particles per unit rapidity. The flux tube
picture is supported by non-perturbative numerical solutions of the
classical Yang-Mills equations~\cite{LappiSV1}. The ``Glasma flux
tubes'' generate $n$-particle long range rapidity
correlations~\cite{DumitGMV1,factorization,DusliFV1,DusliGLV1}. These
distributions are negative binomial distributions~\cite{GelisLM1}.
They also carry topological charge~\cite{KharzKV1}; the resulting
dynamical topological ``sphaleron'' transitions may result in
observable metastable CP-violating domains~\cite{KharzMW1}.
The evolution of the Glasma into a thermalized Quark Gluon Plasma
(QGP) is not understood. An important ingredient is the role of
instabilities \cite{instabilities}. At early times, these arise at NLO
from terms that break the boost invariance of the LO
term~\cite{GelisLV2,factorization}. The modification to the evolution
of the Glasma is obtained by solving 3+1-D Yang-Mills
equations~\cite{insta1} for the (now) rapidity dependent gauge fields
convolved with a distribution giving the spectrum of
fluctuations~\cite{FukusGM1}. While these effects may isotropize the
system, early thermalization may also require collisions whose role
still needs to be clarified~\cite{collisions}.
\section{Phenomenological applications of the CGC}
In this section, we will discuss the applications of the theoretical
formalism outlined in the previous section to analyze and {\sl
predict} a wide range of phenomena ranging from DIS in e+p and e+A
collisions to the scattering of hadronic projectiles ranging from p+p
to p+A to A+A collisions. A unifying ingredient in many of the
applications is the dipole cross-section defined in
eq.~(\ref{eq:opt-thm-LO}), albeit, as apparent in the treatment of A+A
collisions, the fundamental ingredient is really the density matrix
$W_{Y}[\rho]$. Because the JIMWLK equation
(eq.~(\ref{eq:JIMWLK-diff})) for this quantity is time consuming to
solve\footnote{Other unknowns include higher order corrections,
initial conditions at low energy and impact--parameter dependence of
distributions.}, many of the applications are in the context of
models of the dipole cross-section which incorporate key features of
saturation. These models provide an economical description of a wide
range of data with only a few parameters. A good compromise between
the full JIMWLK dynamics and models of the dipole cross-section is the
BK equation, which is a large $N_c$ realization of JIMWLK dynamics.
With the recent availability of the NLO BK equation, global analyzes
of data are in order. Much of our discussion below is in the context
of dipole models; improvements {\it a la} BK are highlighted wherever
available.
\subsection{DIS in e+p and e+A collisions}
A remarkable observation~\cite{StastGK1} is that HERA
data~\cite{StastGK1,GelisPSS1} on the inclusive virtual photon-proton
cross section for $x\leq 0.01$ scale as a function of the ratio
$Q^2/Q_s^2(x)$; see the left part of figure \ref{fig:geom}. This
scaling is violated for larger values of $x$. Here $Q_s^2 = Q_0^2
(x_0/x)^\lambda$ is the saturation scale with $Q_0^2 = 1$ GeV$^2$,
$x_0 = 3\cdot 10^{-4}$ and $\lambda \approx 0.3$. This scaling is referred
to as ``geometrical scaling'', because the survival probability of the
color dipole that the virtual photon fluctuates into is close to unity
or zero respectively depending on whether the ratio of the saturation
radius ($\sim 1/Q_s$) to the size of the dipole (of size $\sim 1/Q$)
is large or small. Recall that the saturation radius denotes the
typical size of regions with strong color fields.
\begin{figure}[htb!]
\begin{center}
\resizebox*{!}{4.5cm}{\includegraphics{scaling_fix.eps}}
\hskip 3mm
\resizebox*{!}{4.5cm}{\includegraphics{rhic_au_bkfc_ext1.eps}}
\end{center}
\caption{\baselineskip 15pt \label{fig:geom}\sl Geometrical scaling and limiting
fragmentation. Left: $\sigma_{\gamma^*p}$ data at HERA for $x\le
0.01$ and all $Q^2$ up to 450~GeV$^2$; $\tau$ is the scaling
variable, $\tau\equiv Q^2/Q^2_s(x)$ \cite{StastGK1,GelisPSS1}.
Right: particle multiplicities for several collision energies at
RHIC \cite{Backa1}, compared to the computation of \cite{GelisSV1}.}
\end{figure}
Geometric scaling has also been observed in inclusive diffraction,
exclusive vector meson production and deeply virtual Compton
scattering data at HERA~\cite{MarquS1}. In detail, the data also show
{\em violations} of geometric scaling, which can be interpreted as
consequences of BFKL diffusion \cite{IancuIM3}, non--zero quark masses
\cite{Soyez1} and possibly DGLAP evolution as well~\cite{KowalMW1}.
Note that the best scaling is obtained with a saturation scale that
behaves like $Q_s^2(x)\propto x^{-0.3}$, a slower $x$-dependence than
predicted by the LO BK equation. This discrepancy is resolved by a
resummed NLO computation of the saturation exponent \cite{Trian1}
which indeed gives $0.3$.
While geometrical scaling is very suggestive of the presence of
semi-hard dynamical scales in the proton, it is not conclusive in and
of itself~\cite{CaolaF1}; more detailed comparisons to the data are
essential. Despite their simplicity, saturation
models~\cite{GBW,BarteGK1,IancuIM3,KowalT1,KowalMW1,Soyez1,forshaw,AlbacAMS1}
provide remarkably good descriptions of HERA data at small $x\le
0.01$. The free parameters are fixed from fits to the total
cross-section data alone; once these are fixed, the models {\sl
predict} a large variety of results, including the longitudinal
($F_{_L}$), diffractive ($F_2^{_D}$), and charm ($F_2^c$) structure
functions, the virtual photon production of vector mesons
($\rho,\,J/\psi$), and the deeply virtual Compton scattering (DVCS).
The most recent analysis of inclusive data~\cite{AlbacAMS1} of DIS in
e+p collisions is quite sophisticated; the energy dependence is given
by the running--coupling BK equation, and the free parameters refer
solely to the initial conditions and to the proton transverse area.
The phenomenon of {\em hard diffraction} in DIS is particularly
sensitive to saturation. The simplest diffractive processes are events
in which the proton remains intact and a large gap in rapidity with no
particles extends between the rapidity of the proton and that of the
fragmentation products of the virtual photon.
\begin{figure*}[htb!]
\begin{center}
\includegraphics[width=0.45\textwidth]{rdiffmar11.eps}
\hskip 3mm
\includegraphics[width=0.35\textwidth]{crossq_dvcs.eps}
\end{center}
\caption{\baselineskip 15pt \label{fig:diff}\sl Left: ZEUS data for the ratio
$\sigma_{\rm diff}/\sigma_{\rm tot}$ together with the
corresponding prediction of the saturation model in
\cite{BarteGK1}. Right: DVCS measurement at HERA, and
comparison with \cite{KowalMW1}.}
\end{figure*}
For small invariant masses, this process corresponds to elastic
scattering of the $q\bar q$ dipole off the target. Its cross--section
is evaluated as\footnote{Forward diffraction (corresponding to $t=0$,
where $t=(P-P^\prime)^2$ is the momentum transfer squared between
the incoming and outgoing proton) can be compared directly to
inclusive DIS within a dipole model since it depends only on the
dipole cross-section.}
\begin{equation}
\label{sigmadiff}
\left.\frac{\rmd\sigma_{\rm diff}}{\rmd t}\right|_{t=0}
=\frac{1}{16\pi}\int\limits_0^1 \rmd z\int \rmd^2{\boldsymbol r}_\perp
\left|\psi(z,{\boldsymbol r}_\perp)\right|^2
\sigma_{\rm dipole}^2(x,{\boldsymbol r}_\perp)\; .
\end{equation}
The dipole cross-section in this expression, for small dipoles $r^2
\ll 1/Q_s^2$, is a color singlet combination of two gluons that can be
interpreted as Pomeron exchange~\cite{pomeron}; for larger dipoles
$r^2\geq 1/Q_s^2$, the color singlet exchange does not have this
simple interpretation. The $t$ distribution has the form $d\sigma_{\rm
diff}/dt = \exp(-B_{_D} |t|) d\sigma_{\rm diff}/dt|_{t=0}$, where
$B_{_D}$ is the transverse area of the interaction region in the
proton and is closely related to the transverse gluon radius in the
proton estimated to be $0.61\pm 0.04$~fm~\cite{CaldwK1}. From this
form of the diffractive cross-section and eq.~(\ref{sigmadiff}), the
total diffractive cross-section is
\begin{eqnarray}\label{intdiff}
\sigma_{_D} \sim
\frac{B_{_D}}{Q^2}\int\limits_{Q^{-2}}^{Q_s^{-2}}
\frac{\rmd r^2 }{r^4}\, \Big(r^2Q_s^2(x)\Big)^2
\sim B_{_D}\,\frac{Q_s^2(x)}{Q^2}\; .
\end{eqnarray}
We used here the color transparency approximation $\sigma_{\rm
dipole}\propto r^2 Q_s^2(x)$ for the dipole cross-section. Unlike
the inclusive cross-section in eq.~(\ref{eq:X-DIS}) which is dominated
by small dipole sizes $\sim 1/Q$, the integrand of the diffractive
cross-section is dominated by larger size dipoles of size $\sim
1/Q_s$. Comparing eq.~(\ref{intdiff}) with eq.~(\ref{eq:X-DIS}), we
deduce that in the saturation framework the ratio $\sigma_{\rm
diff}/\sigma_{\rm tot}$ is approximately constant as a function of
energy. As shown in figure~\ref{fig:diff} (left), the HERA data
support this qualitative observation and are in quantitative agreement
with the detailed saturation model of~\cite{BarteGK1}. We note further
that excellent fits with $\chi^2 \sim 1$ are obtained in this
saturation framework for exclusive vector meson production and deeply
virtual Compton scattering~\cite{KowalMW1} in addition to inclusive
diffraction~\cite{KowalLMV1} (see figure \ref{fig:diff}, right). These
exclusive processes provide detailed information about the impact
parameter dependence of the dipole cross-section~\cite{impact} and may
even provide unique information about the partonic nature of short
range nuclear forces~\cite{CaldwK1}.
In contrast, if unitarization were due to soft physics (the prevailing
viewpoint before the advent of saturation~\cite{Bjork6}), the
natural cutoff for the integral would be $1/\Lambda_{_{\rm QCD}}$. Diffraction would
be non--perturbative even for hard $Q$ and the energy dependence of
the diffractive cross-section would be the square of the inclusive
cross-section, in disagreement with data.
Fits based on geometrical scaling of the e+A fixed target data with
the functional form $Q_s^2 \propto A^\delta$ (with $\delta$ naively
$1/3$) give values of $\delta$ with range from $1/4$ to
$4/9$~\cite{delta}. However, the dipole formalism which describes e+p
data so successfully can be straightforwardly generalized to nuclei to
construct the corresponding nuclear dipole
cross-section~\cite{KowalT1}. When compared to results for the $x,Q^2$
dependence of inclusive e+A data, one finds effectively that
$Q_{s,A}^2 (x)= Q_{s,p}^2 (x) A^{1/3}$~\cite{KowalLV1}. Nuclear
diffractive distributions can be computed in saturation
models~\cite{KowalLMV1,KugerGN1}; predictions have been made as well
for semi-inclusive hadron production~\cite{MarquXY1}, exclusive vector
meson production~\cite{vector} and nuclear DVCS~\cite{Macha1} in this
framework.
In saturation models of nuclei, the small $x$ distributions in a
nucleon are convolved with nuclear geometry to give the nuclear
distributions. This process however does not commute with the RG
evolution in $x$ of nuclear distributions determined at some initial
scale. First computations of e+A inclusive distributions in the NLO BK
framework have been performed and good agreement obtained for existing
fixed target data~\cite{DusliGLV1}.
\subsection{Particle multiplicities in d+A and A+A collisions}
The CGC EFT is most reliable when at least one of the projectiles is
dense in the sense discussed in the previous section. At RHIC, the
world's first deu\-te\-ron+hea\-vy nucleus (d+A) and A+A collider, many
features of the CGC are being tested. These include bulk features such
as the rapidity and centrality dependence of particle multiplicities
in d+A and A+A collisions, limiting fragmentation, particle spectra
and correlations, and even possibly more exotic features such as long
range rapidity correlations (``the ridge'') in A+A collisions and
local CP violation arising from topological fluctuations in A+A
collisions. In this sub-section, we will focus on particle
multiplicities and discuss the other features in subsequent
sub-sections.
``Limiting fragmentation'' is the well known property of the strong
interactions that the rapidity distribution in the fragmentation
region becomes independent of the collision energy. When one
increases the beam energy (see figure \ref{fig:geom}, right),
$dN/d\eta^\prime$ is the same at large $\eta^\prime$ for all energies,
where $\eta^\prime\equiv \eta-\eta_{\rm beam}$. In the fragmentation
region, large $x_1$ modes are probed in one hadron or nucleus and
small $x_2$ modes in the other. At small $x_2$, if gluonic matter is
saturated, parton distributions have a very weak dependence on $x_2$,
or equivalently on $\eta+\eta_{\rm beam}$; cross-sections will only
depend on $x_1$ or $\eta-\eta_{\rm beam}$. Deviations from limiting
fragmentation at high energies are especially interesting because of a
significant window in phase space where the RG evolution in $x_2$ to
(or away) from the universal ``black disk'' of saturated gluon matter
can be explored~\cite{Jalil3,GelisSV1}. These constraints lead to
specific predictions for the rapidity dependence of the multiplicity
at the LHC in the BK RG framework~\cite{GelisSV1,Albac1}.
The CGC can only predict the distribution of initial gluons, at a
proper time $\tau\sim Q_s^{-1}$. In p/d+A collisions this is not a
significant limitation because only a few gluons are produced in the
final state. The situation is vastly different in A+A collisions where
the particle multiplicity is significantly larger and the system
evolves from the non-equilibrium Glasma to a QGP, the latter evolving
subsequently as a fluid with low viscosity. For p+A collisions, the
problem is solvable analytically~\cite{KovchM3,DumitM1,BlaizGV1}. For
A+A collisions, two kinds of calculations have been performed:
(a) Exact numerical solutions of the Yang-Mills
equations~\cite{class0,KrasnV2,Lappi}. Quantum evolution effects are
not included systematically and the ensemble of color charges is
assumed to be Gaussian (MV model), which is reasonable at RHIC
energies of $x\sim 10^{-2}$ at RHIC. The inclusion of quantum effects
from the wavefunctions~\cite{factorization,GelisLM1} is under control,
but is still an unsolved problem for those present in the final state
evolution~\cite{insta1,FukusGM1}.
(b) Approximate analytical calculations of the initial gluon spectrum
\cite{KharzL1}. These calculations assume $k_\perp$-factorization
although it is violated for momenta $p_\perp \leq Q_s$. In this
model, saturation effects are introduced via the unintegrated gluon
distribution of the nuclei with the rapidity dependence of the gluon
spectrum governed by that of the saturation scale $Q_s^2(x)\sim
x^{-0.3}$. In the CGC framework, $dN/d\eta = Q_s^2
R_A^2/\alpha_s(Q_s)$; a unique feature is that the coupling runs as a
function of $Q_s$~\cite{Muell12}, an observation which is in agreement
with the centrality dependence of RHIC data~\cite{KharzN1}. Similar
analyzes of multiplicity distributions were extended to the case of
d+A collisions \cite{KharzLN2}, with results in fair agreement with
RHIC data.
\subsection{Inclusive spectra in d+A and A+A collisions}
\subsubsection{$p_\perp$ dependence}
One does not expect significant final state interactions with on-shell
partons in $p+A$ collisions. Moreover, the inclusive gluon spectrum
depends only on the Wilson line correlator
$\big<U(0)U^\dagger({\boldsymbol x}_\perp)\big>$~\cite{DumitM1,pA,BlaizGV1}, that
can be determined from the dipole cross-section used in
DIS~\cite{GelisJ3}. The hadron spectrum in d+A collisions has been
evaluated in this approach in \cite{DumitHJ1}, with results in good
agreement with RHIC data. Very recently, good agreement of single
inclusive distributions in d+A collisions were obtained in the NLO RG
approach~\cite{AlbacM1} consistent with its application in e+A
collisions~\cite{DusliGLV1}.
In A+A collisions, the strong final state interactions likely alter
the momentum distribution to be more isotropic at low momenta; the
hard tail is modified by parton energy loss. While these effects may
in part be included in the Glasma, the additional contributions at
later stages are important.
\subsubsection{Nuclear modification ratios}
To quantify nuclear effects, $p_\perp$ spectra in A+A and d+A
can be compared to those in p+p collisions by the ratios
\begin{eqnarray}
R_{AA}\equiv
\frac{\left.\frac{dN}{dyd^2{\boldsymbol p}_\perp}\right|_{AA}}
{N_{\rm coll}\left.\frac{dN}{dyd^2{\boldsymbol p}_\perp}\right|_{pp}}
\; ,\quad
R_{dA}\equiv
\frac{\left.\frac{dN}{dyd^2{\boldsymbol p}_\perp}\right|_{dA}}
{N_{\rm coll}\left.\frac{dN}{dyd^2{\boldsymbol p}_\perp}\right|_{pp}}\; ,
\end{eqnarray}
where $N_{\rm coll}\sim A^{4/3}$ (resp. $A$) is the number of binary
collisions in A+A (resp. d+A) collisions. At high $p_\perp$, one
expects the ratios to scale with $N_{\rm coll}$. In A+A collisions at
RHIC, because no suppression is seen in d+A collisions at central
rapidities, the large observed suppression in $R_{AA}$ is rightly
interpreted as a final state effect, with a strong candidate being
energy loss induced by the dense medium~\cite{Eloss}.
However, at small $x$, RG evolution {\it a la} BK
predicts~\cite{forward} that this ratio should be suppressed by
saturation effects in the wavefunction of the nucleus. This can be
studied in the clean environment of d+A collisions where small $x$ is
probed measuring the ratio at a positive rapidity in the direction of
the deuterium nucleus. The ratio $R_{dA}$ was measured by the
BRAHMS~\cite{Arsena1} collaboration up to rapidities of $\eta=3.2$ and
by the STAR collaboration~\cite{Adamsa6} at $\eta=4$. At these large
$\eta$, the value of $R_{dA}$ is significantly below unity, consistent
with predicted trend. In addition, as anticipated, the suppression is
greater for more central collisions. The onset of this forward
suppression was also studied semi-analytically in
\cite{BlaizGV1,IancuIT2}, and more quantitative calculations have been
performed in \cite{KKT,DumitHJ1,GoncaKMN1}, using models
of the dipole cross-section that have a realistic $x$-dependence and
most recently in the NLO BK framework~\cite{AlbacM1}.
\subsection{Two hadron correlations in d+A and A+A collisions}
\begin{figure}[htb!]
\begin{center}
\resizebox*{7.2cm}{!}{\includegraphics{horizon.eps}}
\resizebox*{4.8cm}{!}{\includegraphics{STAR_corr_data_sub.eps}}
\end{center}
\caption{\baselineskip 15pt \label{fig:2hadron}Left: causal relations
between two particles separated in rapidity. Right: 2-hadron
correlation measured by STAR as a function of $\Delta\eta$ and
$\Delta\phi$. }
\end{figure}
Two hadron correlations\footnote{Hadron-photon correlations have also
been studied in \cite{Jalil6}.} are more sensitive in distinguishing
between model predictions than the single inclusive results for which
alternative explanations of the data are feasible. Very recently,
striking preliminary results on these correlations in d+Au collisions
have been presented by the STAR and PHENIX collaborations. In the STAR
measurements, back-to-back correlations of pairs have been studied
where i) one particle in the pair is forward in rapidity and the other
at central rapidity and ii) both particles in the pair are at forward
rapidities. The ``forward-central'' (FC) pairs probe $x\sim 10^{-2}$
in the gold nucleus while the``forward-forward'' (FF) pairs probe
$x\sim 10^{-3}$ in the gold nucleus. A clear broadening of the
backward hadron peak is seen in the transition from FC to FF, as well
as with increasing centrality in the FF events--in the latter case,
the distribution is so broad that no peak is visible! This particular
effect was predicted in the CGC~\cite{Marqu1} and is in quantitative
agreement the STAR results~\cite{decorr}. These results are also
consistent with measurements by the PHENIX collaboration on pair
distributions in d+Au collisions. Related discussions can be found
in~\cite{pair}.
While two particle correlations in A+A collisions are typically much
altered by final state interactions, there is an important exception
for particles widely separated in rapidity. Causality implies that
these correlations are created at very early times, as illustrated in
the left panel of figure \ref{fig:2hadron}. A simple estimate for
ultra-relativistic particles whose space-time and momentum rapidities
are strongly correlated gives
\begin{equation}
\tau_{\rm max}=\tau_{\rm freezeout}\; e^{-\frac{|\Delta\eta|}{2}}\; ,
\end{equation}
for the latest time at which these particles could have been
correlated. For a freeze-out time $\tau_{\rm freezeout}\approx
10~$fm/c, and rapidity separations $\Delta\eta\geq 4$, one sees that
these correlations must have been generated before $1~$fm/c. Thus long
range rapidity correlations are a ``chronometer'' of the evolution of
the very strong Glasma color fields produced early in the heavy ion
collision.
Striking long range rapidity correlations were seen in A+A collisions
in events with prominent jet like structures, where the spectrum of
associated particles is observed to be collimated in the azimuthal
separation $\Delta \varphi$ relative to the jet and shows a nearly
constant amplitude in the strength of the pseudo-rapidity correlation
$\Delta \eta$ up to $\Delta \eta\sim 1.5$~\cite{Adamsa4}. These events
were coined ``ridge'' events following from the visual appearance of
these structures as an extended mountain ridge in the $\Delta
\eta$-$\Delta \varphi$ plane associated with a narrow jet peak. This
collimated correlation persists up to $\Delta \eta ~\sim
4$~\cite{Molna1,Alvera1}. An important feature of ridge correlations
is that the above described structure is seen in two particle
correlations without a jet trigger (see figure~\ref{fig:2hadron}) and
persists without significant modification for the triggered
events~\cite{Adamsa5,Abelea1}. These events include all hadrons with
momenta $p_\perp \geq 150$ MeV. In such events, a sharp rise in the
amplitude of the ridge is seen~\cite{Adamsa5,Daugh2} in going from
peripheral to central collisions.
The ridge structures seen in two particle and three particle~\cite{3p}
A+A correlations can be explained as resulting from the transverse
radial flow of the Glasma flux tubes we discussed previously in
section \ref{sec:AA}. The flux tubes are responsible for the long
range rapidity correlation in the ridge; the angular collimation
occurs because particles produced isotropically in a given flux tube
are collimated by the radial outward ``Hubble'' hydrodynamic flow of
the flux tubes. Ideas on the angular collimation of particle
distributions by flow were discussed in refs.~\cite{flow}. When
combined with the long range rapidity correlations provided by the
Glasma flux tubes, they provide a semi-quantitative description of the
ridge measurements~\cite{DumitGMV1,gavin,DusliGLV1,DusliFV1}. More
detailed studies are feasible with realistic hydrodynamic
simulations~\cite{hydro}. We note that the PHENIX collaboration which
also observed ridge-like structures~\cite{Adarea3} have presented
preliminary data showing that ridge-like structures disappear as
the $p_\perp$ of the trigger particle is increased~\cite{Chen1}; this
result is in qualitative agreement with expectations in the Glasma
flux tube picture. For a recent critical evaluation of this and
alternative models, we refer the reader to ~\cite{Nagle1}.
\section{Outlook: LHC and future DIS colliders}
At the LHC, very low values of $x$ will be probed in p+p, A+A
(including the photo-production induced dynamics of peripheral A+A
collisions) and possibly p+A collisions through a wide range of final
states and diverse kinematic ranges. In nuclei, $Q_s^2$ will be large;
estimates range from $\sim 2.6$-$4$ GeV$^2$ in central collisions to
$\sim 10$ GeV$^2$ for $y=\pm 3$ units. The picture of strongly
correlated albeit weakly coupled dynamics of glue in the CGC EFT
outlined here will be tested as never before. Detailed tests of BFKL
dynamics and possibly even the CGC are feasible in p+p collisions in
studies respectively of Mueller-Navelet~\cite{MuellN1} and forward
jets. Diffractive final states, while always challenging to
interpret, offer opportunities to understand deeply how saturated
gluons generate rapidity gaps. A strong test of CGC dynamics will
already be available in ``Day 1'' physics of bulk dynamics in A+A
collisions. More subtle tests of the RG flow of multi-parton
correlators are available in a variety of final states ranging from
quarkonia to electromagnetic probes to long range rapidity
correlations. In p+A collisions at the LHC, many of the patterns seen
in forward single inclusive and di-hadron correlation spectra will be
present already at central rapidities and will be much more dramatic
at forward rapidities. Finally, the role of Glasma dynamics
relative to that of the Quark-Gluon Plasma in A+A collisions at the
LHC is still unclear. If RHIC has reached the perfect hydrodynamic
limit of maximal flow, what happens at the LHC?
Studies of the dynamics of gluon saturation and the CGC that are
complementary to the LHC can be performed at a future Electron-Ion
Collider (EIC)\footnote{Current proposals include the EIC in the eRHIC
version at BNL and the eLIC version at Jlab, the LHeC proposal at
CERN and an electron-ion collider at FAIR in GSI.}. Firstly, with a
wide lever arm in $Q^2$, the dynamics of saturation can be studied
with precision in the regime where $Q^2\sim Q_s^2\gg \Lambda_{_{\rm
QCD}}^2$; this is difficult to achieve at a hadron collider.
Secondly, clarifying what aspects of the dynamics probed are universal
and novel calls for an electron probe. The factorization theorems we
have developed here suggest that the density matrices $W$ are
universal--how can we confirm this and cleanly extract their rich
dynamics? These issues are not merely academic and are strongly
reflected in the structure of final states. For example, rapidity
gaps are a large fraction of the cross-section at HERA but are a much
smaller contribution in hadronic collisions at Fermilab, demonstrating
a breakdown of factorization~\cite{AlverCTW1}. The physics case for
an EIC/LHeC has been outlined in \cite{EIC} and in a
number of white papers and reports.
In summary, the prospects for unambiguous discovery and exploration of
a novel many body QCD regime of gluon saturation are very bright in
the next decade. The theoretical status of studies of this regime
within the framework of the CGC EFT are increasingly robust albeit may
questions remain. It is hoped that experiments, as usual, will provide
definitive answers.
\section*{Acknowledgements}
F.G. and E.I. are supported in part by Agence Nationale de la
Recherche via the program ANR-06-BLAN-0285-01. J.J.-M. is supported by
the DOE Office of Nuclear Physics through Grant No.~DE-FG02-09ER41620
and by the City University of New York through the PSC-CUNY Research
Award Program, grant 62625-40. R.V.'s research is supported
by the US Department of Energy under DOE Contract No.
DE-AC02-98CH10886.
|
\section{Introduction}
Let $M$ be a compact manifold of dimension $n$ and $P = P(h)$ a semiclassical pseudodifferential operator on $M$ parametrised by the positive number $h \in (0, h_0]$. Suppose that $u = u(h)$ is an $O(h)$ quasimode, i.e. an $L^2$-normalised family of functions, defined for some subset of $(0, h_0]$ accumulating at $0$, such that $P(h) u(h)$ is $O(h)$ in $L^2(M)$. We assume $P$ has real principal symbol $p(x, \xi)$ and that its full symbol is smooth in $h$. We also put technical assumptions on $p(x,\xi)$ (see Definition \ref{admissible} and \ref{curvedtoflow}) and assume $u$ is localised (see Definition \ref{localised}). One important special case is when $P(h) = h^2 \Delta - 1$ where $\Delta$ is the Laplacian with respect to a Riemannian metric on $M$. Then $u(h)$ is an approximate eigenfunction with eigenvalue $h^{-2}$:
$$
(\Delta - h^{-2}) u(h) = O(h^{-1}) \text{ in } L^2(M).
$$
Other cases of interest are discussed in \cite{koch07}, where this framework was introduced.
The aim of this paper is to bound the extent to which $u$ can concentrate as $h \to 0$ by estimating the $L^{p}$ norm of $u$ restricted to hypersurfaces, in a manner that is sharp (up to a constant independent of $h$) as $h \to 0$. In particular, we wish to relate the degree of concentration to the geometry of the hypersurface relative to the bicharacteristic flow of $P(h)$.
There are a number of ways to study concentration of eigenfunctions. One can for example study semiclassical measures as in G\'{e}rard-Leichtnam \cite{gerard}, Zel\-ditch \cite{zelditch}, Zelditch-Zworski \cite{zelditch2}, Anantharaman \cite{anantharaman}, Anantharaman-Koch-Non\-nen\-macher \cite{anantharaman07a}, Anantharaman-Nonnenmacher \cite{anantharaman07}. The aim of these studies is generally to prove non-concentration theorems under geometric conditions on the geodesic flow (such as Anosov flow).
In 1988 Sogge \cite{sogge88} produced sharp $L^{p}$ estimates for spectral clusters (and therefore eigenfunctions) of elliptic operators, comparing the size of the $L^{p}$ norm over the full manifold to the $L^{2}$ norm in terms of powers of the eigenvalue $\lambda$. In 2004 Reznikov \cite{reznikov} proved bounds for restrictions of Laplacian eigenfunctions to curves where the underlying manifold is a hyperbolic surface and in 2007 Burq, G\'{e}rard and Tzvetkov \cite{burq07} proved estimates for general submanifolds and Laplacian eigenfunctions. Their estimates are sharp for sub-sequences of spherical harmonics. For high $p$ these estimates are optimised by eigenfunctions concentrating at a point. For low $p$ the optimising examples are eigenfunctions concentrating in a small tube around a stable periodic geodesic. Burq, G\'{e}rard and Tzvetkov \cite{burq07} were also able to obtain better estimates for small $p$ in dimension two when the submanifold is a curve with positive geodesic curvature. Hu \cite{hu09} extended this to hypersurfaces in $n$ dimensions where the hypersurface has positive curvature. In the special case of a flat two or three dimensional torus Bourgain and Rudnick obtain an improved nonconcentration result for curved hypersurfaces \cite{bourgain09}.
In 2009 Tacy \cite{tacy09} extended Burq, G\'{e}rard and Tzvetkov's results on Laplacian eigenfunctions to quasimodes of semiclassical operators. This extension uses the semiclassical framework set up in Burq-G\'{e}rard-Tzvetkov \cite{burq2} and Koch-Tatatru-Zworski \cite{koch07}. The main result of \cite{tacy09} is the following, where we refer to Definitions~\ref{localised} and \ref{admissible} for the precise definitions of localisation and admissibility.
\begin{thm}\label{tacytheorem}
Let $(M,g)$ be a smooth manifold without boundary and let $H$ be a smooth embedded hypersurface. Let $u(h)$ be a family of $L^{2}$ normalised functions that satisfy $Pu=O_{L^{2}}(h)$ for $P$ a semiclassical operator with symbol $p(x,\xi)$. Assume further that $u$ satisfies the localisation property and that the symbol $p(x,\xi)$ is admissible. Then
$$\norm{u}_{L^{p}(H)}\lesssim{}h^{-\delta(n,p)},$$
\begin{equation}\delta(n,p)=
\begin{cases}
\frac{n-1}{2}-\frac{n-1}{p},&\frac{2n}{n-1}\leq{}p\leq\infty\\
\frac{n-1}{4}-\frac{n-2}{2p},&2\leq{}p\leq\frac{2n}{n-1}
\end{cases}.
\label{delta(n,p)}\end{equation}
\end{thm}
\begin{remark} We have only given the results of \cite{tacy09} pertaining to hypersurfaces. Higher codimension submanifolds were
also treated there.
\end{remark}
\begin{figure}
\centering
\includegraphics[width=9cm,height=7cm]{hypersurfacegraph.eps
\label{fig:hypersurfacebetter}
\caption{$\delta(p)$ plotted against $1/p$ for a general hypersurface and for a hypersurface curved with respect to the flow.}
\end{figure}
This paper extends the estimates of Burq-G\'erard-Tzvetkov and Hu for curved hypersurfaces to the semiclassical regime, framing the geometric conditions in terms of the classical (bicharacteristic) flow.
To motivate the condition of curvature, recall that the classical flow defined by
\begin{equation}\begin{cases}
\dot{x}=\partial_{\xi}p(x,\xi)\\
\dot{\xi}=-\partial_{x}p(x,\xi)\end{cases}\label{classicalflow}\end{equation}
describes the movement in phase space of a classical particle with classical Hamiltonian $p(x,\xi)$. For the model case of the Laplacian the flow defined by (\ref{classicalflow}) is the geodesic flow. In the semiclassical regime we wish to find estimates that link the properties of this classical flow to concentrations of quasimodes. Intuitively we can think of highly localised packets moving on trajectories defined by the flow. The more time a packet spends near a hypersurface the move concentration we would expect to see there. In \cite{koch07} and \cite{tacy09} it is shown that for a hypersurface $H'$ with boundary defining function\footnote{We say that the real function $r$ is a boundary defining function for $H'$ if $H' = \{ r = 0 \}$ and if $r$ vanishes simply at $H'$, i.e. $dr \neq 0$ at $H'$.} $\bdf$, if at some point $(x_0, \xi_0)$ we have $\dot \bdf \neq 0$, where the dot indicates derivative with respect to bicharacteristic flow, and if $u$ is a quasimode sufficiently localized near $(x_0, \xi_0)$, then $u$ does not concentrate at $H'$. That is, if $\chi\in{}C_{0}^{\infty}(\ensuremath{\mathbb{R}}^{n}\times{}\ensuremath{\mathbb{R}}^{n})$ is a cut off function with small enough support around $(x_{0},\xi_{0})$ then
\begin{equation}
\norm{\chi(x,hD)u}_{L^{2}(H')}\lesssim{}\norm{u}_{L^{2}(M)}.
\label{nonconcentration}\end{equation}
However, in the general case, a bicharacteristic may stay inside $H$, allowing considerable concentration of an associated wave packet on $H$. As shown in \cite{koch07} and \cite{tacy09}, concentration (as measured by $L^2$ norm) could be as bad as $\sim h^{-1/2}$ assuming just the localisation condition and assumption (A1) below, while additionally assuming (A2) introduces dispersion effects which reduces the concentration to $\sim h^{-1/4}$. To improve on this,
we need to rule out bicharacteristics that stay inside $H$. A natural assumption to make is that the projections of bicharacteristics are only simply tangent to $H$. In local coordinates this is the same as saying the whenever a bicharacteristic is tangent to $H$, i.e. $\dot \bdf(x_0, \xi_0)$ vanishes, $x_0 \in H$, then the normal acceleration $\ddot{\bdf}(x_{0},\xi_{0})$ is nonzero. We phrase this by saying that $H$ is \emph{curved} with respect to the bicharacteristic flow.
Under this additional assumption, which we label (A3) below, we show that the concentration is at most $\sim h^{-1/6}$:
\begin{thm}\label{glancingflowtheorem}
Let $M$, $H$, $P(h)$ and $u(h)$ be as in Theorem~\ref{tacytheorem}.
If $H$ is curved with respect to the flow given by $p(x,\xi)$, i.e. satisfies assumption (A3) below, then the estimate \eqref{delta(n,p)} for $p=2$ can be improved from $\delta = 1/4$ to $\tilde \delta = 1/6$. By interpolation with the result for $p = 2n/(n-1)$, we obtain
\begin{equation}\begin{gathered} \norm{u}_{L^{p}(H)}\lesssim{}h^{-\tilde\delta(n,p)}, \quad 2\leq{}p\leq\frac{2n}{n-1}, \\
\tilde\delta(n,p)=
\frac{n-1}{3}-\frac{2n-3}{3p},
\end{gathered}\label{mainestimate}\end{equation}
under assumption (A3).
\end{thm}
\begin{remark}
For $p\geq{}2n/(n-1)$ there is no improvement in the curved case.
In this case the $\| \cdot \|_{L^p(H)}$ norm is maximised by functions that concentrate at points so we would not expect the geometry of the hypersurface to affect such estimates. \end{remark}
\section{Semiclassical Analysis}\label{semiclassical}
We work with semiclassical pseudodifferential operators (for a full introduction see \cite{burq2}, \cite{evans} or \cite{koch07}). Such operators are defined by their symbol $p(x,\xi,h)$ and a quantisation procedure
$$P(h) u(h) = p(x,hD, h)u(h)=\frac{1}{(2\pi{}h)^{n}}\int{}e^{\frac{i}{h}<x-y,\xi>}p(x,\xi,h)u(y,h)d\xi{}dy$$
where $h$ is a small parameter.
Because we are just about to assume that $u$ is localised (Definition \ref{localised}), it is harmless to assume that $p$ is a $C_c^\infty$ function of $(x, \xi)$, and for simplicity we take it to be smooth in $h \in [0, h_0]$.
By abuse of notation we denote the principal symbol $p(x, \xi, 0)$ by $p(x, \xi)$, and we will write $p(x, hD)$ for $p(x, hD, h)$.
Following \cite{koch07}, we assume that our family of quasimodes $p(x,hD)u(h)=O_{L^{2}}(h)$ is semiclassically localised:
\begin{defin}\label{localised}
A function $u$ depending parametrically on $h$ is localised if there exists $\chi\in{}C_{c}^{\infty}(T^{\star}M)$ such that
$$u=\chi(x,hD)u+O_{\mathcal{S}}(h^{\infty})$$
where $\mathcal{S}$ is the space of Schwartz functions, and $ g \in O_{\mathcal{S}}(h^{\infty})$ means that each seminorm of $g$ is $O(h^\infty)$.
\end{defin}
Localisation is compatible with the assumption that $p(x,hD)u=O_{L^{2}}(h)$: that is, if $\chi\in{}C_{c}^{\infty}(T^{\star}M)$ then
$$p(x,hD)u=O_{L^{2}}(h)\Rightarrow{}p(x,hD)(\chi(x,hD)u)=O_{L^{2}}(h).$$
Using this localisation assumption we are able to turn the global problem into a local problem on small patches in $T^{\star}M$. If $\chi\in{}C_{c}(T^{\star}M)$ such that
$$u=\chi(x,hD)u+O_{\mathcal{S}}(h^{\infty})$$ then, using compactness of the support of $\chi$, we can write
$$\chi(x,\xi)=\sum_{i=1}^{N}\chi_{i}(x,\xi)$$
for some $N<\infty$ where each $\chi_{i}$ has arbitrarily small support. In this fashion we reduce estimating $\norm{\chi(x,hD)u}_{L^{p}(H)}$ to estimates on $\norm{\chi_{i}(x,hD)u}_{L^{p}(H)}$ (the error term $O_{\mathcal{S}}(h^{\infty})$ is of course trivial to estimate). Due to this localisation we can replace $M$ with $\ensuremath{\mathbb{R}}^{n}$, $H$ with $\ensuremath{\mathbb{R}}^{n-1}$ and $T^{\star}M$ with $\ensuremath{\mathbb{R}}^{n}\times{}\ensuremath{\mathbb{R}}^{n}$. We write $x\in{}M$ as $x=(y,\bdf)$ where $y\in{}H$ and $\bdf$ is the normal direction to $H$.
Still following \cite{koch07}, we further reduce this problem to localising around points $(x_{0},\xi_{0})$ where $p(x_{0},\xi_{0})=0$. To achieve this we use Lemma 2.1 of \cite{koch07} which shows that if $|p(x,\xi)|\geq{1/C}$ on a local patch then we can invert $p(x,hD)$ up to order $h^{\infty}$. That is, choosing $\chi(x,\xi)$ supported on this patch, we can find some $q(x,hD)$ such that
$$q(x,hD)p(x,hD)\chi(x,hD)=\chi(x,hD)+O_{L^{2}\rightarrow{}L^{2}}(h^{\infty})$$
and
$$p(x,hD)q(x,hD)\chi(x,hD)=\chi(x,hD)+O_{L^{2}\rightarrow{}L^{2}}(h^{\infty}).$$
So if $p(x,hD)u=O_{L^{2}}(h)$ and $|p(x,\xi)|>1/C$ we can invert $p(x,hD)$ to get
$$\chi(x,hD)u=O_{L^{2}}(h).$$
We can combine this estimate with the following `semiclassical Sobolev inequality' (see \cite{burq2}, \cite{evans} or \cite{koch07} for proof) to obtain hypersurface restriction estimates.
\begin{lem}[semiclassical Sobolev estimates]\label{semiclassicalsobolev}
Suppose that a family $u=u(h)$ satisfies the localisation condition. Then for $1\leq{}q\leq{}p\leq\infty$
$$\norm{u}_{L^{p}}\lesssim{}h^{n(1/p-1/q)}\norm{u}_{L^{q}}+O(h^{\infty}).$$
\end{lem}
To get the $L^{2}$ norm of the restriction of $u$ to $H$ we use Lemma \ref{semiclassicalsobolev} in only the $\bdf$ coordinates. This is justified as localisation in $T^{\star}\ensuremath{\mathbb{R}}^{n}$ implies localisation in $T^{\star}\ensuremath{\mathbb{R}}^{n-1}$ (see \cite{tacy09}). We have
\begin{equation}\norm{u(y,0)}_{L^{2}_{y}}\lesssim\norm{u(y,\bdf)}_{L^{\infty}_{\bdf}L^{2}_{y}}\lesssim{}h^{-\frac{1}{2}}\norm{u(y,z)}_{L^{2}_{z}L^{2}_{y}}.\label{goodL2}\end{equation}
So, if $|p(x,\xi)|\geq{1/C}$, and $Pu = O(h)$, the $L^{2}$ norm of $u$ when restricted to a hypersurface $H$ containing $x_0$ is $O(h^{\frac{1}{2}})$. This is significantly better than the $L^{2}$ estimate given by Theorem \ref{glancingflowtheorem}. Consequently we can ignore regions where $p(x,\xi)$ is bounded away from zero.
To get better estimates when $p(x_0, \xi_0) = 0$ than what can be obtained from Lemma~\ref{semiclassicalsobolev} (which uses only localisation), we need to make assumptions on the function $p$ (to prevent $p$ vanishing identically, for example, in which case the assumption $Pu = O(h)$ is vacuous!). Our first assumption (A1) is that $p$ vanishes simply on each cotangent fibre:
\begin{itemize}
\item[(A1)]
\text{ for any point $(x_{0},\xi_{0})$ such that $p(x_{0},\xi_{0})=0$, $\partial_{\xi}p(x_{0},\xi_{0})\neq{}0$.}
\end{itemize}
Our second condition is a geometric condition on the characteristic variety. The condition eliminates examples such as $p(x, \xi) = \xi_1$, i.e. $P = h D_{x_1}$, for which we cannot estimate $\| u \|_{L^2(H)}$ by better than the $h^{-1/2}$ estimate given by Lemma~\ref{semiclassicalsobolev} alone. Let us note that (A1) implies that the set
\begin{equation}
\{\xi\mid{}p(x_{0},\xi)=0\}\subset{}T_{x_{0}}^{\star}M
\label{charvar}\end{equation}
is a smooth hypersurface in $T_{x_{0}}^{\star}M$.
\begin{itemize}
\item[(A2)]
\text{For each $x_0 \in M$, the second fundamental form of \eqref{charvar} is positive definite.}
\end{itemize}
\begin{defin}\label{admissible}
A symbol $p(x,\xi)$ is \emph{admissible} if it satisfies condition (A1) and (A2). \end{defin}
In addition we make the geometric assumption of curvature with respect to the flow.
\begin{defin}\label{curvedtoflow}
A hypersurface $H$ of $M$ is \emph{curved} with respect to the flow if the projection of the bicharacteristic flow to $M$ is at most simply tangent to $H$, or in other words, if for one (and hence any) boundary defining function $\bdf$ for $H$, we have
\end{defin}
\begin{itemize}
\item[(A3)] \text{ For any $(x_0, \xi_0)$, $\dot \bdf(x_0, \xi_0) = 0$ implies that $\ddot \bdf (x_0, \xi_0) \neq 0$.}
\end{itemize}
\begin{remark} In the case $P(h) = h^2 \Delta - 1$, where $\Delta$ is the Laplacian on $M$ with respect to a Riemannian metric, assumptions (A1) and (A2) are satisfied, and (A3) is satisfied iff $H$ has positive definite second fundamental form. Thus, in this case our result reduces to that of Burq-G\'erard-Tzvetkov \cite{burq07} ($n=2$) and Hu \cite{hu09} ($n \geq 2$).
\end{remark}
\section{Evolution equation}\label{symbolfac}
Using the argument in the previous section we can assume that $p(x_0, \xi_0) = 0$.
Assumption (A1) then tells us that $\partial_{\xi}p(x_{0},\xi_{0})\neq{}0$. Let us choose coordinates $x = (y, \bdf)$ where $y \in \ensuremath{\mathbb{R}}^{n-1}$ and $\bdf \in \ensuremath{\mathbb{R}}$ is a boundary defining function for $H$. Let $\xi = (\eta, \nu)$ be the dual coordinates. If $\partial_\nu p(x_0, \xi_0) \neq 0$ then we have $\dot r \neq 0$ and, as mentioned in the Introduction
(see \eqref{nonconcentration}), $u$ does not concentrate at $H$ at all.
Therefore we may assume that $\partial_\nu p(x_0, \xi_0) = 0$.
Therefore we have $\partial_\eta p(x_0, \xi_0) \neq 0$. By a linear change of $y$ coordinates we can assume that $\partial_{\eta_1} p(x_0, \xi_0) \neq 0$ and $\partial_{\eta_j} p(x_0, \xi_0) = 0$ for $j \geq 2$.
Now we apply the implicit function theorem and deduce that the characteristic variety $\{ p = 0 \}$ implicitly defines $\xi_1$ as a smooth function of $(x, \xi_2, \dots, \xi_n)$:
\begin{equation}
p = 0 \implies \xi_1 = a(x, \xi_2, \dots, \xi_n).
\label{adefn}\end{equation}
We shall now write $x_1 = t$ and think of it as a time variable. We write $x = (t, \xbar)$ and similarly, $\xi_1 = \tau$ and $\xi = (\tau, \xibar)$. We also write $y = (t, y')$ and $\eta = (\tau, \eta')$. Thus $x = (t, y', r)$ and correspondingly $\xi = (\tau, \eta', \nu)$.
We write $T$ for the `initial' hypersurface $ \{ t = 0 \}$, and recall that $H = \{ r = 0 \}$. We assume that $t = 0$ at $(x_0, \xi_0)$ and write
$(x_0, \xi_0) = ((0, \xbar_0), \xi_0) = ((0, y_0', 0), (\tau_0, \eta'_0, \nu_0))$.
As a consequence of \eqref{adefn}, we have
$$p = e(x, \xi) \big(\tau - a(x, \xibar)\big)$$
near $(x_0, \xi_0)$, where $e(x_0, \xi_0) \neq 0$. By localising suitably we may assume that $e \neq 0$ on the support of our localising function $\chi$. The condition $Pu = O(h)$ in $L^2$ then implies that
$$
e(x, hD_x) \big(h D_t - a(x, hD_{\xbar}) \big) u = O_{L^2(M)}(h)
$$
and using the local invertibility modulo $O(h^\infty)$ of $e(x, hD_x)$, we find that
\begin{equation}\big(hD_{t}-a(x,hD_{\xbar}) \big)u=hf(t,\xbar)
\label{evolutionequation}\end{equation}
where $\norm{f}_{L^{2}(M)}=O_{L^{2}}(1)$.
We view \eqref{evolutionequation} as an evolution equation for $u$, which determines $u$ given the `initial data' $u(0, \xbar)$ and the
inhomogeneous term $f(t, \xbar)$. This determines a family of solution operators $U_{s}(t)$, such that $U_{s}(t)$ is the solution operator for the evolution equation
$$
\big(hD_{t}-a(s+t,\xbar,hD_{\xbar}) \big)u=0, \quad u(0, \xbar ) = u(\xbar)
$$
Using Duhamel's principle we write
$$u(t,\bar{x})=U_{0}(t)u(0,\bar{x})+i\int_{0}^{t}U_{s}(t-s)f(s,\bar{x})ds.$$
Now let $R_H$ be the operation of restriction to the hypersurface $H$, and let $W_{s}(t) = R_H \circ U_{s}(t)$. Also, let $u_{0}=u(0,\xbar)$ be the restriction of $u$ to the initial hypersurface $T = \{ t = 0 \}$.
We then have
$$u(t,y',0)= W_{0}(t)u_0 +i\int_{0}^{t}W_{s}(t-s)f(s,\xbar)ds.$$
Using Minkowski's inequality we have
\begin{multline}
\norm{u}_{L^{2}(H)}\lesssim\left(\int\norm{W_{0}(t)u_{0}}_{L^{2}_{y'}}^{2}dt\right)^{1/2}+\\
\int_{\ensuremath{\mathbb{R}}}\left(\int\norm{W_{s}(t-s)f(s,\xbar)}_{L^{2}_{y'}}^{2}dt\right)^{1/2}ds
\label{qnorm}\end{multline}
We recall from \eqref{nonconcentration} (with $H' = T$) that $\| u_0 \|_{L^2(T)} \lesssim \| u \|_{L^2(M)}$.
Therefore, to prove Theorem~\ref{glancingflowtheorem}, i.e. obtain a $L^{2}$ bound of
$$\norm{u}_{L^{2}(H)}\lesssim{}h^{-1/6}\norm{u}_{L^{2}(M)}$$
it suffices to obtain an estimate, uniform in $s$, of the form
\begin{equation}\left(\int\norm{W_{s}(t-s) f}_{L_{y'}^{2}}^{2}dt\right)^{1/2}\lesssim{}h^{-1/6}\norm{f}_{L^{2}(T)}.\label{Wbounds}\end{equation}
For each $s$ we will show that \eqref{Wbounds} holds with a constant that depends only on the seminorms of $a(x,\xibar)$.
In fact, the estimates are uniform given uniform bounds on a finite number of derivatives of $a$, and given uniform lower bounds on the nondegeneracies involved in the computation in Section~\ref{canonical} --- see Remark~\ref{comech}. Such uniform bounds hold provided that the patch size is chosen sufficiently small. Therefore we only address the estimate for $W_{0}(t)$, which we denote by $W(t)$ from here on.
To obtain this estimate we view $W(t)$, thought of as a single operator from $L^2(T)$ to $L^2(H)$ instead of as a family parametrised by $t$, as an Fourier integral operator.
\section{Fourier integral representation}\label{oscillatoryevol}
We need to express the solution operator for the evolution equation
\begin{equation}hD_{t}-a(t,\bar{x},hD_{\bar{x}})=0\label{evolution}\end{equation}
as an Fourier integral operator. We will then be able to transfer properties of the flow to properties of the phase function defining the operator $U(t)$.
\begin{prop}\label{FIO}
Suppose $U(t):L^{2}(\ensuremath{\mathbb{R}}^{d})\rightarrow{}L^{2}(\ensuremath{\mathbb{R}}^{d})$ satisfies
$$hD_{t}U(t)-A(t)U(t)=0,\quad{}U(0)=Id$$
where A(t) is a semiclassical pseudodifferential operator such that the symbol $a(t,\bar{x},\eta)$ of $A(t)$ is real and is smooth in $h$. Then there exists some $t_{0}>0$ independent of $h$ such that for $0\leq{}t\leq{}t_{0}$
$$U(t)u(\bar{x})=\frac{1}{(2\pi{}h)^{d}}\int\int{}e^{\frac{i}{h}(\phi(t,\bar{x},\xibar)-\wbar\cdot\xibar)}b(t,\bar{x},\xibar,h)u(\wbar)d\wbar d\xibar+E(t)u(\bar{x})$$
where
$$\partial_{t}\phi(t,\bar{x},\xibar)-a(t,\bar{x},\partial_{\bar{x}}\phi(t,\bar{x},\xibar))=0,\quad{}\phi(0,\bar{x},\xibar)=\bar{x}\cdot\xibar$$
$$b(t,\bar{x},\xibar,h)\in{}C^{\infty}_{c}(\ensuremath{\mathbb{R}}\times{}T^{\star}\ensuremath{\mathbb{R}}^{d}\times{}\ensuremath{\mathbb{R}})\quad{}E(t)=O(h^{\infty}):S'\rightarrow{}S$$
\end{prop}
\begin{proof}
This is in fact the normal parametrix construction yielding the eikonal equation for the phase function. See \cite{evans} Section 10.2.
\end{proof}
Recall that $W(t)=R_{H}\circ{}U(t)$ so we have
$$
W(t)f(y')=\frac{1}{(2\pi{}h)^{n-1}}\iint{}e^{\frac{i}{h}(\phi(t,(y',0),\xibar)-\wbar\cdot\xibar)}b(t,y',\eta,h)f(\wbar)d\wbar d\xibar
$$
In what follows we will write $\phi(t,y', \eta', \nu)$ for $\phi(t,(y', 0),\xibar)$ (recall that $\xibar = (\eta', \nu)$). We want to estimate the operator norm of $W(t)$ regarded as a single operator acting from $L^2(T)$ to $L^2(H)$. Note that $W(t)=Z \circ \mathcal{F}_h$ where $\mathcal{F}_h$ is the semiclassical Fourier transform:
$$\mathcal{F}_h f(\xibar)=\frac{1}{(2\pi{}h)^{\frac{n-1}{2}}}\int{}e^{-\frac{i}{h}\xibar\cdot{}\vbar}f(\vbar)d\vbar$$
and the operator $Z$ is given by
$$
Z g(t, y') = \frac{1}{(2\pi{}h)^{\frac{n-1}{2}}}\iint{}e^{\frac{i}{h}\phi(t, y', \eta', \nu)}b(t,y',\eta', \nu,h) g(\eta', \nu) \, d\eta' \, d\nu.
$$
As $\norm{\mathcal{F}_{h}f}_{L^{2}}=\norm{f}_{L^{2}}$ it is enough to estimates $L^{2}\to{}L^{2}$ operator norm of $Z$. To estimate the operator norm of $Z$ we view it as a semiclassical Fourier integral operator and analyse its canonical relation.
\section{Canonical relation}\label{canonical}
To prove Theorem~\ref{glancingflowtheorem} we need to show that the operator norm of $Z$ is bounded by $Ch^{-1/6}$. To do this we use the following theorem of Pan and Sogge \cite{pan90} which is the analogue for oscillatory integral operators of Melrose and Talyor's \cite{melrose85} theorem on Fourier integral operators with folding canonical relations.
\begin{thm}\label{foldingrelations}
Let the oscillatory integral operator $T_{\lambda}$ be defined by
$$T_{\lambda}f(x)=\int_{\ensuremath{\mathbb{R}}^{d}}e^{i\lambda\psi(x,y)}\beta(x,y)f(y)dy$$
where $\beta\in{}C_{0}^{\infty}(\ensuremath{\mathbb{R}}^{d}\times{}\ensuremath{\mathbb{R}}^{d})$ and the phase function $\psi\in{}C^{\infty}(\ensuremath{\mathbb{R}}^{d}\times{}\ensuremath{\mathbb{R}}^{d})$ is real. If the left and right projections from the associated canonical relation
$$\canon_{\psi}=\{(x,\psi'_{x}(x,y),y,-\psi'_{y}(x,y))\}$$
are at most folding singularities then
$$\norm{T_{\lambda}f}_{L^{2}(\ensuremath{\mathbb{R}}^{d})}\lesssim{}\lambda^{-\frac{d}{2}+1/6}\norm{f}_{L^{2}(\ensuremath{\mathbb{R}}^{d})}$$
\end{thm}
Let us recall (see for example \cite{guillemingolubitsky}) that a smooth map $F : \ensuremath{\mathbb{R}}^d \to \ensuremath{\mathbb{R}}^d$ has a folding singularity
at $x \in \ensuremath{\mathbb{R}}^d$ if
\begin{itemize}
\item[(i)] $dF(x)$ is rank $d-1$,
\item[(ii)] the function $\det dF$ vanishes simply at $x$, implying in particular that locally near $x$, the set of $y \in \ensuremath{\mathbb{R}}^d$ such that $dF(y)$ has rank $d-1$ is a smooth hypersurface $S$ containing $x$, and \item[(iii)] the kernel of $dF(x)$ is not contained in the tangent space to $S$:
$$
T_x S + \ker dF(x) = T_x \ensuremath{\mathbb{R}}^d.
$$
\end{itemize}
Given (i) an equivalent condition to (ii) and (iii) is that, if $v$ is a nonzero element of $\ker dF(x)$, then
\begin{equation}
D_v(\det dF(x)) \neq 0.
\label{iii}\end{equation}
The operator $Z$ is a Fourier integral operator with canonical relation
$$
\canon = \big\{ (t, y', \partial_t \phi, \partial_{y'} \phi, \eta', \nu, -\partial_{\eta'} \phi, -\partial_{\nu} \phi ) \big\}.
$$
The left and right projections on $\canon$ are represented in local coordinates by
$$
\pi_L : (t, y', \eta', \nu) \mapsto (t, y', \partial_t \phi, \partial_{y'} \phi)
$$
and
$$
\pi_R : (t, y', \eta', \nu) \mapsto (\eta', \nu, \partial_{\eta'} \phi, \partial_{\nu} \phi )
$$
(where we removed the irrelevant minus signs from $\pi_R$ for notational convenience).
The matrix $d\pi_L$ takes the form
$$d\pi_{L}=\left(\begin{BMAT}(e)[2pt,1.5cm,1.5cm]{c.c}{c.c}
\Id & 0 \\
* & B \end{BMAT}\right)
$$
where
$$
B=\left(\begin{BMAT}(e){ccc.c}{c.ct}
&\partial^{2}_{t\eta'}\phi& & \partial^{2}_{t\nu}\phi \\
&&&\\
&\partial^{2}_{y'\eta'}\phi& & \partial^{2}_{y'\nu}\phi \end{BMAT}\right)
$$
At $(x_0, \xi_0)$ we have $\partial^2_{y' \eta'} \phi = \Id$, $\partial^2_{t\eta'} \phi = \partial_{\eta'} a = 0$, $\partial^2_{y' \nu} \phi = 0$ and $\partial^2_{t \nu} \phi = \partial_{\nu} a = 0$, so we get
$$
B =\left(\begin{BMAT}(e){ccc.c}{c.ct}
&0& & 0 \\
&&&\\
&\Id& & 0 \end{BMAT}\right).
$$
It is clear that the vector field $\partial_\nu$ is in the kernel of $d\pi_L(x_0, \xi_0)$. Moreover, $\det d\pi_L$ is given by $\partial^2_{t\nu} \phi \cdot \det (\partial^2_{y' \eta'} \phi)$ plus terms vanishing to second order at $(x_0, \xi_0)$. To show that $\pi_L$ has a fold at $(x_0, \xi_0)$ we need by \eqref{iii} to show that $\partial_\nu (\det d\pi_L)$ is nonzero at $(x_0, \xi_0)$.
Due to the vanishing of both `off-diagonal' terms $\partial^2_{t\eta'} \phi$ and $\partial^2_{y' \nu} \phi$, the nonvanishing of $\partial_\nu (\det d\pi_L)$ at
$(x_0, \xi_0)$ is equivalent to the nonvanishing of $\partial_{\nu} (\partial^2_{t\nu} \phi) = \partial^3_{t\nu \nu} \phi$.
The matrix $d\pi_R$ takes the form
$$
d\pi_R=\left(\begin{BMAT}(e)[2pt,1.5cm,1.5cm]{c.c}{c.c}
0 & \Id \\
D & *
\end{BMAT}\right)$$
$$
D=\left(\begin{BMAT}(e){c.ccc}{bc.c}
\partial^{2}_{\eta't}\phi& & \partial^{2}_{y'\eta'}\phi &\\
&&&\\
\partial^{2}_{\nu{}t}\phi& &\partial^{2}_{\nu{}y'}\phi &\end{BMAT}\right)
$$
and we see that $\partial_t$ is in the kernel of $d\pi_R(x_0, \xi_0)$.
To show that $\pi_R$ has a fold at $(x_0, \xi_0)$ we need by \eqref{iii} to show that $\partial_t (\det d\pi_L)$ is nonzero at $(x_0, \xi_0)$.
As above, due to the vanishing of the `off-diagonal' terms $\partial^2_{t\eta'} \phi$ and $\partial^2_{y' \nu} \phi$, the nonvanishing of $\partial_t (\det d\pi_L)$ at
$(x_0, \xi_0)$ is equivalent to the nonvanishing of $\partial_t (\partial^2_{t\nu} \phi) = \partial^3_{t t \nu} \phi$.
The proof of Theorem~\ref{glancingflowtheorem} is therefore completed by the following Lemma:
\begin{lem} Under assumptions (A1), (A2) and (A3), we have
$$
\partial^3_{t \nu \nu} \phi(x_0, \xi_0) \neq 0, \text{ and } \partial^3_{t t \nu} \phi(x_0, \xi_0) \neq 0.
$$
\end{lem}
\begin{remark} To simplify notation we write $(x_0, \xi_0)$ for the argument of $\phi$ corresponding to this point, although $(0, y_0',0,\tau_{0}, \eta_0', \nu_{0})$ would be more accurate.
\end{remark}
\begin{proof}
We use the Hamilton-Jacobi equation
\begin{equation}
\partial_{t}\phi(t, \xbar, \eta', \nu) = a\big(t, \xbar,\partial_{\xbar}\phi(t, \xbar,\eta', \nu)\big).
\label{HJ}\end{equation}
Since at $t=0$ we have $\phi(0, \xbar, \eta', \nu) = y' \cdot \eta' + \bdf \nu$ (recall that $\xbar = (y', \bdf)$), we have
$$
\partial^3_{t \nu \nu} \phi(0,\xbar,\eta', \nu) = \partial^2_{\nu \nu} a(0, \xbar, \eta', \nu).
$$
Now we apply assumption (A2): it says that the second fundamental form of the submanifold $\{ \tau = a(x, \xibar_0) \} \subset T_{x_0} M$ is positive definite. Since $\partial_{\xibar} a = 0$ at $(x_0, \xibar_0)$, the second fundamental form of this submanifold at $(x_0, \xi_0)$ is given by the matrix of second derivatives of $a$:
$$
h_{ij}(\xi_0) = \partial^2_{\xi_i \xi_j} a(x_0, \xibar_0), \quad 2 \leq i, j \leq n.
$$
Therefore, $\partial^2_{\nu \nu} a \neq 0$ at $(x_0, \xibar_0)$, showing that $\pi_L$ has a fold singularity at $(x_0, \xi_0)$.
To treat the term $\partial^3_{t t \nu} \phi(x_0, \xi_0)$, we differentiate \eqref{HJ} in $t$, obtaining
$$
\partial^2_{tt} \phi = \partial_t a + \partial_{\xibar} a\cdot \partial^2_{\xbar t} \phi.
$$
Using \eqref{HJ} again on the term $\partial^2_{\xbar t} \phi$ we obtain
$$
\partial^2_{tt} \phi = \partial_t a + \partial_{\xibar} a \cdot\big( \partial_{\xbar} a + \partial_{\xibar} a \cdot\partial^2_{\xbar \xbar} \phi \big).
$$
We evaluate this at $t=0$ since the next derivative to be applied, namely $\partial_\nu$, is tangent to $\{ t = 0 \}$. At $t=0$, we have $\partial^2_{\xbar \xbar} \phi = 0$, so we get
$$
\partial^2_{tt} \phi \Big|_{t=0} = \partial_t a + \partial_{\xibar} a \cdot \partial_{\xbar} a.
$$
Now when we differentiate in $\nu$, we get
$$
\partial^3_{tt\nu} \phi(x_0, \xi_0) = \partial^2_{t \nu} a(x_0, \xibar_0) + \partial^2_{\xibar \nu} a(x_0, \xibar_0)\cdot \partial_{\xbar} a(x_0, \xibar_0)
$$
since $\partial_{\xibar} a(x_0, \xibar_0) = 0$.
At this point we remind the reader that we have chosen coordinate $(t,y',\bdf)$ and $(\tau,\eta',\nu)=(\tau,\xibar)$ such that
$$\partial_{\xibar}p(x_{0},\xi_{0})=0$$
and
$$\tau_{0}-a(t_{0},\xbar_{0},\xibar_{0})=0.$$
It follows that
\begin{equation}\begin{cases}
\partial_{\xi}a(x_{0},\xibar_{0})=0\\
\partial_{\tau}p(x_{0}.\xi_{0})=e(x_{0},\xi_{0})\\
\partial_{\bar{x}}p(x_{0},\xi_{0})=-e(x_{0},\xi_{0})\partial_{\bar{x}}a(x_{0},\xibar_{0})\\
\partial_{t}p(x_{0},\xi_{0})=-e(x_{0},\xi_{0})\partial_{t}a(x_{0},\xi_{0}).
\end{cases}
\label{miscid}\end{equation}
Now we apply assumption (A3), which says that $\ddot \bdf \neq 0$. We express $\ddot \bdf$ in terms of $a$. We have
$$
\dot \bdf = \partial_\nu p = \partial_\nu \big(e (\tau - a) \big).
$$
Differentiating a second time and using the flow identities
$$\dot{x}=\partial_{\xi}p(x,\xi)\quad\dot{\xi}=-\partial_{x}p(x,\xi),$$
we have
$$
\ddot \bdf = \Big( \partial_\tau p \partial_t + \partial_{\xibar} p \partial_{\xbar} - \partial_t p \partial_\tau - \partial_{\xbar} p \partial_{\xibar} \Big) \Big( (\tau - a)\partial_\nu e - e \partial_\nu a \Big).
$$
At $(x_{0},\xi_{0})$ using the identities given in \eqref{miscid} we can simplify this to
$$
\ddot \bdf(x_0, \xi_0) = -e \Big( \partial^2_{\nu t} a(x_0, \xibar_0) + \partial_{\xbar} a(x_0, \xibar_0)\cdot \partial^2_{\nu \xibar} a(x_0, \xibar_0) \Big).
$$
Therefore, applying assumption (A3), we find
$$
\partial^3_{tt\nu} \phi(x_0, \xi_0) = - \frac{\ddot \bdf(x_0, \xi_0)}{e(x_0, \xi_0)} \neq 0.
$$
This shows that $\pi_R$ has a fold singularity at $(x_0, \xi_0)$ and completes the proof.
\end{proof}
\begin{remark} It is easy to see from the calculations above that
assumption (A3) is \emph{equivalent} to the statement that $\pi_R$ has a folding singularity. Similarly,
assumption (A2) is equivalent to the statement that $\pi_L$ has a folding singularity for every hypersurface $H$ whose tangent space $T_{x_0} H$ at $x_0$ contains $\partial_\xi p(x_0, \xi_0) \partial_{x}$,
i.e. the tangent vector of the projected bicharacteristic through $(x_0, \xi_0)$.
\end{remark}
\begin{remark}\label{comech}
According to \cite{comech}, Theorem 2.2, one obtains uniform bounds of the form $C h^{-1/6}$ on the norms of the operators $W_s$ given by \eqref{Wbounds} provided that there are uniform bounds on a finite number of derivatives of the symbol of $W_s$, and uniform lower bounds on the determinant of $\partial^2_{y' \eta'} \phi$, $\partial_\nu (\partial^2_{t \nu} \phi)$, and $\partial_t (\partial^2_{t \nu} \phi)$. These lower bounds are achieved simply by shrinking the patch size sufficiently and using continuity. Thus we obtain a bound as in \eqref{Wbounds} uniformly in $s$, as desired.
\end{remark}
\section{Optimality of Theorem~\ref{glancingflowtheorem}}
All the estimates given by Thereom \ref{glancingflowtheorem} are sharp. We study a simple local model around $(0,0)$ for hypersurface curved with respect to the flow. Let $H=\{x\mid{}x_{n}=0\}$ and $p(x,\xi)$ be given by
$$p(x,\xi)=\xi_{1}-x_{n}-\sum_{i=2}^{n}\xi_{i}^{2}$$
Note that
\begin{align*} \dot{t}&=1&\dot{y'}&=-2\eta'&\dot{r}&=-2{}\nu
\\
\dot{\tau}&=0&\dot{\eta'}&=0&\dot{\nu}&=1
\end{align*}
Therefore the flow $(x(s),\xi(s))$ with intial point $(0,0)$ is given by
\begin{align*}
t(s)&=s&{}y'(s)&=0&r(s)&=-s^{2}\\
\tau(s)&=0&\eta'(s)&=0&\nu(s)&=s
\end{align*}
So we have that condition (A3) is clearly satified as $\ddot{r}(0)=-2$.
$$p(x,hD)=hD_{t}-r-h^{2}D_{r}^{2}-\sum_{i=1}^{n-2}h^{2}D_{y'}^{2}$$
It is easier to develop a solution in Fourier space. Note that
$$
\mathcal{F}_{h} \circ p(x,hD) \circ \mathcal{F}_{h}^{-1} =\tau-hD_{\nu}-\nu^{2}-\eta'\cdot\eta'.
$$
As the semiclassical Fourier transform preserves $L^{2}$ norms if
$$\norm{(\tau-hD_{\nu}-\nu^{2}-\eta'\cdot\eta')f}_{L^{2}}=O_{L^{2}}(h)$$
and $u=\mathcal{F}_{h}^{-1}f$, then
$$\norm{p(x,hD)u}_{L^{2}}=O_{L^{2}}(h).$$
We therefore seek a solution for
\begin{equation}(\tau-hD_{\nu} - \nu^2 - \eta'\cdot\eta')f=0;\label{Fdiffeq}\end{equation}
it is obvious that
$$g(\tau,\eta',\nu)=e^{\frac{i}{h}(\frac{1}{3}\nu^{3}+\nu(\tau-\eta'\cdot\eta'))}$$
is a solution to \eqref{Fdiffeq}. The natural scaling $\nu\to{}h^{-1/3}\nu$ induces a scaling of $\tau\to{}h^{-2/3}\tau$ and $\eta'\to{}h^{-1/3}\eta'$ and accordingly we place cut off functions appropriate to that scale. Let
$$f(\tau,\eta',\nu)=h^{-\frac{n-2}{6}-\frac{1}{3}}\chi(|\nu|)\chi(h^{-\frac{2}{3}}|\tau|)\chi(h^{-\frac{1}{3}}|\eta'|)e^{\frac{i}{h}\psi(\tau,\eta',\nu)}$$
where
$$\psi(\tau,\eta',\nu)=\frac{1}{3}\nu^{3}+\nu(\tau-\eta'\cdot\eta')$$
Now $\norm{f}_{L^{2}}=O_{L^{2}}(1)$ and $f$ satisfies \eqref{Fdiffeq} up to an $O(h)$ error coming from the $D_\nu$ hitting the cutoff function $\chi(|\nu|)$. We define the function $u$ as
$$u=\chi(|x|)\mathcal{F}_{h}^{-1}f$$
Now $R_{H}u$ is given by
$$R_{H}u(y)=\frac{h^{-\frac{n-2}{6}-\frac{1}{3}}\chi(|y|)}{(2\pi{}h)^{\frac{n}{2}}}\int{}e^{\frac{i}{h}(t\tau+y' \cdot \eta'+\psi(\tau,\eta',\nu))}\chi(|\nu|)\chi \left( \frac{|\tau|}{h^{2/3}} \right) \chi \left( \frac{|\eta'|}{h^{1/3}} \right)d\tau{}d\nu{}d\eta' .$$
For $|t|\leq{} \epsilon h^{1/3}$, $\epsilon$ small, the factor $e^{\frac{i}{h}t\tau}$ does not oscillate significantly and can be ignored. Similarly for $|y'|\leq{} \epsilon h^{2/3}$ the factor $e^{\frac{i}{h}y' \cdot \eta'}$ does not oscillate significantly and is also ignored. On the other hand, there are oscillations in the $\nu$ variable. At $\tau = \eta' = 0$ there is degenerate stationary phase at $\nu = 0$; Theorem 7.7.18 of \cite{Hormandervol1} applies and shows that there is a lower bound of the form
$$|R_{H}u(t, y')| \sim {}h^{-\frac{n-2}{6}-\frac{1}{3}-\frac{n}{2}+\frac{n-1}{3}+\frac{2}{3}}=h^{-\frac{n-1}{3}}$$
for $|t|\leq{} \epsilon h^{1/3}$, $|y'|\leq{} \epsilon h^{2/3}$. Thus on this set we get a lower bound on the $L^p$ norm:
$$\norm{u}_{L^{p}([0,\epsilon h^{1/3}]_t\times{}B(0,\epsilon h^{2/3})_{y'})}\sim{}h^{-\frac{n-1}{3}+\frac{1}{3p}+\frac{2(n-2)}{3p}}=h^{-\left(\frac{n-1}{3}-\frac{2n-3}{3p}\right)}$$
which saturates the estimate of Theorem \ref{glancingflowtheorem}.
|
\section{INTRODUCTION}
\label{sintro}
Supernova 1986J was one of the most radio luminous supernovae ever
observed. Its long-lasting and strong radio emission and its relative
nearness makes it one of the few supernovae for which it is possible
to produce detailed images with very-long-baseline interferometry
(VLBI), in addition to being one of the few supernovae still
detectable more than $t = 20$ years after the explosion.
SN~1986J was first discovered in the radio, some time after the
explosion \citep{vGorkom+1986, Rupen+1987}. The best estimate
of the explosion epoch is $1983.2 \pm 1.1$ \citep[$t = 0$, Bietenholz
et~al. 2002, see also][]{Rupen+1987, Chevalier1987,
WeilerPS1990}\nocite{SN86J-1}. It occurred in the nearby galaxy
NGC~891, whose distance is $\sim 10$~Mpc \citep[see
e.g.,][]{Tonry+2001, Ferrarese+2000, Tully1988, Kraan-Korteweg1986,
Aaronson+1982}. We will adopt the round value of 10~Mpc throughout
this paper.
Early optical spectra showed that the supernova was unusual, but based
on its prominent H$\alpha$ lines it was classified as a Type~IIn
supernova \citep{Rupen+1987}. Later, \citet{Leibundgut+1991}
suggested that it might be of Type Ib.
VLBI observations started soon after discovery \citep{Bartel+1987},
and an image, the first of any optically identified supernova, was
obtained soon after \citep{Bartel+1991}, and showed a heavily
modulated shell structure with protrusions. VLBI observations at
subsequent epochs up to 1999 led to a series of images showing the
expansion of the shell, and allowing the expansion velocity and
deceleration to be measured \citep{SN86J-1}.
The evolution of the radio spectrum of a supernova is in most cases
predictable. In particular, once supernovae have become optically thin
after a few months or years, they usually display power-law
radio spectra with spectral indices, $\alpha$, in the range of $-0.8$ to
$-0.5 \; (S \propto \nu^\alpha)$. SN~1986J also showed such a
spectrum before 1998, but a remarkable change occurred after that. An
inversion appeared in the spectrum, with the brightness increasing
with increasing frequency above $\sim$10~GHz, and a high-frequency
turnover at $\sim$20~GHz \citep{SN86J-1}.
This spectral inversion was associated with a bright spot in the
(projected) center of the expanding shell, as we showed in
\citet{SN86J-Sci} using phase-referenced multi-frequency VLBI imaging.
At that time (late 2002) the bright spot was clearly present in the
15~GHz image, but not discernible in the 5~GHz one. We report
here on new VLBI and VLA observations of SN~1986J which show the
further evolution of this unique object.
\section{Observations and Data Reduction}
\label{sobs}
\subsection{VLBI Measurements}
\label{svlbiobs}
\begin{deluxetable*}{l c c c c }
\tabletypesize{\scriptsize}
\tablecolumns{5}
\tablecaption{VLBI Observations of SN~1986J}
\tablehead{
~~~Date & Frequency & \colhead{Antennas\tablenotemark{a}}
& Total time & Recording rate \\
& (GHz)& & (hr) & (Mbits~s$^{-1}$)
}
\startdata
2005 Apr 25 & 22 & VLBA, Ef, Gb, Y27 & 12 & 256 \\
2005 Oct 24 & \phn5 & VLBA, Ef, Gb, Y27, Jb, On, Wb, Tr & 12 & 256 \\
2006 Dec 3 & \phn8 & VLBA, Ef, Gb, Y27 & 15 & 512 \\
2006 Dec 10 & 22 & VLBA, Ef, Gb, Y27 & 15 & 512 \\
2008 Oct 26 & \phn5 & VLBA, Ef, GB, Y27, Jb, Mc, Nt, Tr, Wb & 18 & 512 \\
\enddata
\tablenotetext{a}{
VLBA, ten 25~m dishes of the NRAO Very Long Baseline Array;\phn
Ef, 100~m, MPIfR, Effelsberg, Germany;\phn
Gb, 105~m, NRAO, Green Bank, WV, USA;\phn
Y27, equivalent diameter 130~m, NRAO, near Socorro, NM, USA;\phn
Jb, 76~m, Jodrell Bank, UK;\phn
Mc, 32~m, IdR-CNR, Medicina, Italy;\phn
Nt, 32~m, IdR-CNR, Noto, Italy;\phn
On, 20~m, Onsala Space Observatory, Sweden;\phn
Tr, 32~m, Torun, Poland;\phn
Wb, equivalent diameter 94~m, Westerbork, the Netherlands.\phn}
\label{tobs}
\end{deluxetable*}
The VLBI observations of SN~1986J were made between 2005 and 2008 with
global VLBI arrays of 13 to 18 antennas and with durations of 12 to 15
hours. The observations of SN~1986J were interleaved with ones of
3C~66A, only 40\arcmin\ away on the sky, which we used as a
phase-reference source. Details of the four observing runs are given
in Table~\ref{tobs}. The declination of +42\arcdeg\ of SN~1986J
enabled us to obtain dense and only moderately elliptical
\mbox{$u$-$v$}~coverage. As usual, a hydrogen maser was used as a time and
frequency standard at each telescope. The data were recorded with
either the VLBA or the MKIV VLBI system with sampling rates of 256 or
512~Mbits per second. We observed at frequencies of 5.0, 8.4 and
22~GHz. The data were correlated with the NRAO\footnote{The National
Radio Astronomy Observatory, NRAO, is a facility of the National
Science Foundation operated under cooperative agreement by Associated
Universities, Inc.}
VLBA processor in Socorro, NM (USA), and the analysis was carried out
with NRAO's Astronomical Image Processing System (AIPS). The initial
flux density calibration was done through measurements of the system
temperature at each telescope, and improved through self-calibration
of the 3C~66A data.
We phase-referenced to 3C~66A for all of the present observing
runs. Our positions in this paper are given relative to an assumed
position of RA = \Ra{02}{22}{39}{611500}, decl.\ =
\dec{43}{02}{07}{79884} taken from the International Celestial Reference
Frame, ICRF \citep{Fey+2004}.
for the brightness peak of 3C~66A.
\subsection{VLA Observations}
\label{svla}
In addition to the VLBI observation, we also obtained VLA observations
to measure SN~1986J's total flux density at a range of different
frequencies, both during our VLBI observations and at other times. We
also reduced a number of data sets from the VLA archives in order to
obtain additional flux density measurements. The epochs and the
frequencies observed are listed in Table~\ref{tflux}. These
observations allowed us to monitor the evolution of SN~1986J's
integrated radio spectrum.
The VLA data were reduced following standard procedures, with the flux
density scale calibrated by using observations of the standard flux
density calibrators (3C~286 and 3C~48) on the scale of
\citet{Baars+1977}. An atmospheric opacity correction using mean
zenith opacities was applied, and NGC~891/SN~1986J was self-calibrated
in phase to the extent permitted by the signal-to-noise ratio for each
epoch and frequency. At lower resolutions the supernova cannot be
reliably separated from the diffuse emission from the galaxy.
Therefore, we cite only flux densities obtained at FWHM
resolutions\footnote{We take the geometric mean of the major and minor
axes of the elliptical Gaussian fit to the dirty beam as our measure
of resolution.}
of $<7$\arcsec. For the 0.33~GHz measurements we incorporate an
additional uncertainty to reflect the difficulty of accurate galaxy
subtraction.
We note also that the gradual introduction of EVLA antennas into the
VLA array might cause some problems in the flux-density calibration.
Since the bandpass of the EVLA antennas differs from that of the older
VLA-antennas, non-closing offsets are introduced for continuum
observations such as ours, with the gain of an EVLA-EVLA baseline
being typically several percent higher than that of a VLA-VLA
baseline. Such non-closing offsets cannot be calibrated out using
only standard procedures, which assume that the gains are dependent
only on the antennas, and not on the baseline. These closure errors
might compromise the accuracy of the flux density calibration. To
determine whether this effect is significant, we carried out, for
several epochs and frequencies, a more elaborate flux-density
calibration scheme using baseline-dependent factors calculated by the
AIPS task BLCAL\@. Although the more elaborate calibration led to an
improvement in the dynamic range, in no case was the normal
flux-density scale in error by more than 1\%. Since we include in all
cases a flux-density calibration uncertainty of $\ge 5$\%, we can say
that any errors introduced by the use of the EVLA antennas are well
within our stated uncertainties.
\begin{deluxetable*}{l c c c c c c c }
\tabletypesize{\scriptsize}
\tablecaption{Flux Densities of SN~1986J from VLA Observations}
\tablehead{
\colhead{Date} & \multicolumn{6}{c}{Flux densities\tablenotemark{a}}\\
& \colhead{0.32~GHz} & \colhead{1.43 GHz}& \colhead{4.9 GHz\tablenotemark{b}}
& \colhead{8.4~GHz\tablenotemark{c}}
& \colhead{14.9 GHz} & \colhead{22 GHz\tablenotemark{d}}
& \colhead{43.3~GHz}
}
\startdata
1995 Jun 22\tablenotemark{e}
& &$25.6 \pm1.3 $&$14.2 \pm 0.8$& &$9.1 \pm 0.7$&$8.0 \pm 0.8$\\
1996 Oct 25\tablenotemark{e} & & &$11.0 \pm 0.6$&$8.7 \pm0.5$\\
1997 Jan 31\tablenotemark{e} & & & &$8.2 \pm0.5$\\
1998 Feb 9\tablenotemark{e} & & & & &$5.65\pm0.69$&$9.3 \pm 1.5$\\
1998 Jun 5 & &$13.7\pm0.9$\tablenotemark{f}
&$ 8.5 \pm 0.5$&$7.2 \pm 0.4$\\
1999 Feb 2 & & &$ 7.3 \pm 1.5$&$6.1 \pm 0.4$&$6.1 \pm 0.1$&$9.5 \pm 1.0$\\
1999 Jun 13\tablenotemark{e} &
&$11.69\pm0.62$& &$5.68\pm0.30$&$4.92\pm0.52$& \\
2000 Oct 19\tablenotemark{e} &
& $8.97\pm0.47$&$5.43\pm0.29$&$4.76\pm0.26$&$5.88\pm0.42$&$5.67\pm0.65$ \\
2001 Jan 25\tablenotemark{e} &
& $9.02\pm0.22$& &$4.75\pm0.36$\\
2002 Jan 27\tablenotemark{e}
& & $7.68\pm0.46$&$4.15\pm0.26$&$4.13\pm0.28$&$5.02\pm0.54$&$4.22\pm0.44$ \\
2002 May 25 & & $7.10\pm0.70$&$4.20\pm0.30$&$3.80\pm0.30$&$4.70\pm0.50$
&$5.00\pm0.60$&$3.9^{+2.1}_{-0.9}$ \\
2002 Nov 11 & & &$3.86\pm0.28$&$4.16\pm0.25$& &$5.37\pm0.50$\\
2003 Jan 8 & & &$3.75\pm0.45$&$3.75\pm0.30$&$4.25\pm0.44$&$4.41\pm0.48$&$3.90\pm0.50$\\
2003 Jun 22 & & $6.63\pm0.57$&$3.50\pm0.25$&$3.88\pm0.20$&$4.57\pm0.70$& \\
2004 Jan 31\tablenotemark{e}
& & & & &$4.72\pm0.29$&$4.74\pm0.27$\\
2004 Sep 23 & & $4.55\pm0.33$&$3.07\pm0.20$&$3.45\pm0.20$&$4.44\pm0.26$&$4.79\pm0.49$&$3.59\pm0.40$\\
2005 Jan 8 & & $4.49\pm0.33$&$2.91\pm0.16$&$3.21\pm0.19$&$4.01\pm0.27$&$4.11\pm0.23$&$3.18\pm0.38$\\
2005 Apr 26 & & &$2.58\pm0.43$&$2.92\pm0.17$&$3.14\pm1.12$&$3.73\pm0.28$\\
2005 Jun 13\tablenotemark{e}
& & &$2.45\pm0.17$&$3.11\pm0.19$&$4.01\pm0.34$&$3.75\pm0.59$\\
2005 Oct 10 & & & & & &$4.25\pm0.55$\\
2006 Mar 6 & & $3.99\pm0.26$&$2.62\pm0.17$&$3.20\pm0.16$&$3.92\pm0.26$&$3.93\pm0.20$&$2.91\pm0.35$\\
2006 Jun 6 & & &$2.45\pm0.15$&$3.20\pm0.17$&$4.16\pm0.35$&$3.61\pm0.21$&$2.32\pm0.34$\\
2006 Aug 30\tablenotemark{e} & & &$2.12\pm0.13$& & &$3.24\pm0.43$\\
2006 Dec 4 & & & &$3.13\pm0.31$&\\
2006 Dec 11 & & & & & &$4.54\pm0.91$\\
2007 Apr 23 & & & & &$3.69\pm0.29$&$3.57\pm0.20$&$2.07\pm0.29$\\
2007 Jul 31 & & &$2.06\pm0.15$& &$3.77\pm0.32$\\
2007 Aug 19\tablenotemark{g}
&$8.9\pm1.8$& $3.07\pm0.13$& &$2.97\pm0.12$& \\
2007 Aug 23\tablenotemark{h} & & & & & &$3.12\pm0.19$&$2.13\pm0.23$ \\
2008 Oct 26 & & &$2.27\pm0.12$& \\
\enddata
\tablenotetext{a}{Flux densities from fitting images and/or fitting
\mbox{$u$-$v$}~plane models. The tabulated uncertainties include a systematic
uncertainty in the flux density calibration of 5\% or more at
frequencies of 22~GHz and below, and 10\% or more at 43~GHz.
They also include the estimated uncertainty of the galaxy
subtraction for those epochs/frequencies at which the galaxy
contribution was significant. We only tabulate measurements
at resolutions $<7$\arcsec, and ones where the galaxy contribution
was $<50$\% of the brightness of SN~1986J.}
\tablenotetext{b}{Observations between 4.86 and 4.99 GHz.}
\tablenotetext{c}{Observations between 8.41 and 8.46 GHz.}
\tablenotetext{d}{Observations between 22.21 and 22.66 GHz.}
\tablenotetext{e}{Data from the VLA archives.}
\tablenotetext{f}{Flux density at 1.7 GHz, rather than 1.43 GHz.}
\tablenotetext{g}{Values are the average from two observing runs on 2007 Jul 31 and 2007 Sep 6.}
\tablenotetext{h}{Values are the average from two observing runs on 2007 Jul 31 and 2007 Sep 15.}
\label{tflux}
\end{deluxetable*}
In Table~\ref{tflux} we list the total flux density measurements that
we obtained from VLA observations. Note that some of these values
have been previously published \citep[see][]{SN86J-COSPAR-2, SN86J-COSPAR,
SN86J-Sci}, but we repeat them here for the convenience of the reader
and to allow a better interpretation of the evolution of SN~1986J's radio
emission.
We plot a selection of these flux densities in two ways: first, in
Figure~\ref{flightcurve}, we plot the evolution of the flux-density as
a function of time at several different frequencies. Second, in
Figure~\ref{fspectra} we plot the radio spectrum as a function of time
at a number of epochs. We focus in this paper on the flux-density
evolution in the last decade, for lightcurves covering the earlier
part of SN~1986J's evolution, see \citet{SN86J-1} and
\citet{WeilerPS1990}.
\begin{figure*}
\centering
\includegraphics[height=0.67\textheight]{sn86j_fluxevolbw.eps}
\caption{Multi-frequency radio lightcurves for SN~1986J, as determined
from VLA observations. We plot the lightcurves at 1.4, 5.0, 8.4, 15,
and 22~GHz. The flux-density scale on the left axis is relevant
directly for the lightcurves at 1.5 and 5.0~GHz (marked with an asterisk),
while those at 8.4, 15 and 22 GHz are shifted logarithmically
downwards for better visibility, so that the shape of the lightcurve
is preserved even though the flux density scale is not. The curves
are shifted by the following factors: 8.4~GHz by 0.56, 15~GHz by 0.32,
and 22~GHz by 0.14. The uncertainties are estimated standard
errors, including statistical and systematic contributions. The data
include both our own and re-reduced archival data (see
Table~\ref{tflux}).}
\label{flightcurve}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[height=0.67\textheight]{sn86j_multispec6bw.eps}
\caption{The evolving radio spectrum of SN~1986J, as determined from
VLA observations. Each curve shows the radio spectrum at the epoch
indicated at left, with the earliest spectrum at the top. The
flux-density scale on the left axis is relevant directly for the first
three curves from the top (1995.5, 1999.1 and 2002.4), whose epochs
are marked with an asterisk. The other curves are shifted
logarithmically progressively downwards for better visibility, so that
the spectral shape is preserved even though the flux density scale is
not. The curves are shifted by the following factors: 2004.7 by 0.80,
2005.0 by 0.60, 2005.3 by 0.50, 2006.2 by 0.25, 2006.4
by 0.15, and 2007.6 by 0.10. The uncertainties are estimated
standard errors, including statistical and systematic contributions.
The data include both our own and re-reduced archival data (see
Table~\ref{tflux}).}
\label{fspectra}
\end{figure*}
Figure \ref{flightcurve} shows that the flux density at all our
observing frequencies is generally decreasing. In terms of a
power-law decay with $S_\nu \propto t^\beta$,
a weighted least-squares fit gives an average value of $\beta = -2.2$
over the last decade and over the frequencies of 1.5, 5.0, 8.4, 15 and
22~GHz, with the fastest decay seen at 1.5~GHz with $\beta = -3.3$
and the slowest at 15~GHz with $\beta = -1.0$.
\pagebreak[4]
~
\subsection{Radio Spectral Index}
\label{sspix}
As is clear from Figure~\ref{fspectra}, the radio spectral index,
$\alpha$ is a strong function of both time and frequency. Up to
1995.5, however, the spectrum at least between 1.4 and 15~GHz seems
well described by a single power-law. Earlier on, up to 1988,
\citet{WeilerPS1990} reported a spectral index for the optically-thin
part of $-0.67_{-0.04}^{+0.08}$. \citet{BallK1995}, for the period up
to 1991, find a somewhat flatter value of $-0.57 \pm 0.01$ from
fitting a diffusive acceleration model.
Fitting a power-law to the first set of flux densities included in
Table~\ref{tobs}, namely those of 1995.5, we find a notably flatter
spectrum, with a spectral index of $-0.44 \pm 0.03$ (we note that if
we drop the 22~GHz flux density from this fit on the suspicion that it
might already be contaminated by the inversion, the fitted spectral
index changes by less than the uncertainties). So the spectrum seems
to have flattened significantly between 1988 and 1995.
After 1995, the inversion appears in the spectrum, and its evolution
becomes more complex. We will focus first on the low-frequency part
of the spectrum, which we associate with the shell. In
\citet{SN86J-1}, we performed a two-part decomposition of the spectrum
for the period 1998 to 2002, and obtained a spectral index of
$-0.55_{-0.16}^{+0.09}$.
For even more recent times, we can calculate some representative
spectral indices from the flux densities in Table~\ref{tobs}.
From our measurements on 2007 Aug 19, we find $\alpha_{\rm 0.3
GHz}^{\rm 1.4 GHz} = -0.71 \pm 0.14$. Although the uncertainties are
rather large, this suggests also that the part of the spectrum at the
lowest frequencies is again becoming steeper since 1995.5.
In summary, the evidence suggests that the part of the spectrum below
the inversion, which we associate with the shell, was fairly steep
with $\alpha \simeq -0.62$ at times till 1988, flattening to $\alpha
\simeq -0.44$ in 1995, and once again steepening somewhat to $\alpha
\simeq -0.7$ since then. By averaging the spectral indices derived by
us and others for the low-frequency/optically-thin part of the
spectrum, we find an average value over the whole timerange of $-0.61$
with an rms of 0.11.
Another region of interest is the spectrum {\em above} the
high-frequency turnover. This part of the spectrum is dominated by
emission from the central bright spot, and thus presumably represents
its unabsorbed spectrum. A weighted average between 2006 Mar 6 and
2007 Aug 23 gives $\alpha_{\rm 22 GHz}^{\rm 44 GHz} = -0.61 \pm 0.11$.
\section{VLBI Images}
\label{simages}
In Figures~\ref{fimage5G} and \ref{fimage22} we show the VLBI images of
SN~1986J at the different frequencies and dates in Table~\ref{tobs}.
For comparison, we include in Figure~\ref{fimage5G} also some of our
previously published images from 1999 and 2002.
We discuss the images in turn, first those at 5~GHz, which
together with the earlier images present the best picture of the
evolution of the supernova, and then those at 8.4 and 22~GHz.
Since all the observations in this paper were phase-referenced to the
brightness peak of 3C~66A, which is expected to be closely related to
the core of the galaxy, we can accurately align the images of SN~1986J
at different times and frequencies to facilitate inter-comparison.
Unfortunately, since the earliest epochs of VLBI observations of
SN~1986J \citep[between 1987 and 1992, see][]{SN86J-1} were not
phase-referenced, we cannot pinpoint the explosion location
accurately, as we did for SN~1993J \citep[see][]{SN93J-1}. However,
we can visually estimate the position of the center of the shell
sufficiently accurately for our purposes. We use this center position
as the origin of our coordinate system in all the images, and also as
our estimate of the explosion location.
In particular, we estimated by eye the coordinates of the center of
the shell in each of the four phase-referenced, 5-GHz images shown in
Figure~\ref{fimage5G}. We then averaged these four sets of coordinates
and use the average to represent the coordinates of the shell center
and the explosion center position, which is Ra =
\Ra{02}{22}{31}{321457}, decl.\ = \dec{42}{19}{57}{25951}. We note
that a minimum uncertainty in this center position can be derived from
the scatter of the individual estimates, and is $\sim 60 \; \mu$as and
$\sim 180 \; \mu$as in RA and decl.\ respectively, and we estimate
realistic uncertainties in the actual position of the center of the
shell the order of 200~\mbox{$\mu$as}. We note that the largest component of
this uncertainty is not astrometric, but comes from the difficulty of
identifying the center position of the shell in the morphology, and
our measurements of proper motion below do not depend on this estimate
of the position of the shell's center.
\begin{figure*}
\centering
\includegraphics[height=0.8\textheight]{sn86j-5ghz-4panel.eps}
\caption{\small A sequence of VLBI images of SN~1986J at 5~GHz,
showing its evolution over the last decade. In each panel, the
coordinate system is centered on the estimated position of the center
of the shell (\Ra{02}{22}{31}{321457}, \dec{42}{19}{57}{25951}, see
text). The contours are in \% of the peak brightness, with negative
ones being dotted. Each panel has contours at 10, 20, 30, 40, {\bf
50}, 70, and 90\%, with the 50\% one being emphasized and the lowest
white one. The lowest contours, given for each panel below, are three
times the rms background brightness. The FWHM size of the convolving beam
is indicated at lower left in each panel. North is up and east is to
the left. This figure is also available as an mpeg animation in the
electronic edition of the {\em Astrophysical Journal}.
a) 1999 Feb 21, 5 GHz. The peak brightness was 1720~\mbox{$\mu$Jy~beam$^{-1}$}, the
lowest contour 5.5\%, the rms background brightness 31~\mbox{$\mu$Jy~beam$^{-1}$},
and the convolving beam $2.09 \times 1.02$~mas at p.a.\ $-7$\arcdeg.
b) 2002 Nov 10, 5 GHz. The peak brightness was 543~\mbox{$\mu$Jy~beam$^{-1}$}, the
lowest contour 7.6\%, the rms background brightness 14~\mbox{$\mu$Jy~beam$^{-1}$},
and the convolving beam $1.57 \times 1.12$~mas at p.a.\ $-10$\arcdeg.
The visibility data were tapered slightly in $u$ to increase
signal-to-noise and produce a somewhat less elongated beam.
c) 2005 Oct.\ 24, 5.0 GHz. The peak brightness was 414~\mbox{$\mu$Jy~beam$^{-1}$}, the
lowest contour 9.3\%, the rms background brightness 13~\mbox{$\mu$Jy~beam$^{-1}$},
and the convolving beam $1.65 \times 0.83$~mas at p.a.\
$-20$\arcdeg.
d) 2008 Oct.\ 25, 5.0 GHz. The peak brightness was 388~\mbox{$\mu$Jy~beam$^{-1}$}, the
lowest contour 7\%, the rms background brightness 9.0~\mbox{$\mu$Jy~beam$^{-1}$}, and the
convolving beam was $2.02 \times 0.98$~mas at p.a.\ $-13$\arcdeg.}
\label{fimage5G}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[width=0.473\textwidth]{sn86j-dec06x-l.eps}
\includegraphics[width=0.49\textwidth]{sn86j-dec06k-l.eps}
\caption{VLBI images of SN~1986J at 8.4 and 22~GHz.
In each panel, the coordinate system is centered on the estimated
position of the center of the shell (\Ra{02}{22}{31}{321457},
\dec{42}{19}{57}{25951}, see text). The contours are in \% of the
peak brightness, with the lowest being at three times the rms
background brightness, and the 50\% being emphasized and being the
first white contour. The FWHM of the convolving beam is shown at
lower left. North is up and east to the left.
a) Contours are drawn at $-6$, 6, 10, 20, 30, {\bf 50}, 70, and
90\% of the peak brightness, which was 717~\mbox{$\mu$Jy~beam$^{-1}$}. The rms background
brightness was 15~\mbox{$\mu$Jy~beam$^{-1}$}, and the convolving beam FWHM was $1.24 \times
0.68$~mas at p.a.\ $-12$\arcdeg.
b) Contours are drawn at $-17$, 17, 30, {\bf 50}, 70, and 90\% of
the peak brightness, which was 580~\mbox{$\mu$Jy~beam$^{-1}$}. The rms background
brightness was 34~\mbox{$\mu$Jy~beam$^{-1}$}, and the convolving beam FWHM was $0.85 \times
0.35$~mas at p.a.\ $-20$\arcdeg.}
\label{fimage22}
\end{figure*}
\noindent We discuss the images in turn:
\begin{trivlist}
\item{a) 5 GHz, 1999 Feb.\ 21:} (Fig.~\ref{fimage5G}a) The northeast
bright spot dominates. There is a suggestion that it has an extension
towards the center. A version of this image was previously published
in \citet{SN86J-1}.
\item{b) 5 GHz, 2002 Nov.\ 10:} (Fig.~\ref{fimage5G}b) The northeast
bright spot continues to dominate the image. Again, a slight
extension towards the center is seen, but it cannot be clearly
identified as the central bright spot. A version of this image was
previously published in \citet{SN86J-Sci}.
\item{c) 5 GHz, 2005 Oct.\ 24:} (Fig.~\ref{fimage5G}c) The central
bright spot is now clearly seen at 5~GHz although it is not yet
as strong as the northeast bright spot. A version of this image was
previously published in \citet{SN86J-COSPAR-2}.
\item{d) 5 GHz, 2008 Oct.\ 25:} (Fig.~\ref{fimage5G}d) The central
bright spot now dominates at 5~GHz, and the northeast bright spot
is becoming relatively fainter. The shell continues to expand (see
\S~\ref{sexp} below for a determination of the expansion velocity).
\item{e) 8.4 GHz, 2006 Dec.\ 3:} (Fig.~\ref{fimage22}a) This run
benefited from the newly available 512~Mbits~s$^{-1}$ recording
bandwidth. The resulting image is dominated by the central bright
spot, with the northeast bright spot and the remainder of the shell
being only marginally discernible.
\item{f) 22 GHz, 2005 Apr.\ 25:} This image was very similar to the
later 22~GHz image in 2006, discussed below. As it was of notably
lower signal-to-noise ratio, we do not reproduce it here \citep[it was
reproduced in][]{SN86J-COSPAR-2}.
\item{g) 22 GHz, 2006 Dec.\ 10:} (Fig.~\ref{fimage22}b) At this
frequency, the steep-spectrum shell is not visible, and the central
bright spot dominates. The latter is now clearly resolved and appears
somewhat asymmetrical, with a brightness peak to the east. The
finger-like extension to the south is of marginal significance. This
image was deconvolved with CLEAN, but we note that maximum entropy
deconvolution (AIPS task VTESS) produces a very similar image.
\end{trivlist}
\subsection{Expansion}
\label{sexp}
The average expansion velocity and deceleration can be determined from
the VLBI data by estimating the size of the supernova at each epoch.
As SN~1986J's morphology is becoming increasingly irregular,
estimating the size by fitting a simple geometrical model to the
visibility data \citep[as we did for SN~1993J, see e.g.,][]{SN93J-1,
SN93J-2} is no longer practical. We take as a representative
(angular) radius the average radius of the contour which contains 90\%
of the total flux density in the image. Measures of the radius
derived from images are complicated by the convolution of the images
by a clean beam, but a consistent estimate of the expansion rate can
be obtained by comparing radii derived from images convolved with an
approximately co-evolving beam. For a fuller discussion, see
\citet{SN86J-1}, where we showed that the expansion up to 1999 had a
power-law form with the radius, $r \propto t^{0.71 \pm 0.11}$. We
accordingly use images convolved with a co-moving beam,\footnote{This
is the same
procedure we used in \citet{SN86J-1}, where we also found that the
expansion curve is not sensitive to the exact time-dependence of the
restoring beam, as long as it is similar to the time-dependence of the
radius. The use of an approximate value for the coefficient of the
expansion power-law should therefore not significantly bias our
result. Note that the value of 0.70 we use for this coefficient is
slightly different from the value of 0.75 that we used in
\citet{SN86J-1}. The present value is a somewhat better match to the
fit power-law expansion of the supernova. Note also that our present
value of the expansion coefficient is in agreement within the
uncertainties with the value we had obtained in \citet{SN86J-1}, but
does not support the higher value of $0.90 \pm 0.06$ reported by
\citet{Perez-Torres+2002}. }
whose dimensions scale with $t^{0.70}$.
We call this radius \mbox{$\theta_{\rm90\%\;flux}$}. Table~\ref{tthfl} gives the derived radii
for our various observing epochs. For observations after 2000, we do
not determine radii from the observations at 8~GHz or above because at
these frequencies, the shell was not sufficiently distinct for a
useful measurement. We further note that the increasingly
non-self-similar evolution of the morphology makes such average
measures of the radius and expansion less reliable, however, our fit
to the expansion should give a reasonable estimate of the properties
of the expansion averaged over the entire shell.
We fix the explosion date, $t_0$, at 1983.2, the value that we
determined in \citet{SN86J-1} from both the expansion measurements and
the radio light-curve. A least-squares fit of a power-law then gives
an average expansion $\mbox{$\theta_{\rm90\%\;flux}$} \propto t^{0.69 \pm 0.03}$, consistent
with, but more accurate than the value we had derived in
\citet{SN86J-1} for the measurements up to 1999.
\begin{deluxetable*}{l c c c c }
\tabletypesize{\small}
\tablecaption{Angular Sizes of SN~1986J}
\tablewidth{0.7\textwidth}
\tablehead{
\colhead{Date} & \colhead{Frequency}
& \colhead {Age\tablenotemark{a}}
&\colhead{Angular Size\tablenotemark{b}}\\
& \colhead{(GHZ)}
& \colhead{(yrs)}
& \colhead{(mas)}
}
\startdata
1988 Sep 29 & 8.4 & \phn 5.55 & $1.37 \pm 0.10$\tablenotemark{c} \\
1990 Jul 21 & 8.4 & \phn 7.35 & $1.83 \pm 0.10$\tablenotemark{c} \\
1999 Feb 22 & 5.0 & 15.94 & $3.07 \pm 0.10$\tablenotemark{c} \\
2002 Nov 10 & 5.0 & 19.66 & $3.60 \pm 0.14$ \\
2005 Oct 24 & 5.0 & 22.62 & $3.71 \pm 0.15$ \\
2008 Oct 26 & 5.0 & 25.62 & $4.23 \pm 0.17$ \\
\enddata
\tablenotetext{a}{The explosion date is not precisely known. We use a
best-fit explosion epoch of 1983.2 determined in \citet{SN86J-1}.}
\tablenotetext{b}{We take as a representative angular radius \mbox{$\theta_{\rm90\%\;flux}$},
which is the average radius of the area containing 90\% of the clean
flux density in the image. The images are convolved with an
approximately co-moving restoring beam. See text for details.}
\tablenotetext{c}{These values are from data discussed in
\citet{SN86J-1}, although the present values of \mbox{$\theta_{\rm90\%\;flux}$}\ differ slightly
from the ones reported there because we are using a slightly different
co-moving convolving beam. In that earlier reference, we used a
convolving beam whose size increased $\propto t^{0.75}$, whereas the
present values are based on a beam whose size increases $\propto
t^{0.70}$, which better matches the actual expansion of the
supernovae. The difference in derived radii are small, with the
values of \mbox{$\theta_{\rm90\%\;flux}$}\ changing by less than their stated uncertainties.}
\label{tthfl}
\end{deluxetable*}
We plot the resulting expansion curve in Figure~\ref{fexp}. The two
radii after epoch 1999 fall slightly below the earlier power-law fit,
suggesting a possible slight increase in the deceleration. However,
we caution that the increase in brightness of the central bright spot
relative to that of the shell might lead to a \mbox{$\theta_{\rm90\%\;flux}$}\ being a slight
underestimate of the true radius of the shell.
The average expansion velocity measured between the 1999 and 2008
epochs was $5700 \pm 1000$~\mbox{km s$^{-1}$}.
Although this velocity, based on \mbox{$\theta_{\rm90\%\;flux}$}, is likely not the exact
velocity of the outer radius of the shell of radio emission, it should
be a reasonable approximation thereof.
\begin{figure*}
\centering
\includegraphics[height=4.in]{sn86j_exp_thfl_logv3.eps}
\caption{The expansion curve of SN~1986J\@. We plot the angular outer
radius, \mbox{$\theta_{\rm90\%\;flux}$}\ (the angular radius which contains 90\% of the total
flux density in the image, see text, \S~\ref{sexp} for details),
against the age, both on logarithmic scales. The line indicates
expansion of the form $\theta \propto (t - t_0)^m$, with $m = 0.69$,
where $t = 0$ corresponds to the fixed epoch 1983.2, which was the
best weighted least-squares fit to the points plotted.}
\label{fexp}
\end{figure*}
\subsection{Positions and Proper Motion of the Two Bright Spots}
\label{svspot}
How does the expansion of the supernova compare with the proper
motions of the northeast and central bright spots? Since our images
are phase-referenced to the brightness peak of the un-related nearby
compact radio source, 3C~66A, we can accurately determine the proper
motion of the two bright spots relative to 3C~66A.
Before we proceed to the proper motions, we briefly discuss the
uncertainties which will apply to such measurements \citep[for other
discussions on the astrometric accuracy obtainable with VLBI,
see][]{SN93J-1, M81-2004, PradelCL2006}. Our estimated astrometric
uncertainties below consist of the following three components: 1) An
uncertainty due to the noise in the image. 2) An uncertainty in
identifying a particular point in the morphology of the reference source,
3C~66A, which is slightly time and frequency
dependent\footnote{\citet{Kovalev+2008a} show that
frequency-dependent shifts in core position are common among compact
radio sources. For example, see \citet{M81-2004} and \citet{M81-2000}
for measurements of the variability in time and frequency of the
compact core of the nearby galaxy M81.}.
We estimate the latter uncertainty component as 20~\mbox{$\mu$as}\ in RA and
70~\mbox{$\mu$as}\ in decl., with the decl.\ component being larger because the
source is elongated approximately in the north-south direction. 3) An
astrometric uncertainty due to errors in the station coordinates,
earth orientation, and tropospheric correction, which we estimate at
30~\mbox{$\mu$as}\ in each coordinate \citep[see][]{PradelCL2006}.
We first turn to the northeast bright spot. It was clearly detected
already in 1999 \cite[Bietenholz et al., 2002; for a similar image
from an independent reduction of our observations,
see][]{Perez-Torres+2002}\nocite{SN86J-1}. We find that, based on a
linear fit to the position in the 1999 Feb., 2002 Nov., 2005 Oct., and
2008 Oct.\ images at 5~GHz, the proper motion of the northeast bright
spot was $60 \pm 20$~\mbox{$\mu$as~yr$^{-1}$}, corresponding to a speed of $3000 \pm
1000$~\mbox{km s$^{-1}$},
$2400 \pm 250$ and $1500 \pm 700$~\mbox{km s$^{-1}$},
in RA and decl.\ respectively.
Does the current proper motion of the northeast bright spot place it
at the explosion center at the time of explosion? To answer this
question requires knowing the position of the explosion enter. As
mentioned earlier, this center is not accurately known for SN~1986J,
but as an estimate we use the average position of the center of the
shell, as described in \S~\ref{simages} above. Extrapolating the
northeast bright spot's present proper motion the explosion epoch of
1983.2 yields a position consistent with the estimated center position
within 1.9 and 0.8 sigma in RA and decl., respectively. If allowance
for deceleration is made, then the spot's 1983.2 extrapolated position
is even closer to the center. In fact, a radial motion with the same
power-law dependence on time as the overall expansion, which has $r
\propto t^{0.69}$, is compatible with our measurements, as can be seen
in Figure~\ref{fvspot}, where we show the radial motion of the
northeast bright spot.
A homologous, power-law expansion which includes the northeast bright
spot is therefore consistent with our measurements, with the spot's
present position and proper motion being compatible with it being at
the center of the shell at the explosion epoch. We note, however,
that the uncertainties are not small, so considerable deviation from
the power-law expansion, for example stronger deceleration, or motion
with a constant speed, are also consistent with our measurements.
We turn now to the central bright spot. It is best defined in the
images from 2003 June 22 \citep[15~GHz,][]{SN86J-Sci}, 2006 Dec.\ 10
(22~GHz) and 2008 Oct.\ 26 (5~GHz). If we estimate the center bright
spot's position at each epoch by the position of the brightness peak,
we can determine the spot's proper motion. A linear fit to the three
positions gives a proper motion of +4 and +31~\mbox{$\mu$as~yr$^{-1}$}\ in RA and
decl., respectively.
The morphology of the bright spot makes the peak brightness position
somewhat resolution-dependent, (and possibly frequency-dependent), and
we estimate that the accuracy with which any particular spot in the
morphology can reliably be identified to be on the order of 100~\mbox{$\mu$as}.
In addition to the other uncertainties estimated above, we use a
rounded value of 120~\mbox{$\mu$as}\ for the uncertainty in the spot position
in each coordinate, which leads to an uncertainty in the proper motion
of 31~\mbox{$\mu$as~yr$^{-1}$}\ in each coordinate. This implies a speed of $1500 \pm
1500$~\mbox{km s$^{-1}$}\ at 10~Mpc. The estimate of the proper motion is consistent
with the central bright spot being stationary. The average position
of the central bright spot is $300 \pm 150$~\mbox{$\mu$as}\ from the geometric
center of the shell (see \S~\ref{simages}).
However, given the uncertainty in associating the explosion center
with the geometric center of the shell, we only say that the position
of the central bright spot is marginally consistent with that of the
explosion center.
\begin{figure*}
\centering
\includegraphics[angle=-90, width=0.58\textwidth]{sn86j_spot.eps}
\caption{The angular distance of the northeast bright spot from the
center as a function of time. The additional axis labels along the
top give the calendar year. The spot's angular distance is taken to be
that between the spot local maximum and the estimated explosion center
(see text, \S~\ref{svspot}, for details on the estimation of the
explosion center).
The shell expands with outer angular radius $\propto t^{0.69}$ (see
text \S~\ref{sexp}). Of the possible expansion curves which have
$\theta \propto t^{0.69}$, we show as the solid line the one that best
matches the spot's motion. Note that the spot's displacement as a
function of time is somewhat dependent upon the choice of the
explosion center, although for reasonable choices, the nature of the
expansion curve remains unchanged. Our plotted error bars are the
uncertainty in the spot position and an estimated uncertainty of
0.1~mas in the explosion location added in quadrature. The spot's
motion is consistent with homologous expansion.}
\label{fvspot}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[width=0.58\textwidth]{sn86j_spotflux.eps}
\caption{The evolution at 5~GHz of the fractional flux density of the
bright spots in the northeast and in the center of the projected shell
as a function of time. The additional axis labels along the top give
the calendar year. The bright spot flux densities are estimated by
fitting, by least squares, a point source and a baseline level in the
region near the spot. The fraction of the flux density is taken to be
the fitted flux density for the point source divided by the total
CLEAN flux density in the image. For the first three epochs, the
bright spot in the center is not discernible. The chevron shows an
upper limit on the central bright spot flux density at age 19.7~yr,
which we derived by taking half the surface brightness at its
location.}
\label{ffspot}
\end{figure*}
\subsection{Brightness Evolution of the Two Bright Spots}
\label{sfspot}
In order to quantify the evolution of the brightness of the northeast
and central bright spots, we fitted simple models to each bright spot
in the image plane. In particular, in the region near each spot we
fit an elliptical Gaussian of the same dimensions as the restoring
beam, i.e., the equivalent of a point source, and a baseline level.
Although the assumptions that the bright spots are in fact unresolved
and that they can be separated in this way are not unquestionable, we
feel that this procedure gives a reasonable approximation of the
relative flux density of the two spots at each epoch, and is more
accurate than, for example, relying only the spot's peak brightness.
In Figure~\ref{ffspot} we show the 5-GHz evolution of the flux density
of each bright spot relative to the total flux density in the image as
a function of time. The northeast bright spot's relative brightness
increases, reaching a maximum of $\gtrsim$20\% of the total flux
density at an age of $\sim$16~yr, and fades thereafter. Recall,
however, that the total flux density is decreasing relatively rapidly
as a function of time (\S~\ref{svla}, Figure~\ref{flightcurve}), so in
absolute terms, the brightness of the northeast bright spot decreases
monotonically. At its relative peak, its flux density was $\sim
1.3$~mJy, and its radio luminosity, assuming an $\alpha = -0.61$
spectrum up to 100~GHz, was $\sim 6\times 10^{36}$~erg~s$^{-1}$.
The central bright spot was not clearly visible (at 5~GHz) till age
22.6~yr, but by age 25.6~yr, it had increased to $\sim$15\% of the
total flux density. The central bright spot is still increasing in
brightness in absolute terms, but the absolute increase is not as
rapid as the relative one plotted in Figure~\ref{ffspot}. In 2008,
the central bright spot had a flux density of $\sim 390 \;\mu$Jy, and
a radio luminosity of $\sim 2\times 10^{36}$~erg~s$^{-1}$ (again
assuming a spectrum with $\alpha = -0.61$ up to 100~GHz).
\section{DISCUSSION}
\subsection{Overall Evolution: Expansion Curve, Deceleration, and Spectral Index}
\label{sexpdecel}
The expansion curve in Figure~\ref{fexp} is compatible with a power-law
expansion, and earlier we found the best fit to correspond to $r \propto
t^{0.69}$. The flux density decay, however, shows a distinct change
in the power-law slope at some time between 1989 and 1995, i.e.,
between ages of 6 to 12~yr \citep{SN86J-1}. Unfortunately no flux
density measurements are available in this period to determine the
time of change more accurately. Similar behavior was found for
SN~1993J with a steepening of the flux density decay at $t \sim$7~yr
\citep{SN93J-2, Weiler+2007}.
In the case of SN~1993J, this steepening was found to correspond to a
decrease in deceleration \citep{SN93J-2}. This is the general
behavior expected when there is a relative decrease in the CSM
density. Is a similar behavior seen in SN~1986J? Examining the
expansion curve in Figure~\ref{fexp} \citep[see also expansion curves
including earlier points in][]{SN86J-1, SN86J-COSPAR} suggests that a
significant {\em decrease}\/ in deceleration in the age range of 6 to
12~yr is not compatible with the data, although a small {\em increase}\/
in deceleration might be. This would suggest that the increase in the
rate of flux density decay is due not to a relative decrease in the
CSM density but rather to one in the ejecta density. This suggests a
steep density profile in the ejecta in the exterior regions, with a
somewhat flatter profile in the interior. Such a flattening is indeed
expected from models of supergiant stars \citep[see, e.g., discussion
in][]{Chevalier2005}. Although suggestive, we do not regard this
result as definitive because the measurements do not demand any break
in the power-law expansion curve.
The evolution of the integrated spectrum is complex, but at low
frequencies the spectrum seems to have a power-law shape. One can with
reasonable confidence associate the part of the spectrum at the lowest
frequencies with the shell, with the complex evolution due to the
emergence of the central bright spot being restricted to higher
frequencies ($\gtrsim 5$~GHz). As we showed in \S~\ref{sspix} above,
the lowest part of the spectrum seems be initially rather steep, with
$\alpha = -0.67$, then by 1995 it flattens somewhat to $\alpha =
-0.44$, and subsequently it steepens again.
What could cause such a variation of the radio emission spectrum with
time? Since the emission spectrum of synchrotron radiation reflects
the energy spectrum of the relativistic particles, one obvious
interpretation is that the latter energy spectrum is also changing
with time, perhaps in response to varying conditions at the shock
front where the particle acceleration occurs. However, optical depth
effects also offer a possible explanation. Although most of the
radio-emitting shell is optically thin at cm wavelengths, the interior
of the shell is not expected to be so because of the high density of
thermal material. If mixing is occurring, a substantial transition
zone between the optically thin shocked material and the optically
thick ejecta may exist, which could partly absorb the radio emission
from a substantial fraction of the shell, and thus modify the
integrated spectrum. In fact, the slow turn-on of the lightcurve
suggests a significant amount of absorption mixed in with the radio
emission already at early times \citep{WeilerPS1990}.
We note that changes with time in the radio spectral index are seen in
other supernovae. In particular, a flattening with time of the radio
spectrum as seen for SN~1993J
\citep{SN93J-2}\footnote{\citet{ChandraRB2004b} also find a flattening
with time of the optically-thin spectrum of SN~1993J\@. A slight but
systematic flattening is also visible in the plotted data after $t =
500$~d of \citet{Weiler+2007}, although they fitted a model with a
constant spectral index for the optically thin part of the spectrum.}
and at least at the latest times also for SN~1979C
\citep{SN79C-shell}, may be characteristic of the evolution of radio
supernovae, and could be coupled with a steepening of the lightcurve
decay.
\subsection{Comparison of Expansion Curve to Optical Radial Velocities}
\label{soptvel}
From our VLBI measurements, we determined the expansion velocity of
the edge of the radio emission region in the plane of the sky. Since
the forward shock forms the outer boundary of the region where radio
emission is expected to occur, our VLBI expansion velocity can be
reasonably associated with the velocity of the forward shock velocity
\citep[see discussion on this subject for SN~1993J in][]{SN93J-4}.
Note that since the geometry of SN~1986J is somewhat irregular, the
assumption of sphericity which allows a comparison of the velocities
in the sky-plane measured with VLBI with those along the line-of-sight
measured from optical spectra is likely true only to first order.
In the optical, there are a number of spectral lines present which
allow radial velocity measurements. Recently,
\citet{Milisavljevic+2008}, obtained new optical spectra at epoch
2007.7, and compared them to earlier ones from 1989.7 and 1991.8.
The optical emission lines are thought to mainly come from two
distinct locations. Narrow optical lines, chiefly H$\alpha$, He~I and
[N~II] $\lambda 5755$, are thought to originate in dense, shock-heated
circumstellar medium, and therefore have relatively narrow
line-widths, typically $<2000$~\mbox{km s$^{-1}$}. These lines were more prominent
early on and have faded since 1989.
Currently the most prominent emission lines are the [O I] $\lambda
\lambda 6300, 6364$ and [O II] $\lambda \lambda$ 7319, 7330 forbidden
lines of oxygen, with the [O III] $\lambda \lambda 4959, 5007$ line
also being discernible. In 1989.7, these lines showed a double-peaked
structure, with one peak at a velocity of $-1000$~\mbox{km s$^{-1}$}\ (relative to
the systemic one), and the other near $-3500$~\mbox{km s$^{-1}$}, and the maximum
velocity, on the blue sides of these lines, was $\sim -6000$~\mbox{km s$^{-1}$}.
Similar velocities were seen in 1991.8\@. By 2007.7, the double
structure was less clearly discernible, the maximum blue velocity had
decreased and the [O III] line had almost vanished.
\citet{Milisavljevic+2008} attribute these forbidden lines to ejecta
which has been heated (but not yet shocked) by the reverse shock.
This interpretation of the optical lines is consistent with the
expansion velocities measured in the radio. In 1989.7, the forward
shock velocity was $\sim$8300~\mbox{km s$^{-1}$}, and the reverse shock velocity
would be expected to be about 80\% of this, or $\sim$6600~\mbox{km s$^{-1}$}.
The velocities implied by both the line-center and the total widths
seen in the forbidden oxygen lines are consistent with being inside
the reverse shock in 1989.7. By 2007.7, the forward shock velocity is
$\sim$5500~\mbox{km s$^{-1}$}, and the expected reverse shock velocity
$\sim$4400~\mbox{km s$^{-1}$}. The evolution of the optical line profiles is
consistent with this reduction in the velocity of the reverse shock in
that, particularly for the [O II] line, with the highest (blue)
velocities now being $\sim -4300$~\mbox{km s$^{-1}$}. This suggests that, of the
ejecta responsible for the oxygen forbidden line emission, the highest
velocity fraction passed through the reverse shock between 1989.7 and
2007.7, and the remainder is now situated relatively close to the
reverse shock.
\subsection{Magnetic Field}
\label{sbfield}
The magnetic field can be estimated from the radio brightness and the
source size using minimum energy arguments
\citep{Pacholczyk1970}\footnote{We note that \citet{BeckK2005} point
out a number of shortcomings in the standard formulae of Pacholczyk
and give revised formulae. However, for relatively flat spectra, the
differences between the standard and revised formulae are not large
and only weakly dependent on the field strength, so a field calculated
using the revised formula would have almost the same dependence on
time and radius. In the interests of easier comparison with earlier
results we estimate our magnetic field using the standard formula from
\citet{Pacholczyk1970}.}.
Given that the spectrum shows deviations from a power-law, we will
perform the calculation only for the shell component, which dominates
below $\nu = 5$~GHz and is observed to have a power-law spectrum with
an average $\alpha = -0.5$ (\S~\ref{sspix}). The derived field
strength depends on the ratio, $k$, of energies of relativistic
protons and electrons. We use a value of $k = 100$, as was found to
be reasonable for strong shocks in supernova remnants by
\citet{BeckK2005}. We estimate first the field strength in the shell,
for which we use the angular radii derived in \S~\ref{sexp} above, and
we take the flux density at 1.4~GHz to be representative of the shell
(as opposed to that of the central bright spot).
Using the values from 1988 to 2008, we find that $B$ has an
approximate power-law dependence on time and radius, as might be
expected from the approximately power-law evolution of both the flux
density and radius. In particular, we find that for minimum energy,
$B \simeq 60 \, (t/{\rm 10 \, yr})^{-1.3}$~mG,
and that $B \propto r^{-1.8}$ for $t > 1500$~d. Although such
estimates of the field are uncertain by factors of several, the
estimates of its dependence on $t$ and $r$ are likely to be somewhat
better constrained with the exponents being uncertain by perhaps
$\sim$70\%.
\subsection{The Mass Loss Rate of the Progenitor}
\label{smassloss}
Can we estimate the mass-loss rate of the progenitor from our
observations? This mass-loss rate is often estimated from studies of
the radio lightcurves. It is determined mostly by the rising part of
the lightcurve, while the emission is optically thick, and is usually
expressed as $\dot M$/\mbox{$v_{\rm w}$}, where \mbox{$v_{\rm w}$}\ is the wind velocity. However,
such calculations can significantly overestimate the value of $\dot
M$/\mbox{$v_{\rm w}$}\ \citep[see, e.g.,][]{SN79C}, and in general have been shown to
be somewhat problematic if not based on complete physical models
\citep[e.g.,][]{FranssonB2005}. In particular, in the case of
SN~1986J, \citet{WeilerPS1990} cited a value of $\dot M$/\mbox{$v_{\rm w}$}\ =
2.4\ex{-4} \mbox{$M_\sun$~yr$^{-1}$}\ (for \mbox{$v_{\rm w}$}\ = 10~\mbox{km s$^{-1}$}), based not on the rising
part of the lightcurve, but rather on a short term variation in
the optical depth seen in late 1988\@. Later, \citet{Weiler+2002}
reported a notably lower value of 4.3\ex{-5} \mbox{$M_\sun$~yr$^{-1}$}, which is based
on a model involving clumpy absorption in the CSM\@.
\citet{Houck+1998} suggest a value of a few times $10^{-4}$~\mbox{$M_\sun$~yr$^{-1}$}\
based on X-ray observations. In any case, given the departures from
self-similarity and the considerable evidence of clumpiness in the CSM,
which will have a strong effect on both radio brightness and
absorption, any estimated mass-loss rate is likely to be quite
uncertain.
A mass-loss rate between 4\ex{-5} and, say, 3\ex{-4} \mbox{$M_\sun$~yr$^{-1}$}\ with \mbox{$v_{\rm w}$}
= 10~\mbox{km s$^{-1}$}\ would imply that by age 25.6~yr (2008.8), when the radius
of the shock-front is 6.3\ex{17}~cm, the swept-up mass would be $0.8
\sim 6$~\mbox{$M_\sun$}.
For this range in mass-loss rate, the kinetic energy of the swept-up
mass would be $(0.2 \sim 1.7) \times 10^{51}$~erg ($v \simeq
5400$~\mbox{km s$^{-1}$}). Since the kinetic energy of the swept-up mass can be at
most a fraction of the original kinetic energy of the ejecta, which
was likely not much larger than $10^{51}$~erg
\citep[e.g.,][]{Jones+1998}, we regard the higher end of this
mass-loss range as unlikely on energetic grounds. We therefore
estimate a present swept-up mass of $0.8 \sim 2$~\mbox{$M_\sun$}, and an average
mass-loss rate ($\dot M / v_{\rm w}$) in the range of $(4 \sim 10)
\times 10^{-5}$~\mbox{$M_\sun$~yr$^{-1}$}.
Such a range of mass-loss rate implies a CSM density at the 2008.8
outer shock radius of 6.3\ex{17}~cm
of $(5 \sim 13)\ex{-22}$~g~cm$^{-3}$. If we assume that the fraction
of hydrogen atoms is 75\% of the total, then the hydrogen number
density is between $230 \sim 570$~cm$^{-3}$.
\subsection{The Northeast Bright Spot}
\label{sshellspot}
The images are characterized by two bright spots, whose flux densities
vary. The first one, in the northeast, appears to be a component of
the shell. It was first clearly discernible at age $\sim$16~yr,
although it may have been present already earlier. Its proper motion
is consistent with homologous expansion with the bulk of the shell
(Fig.~\ref{fvspot}). As can be seen in Figure~\ref{ffspot}, at its
peak relative brightness (age $\sim$16~yr) approximately $\sim$20\% of
the total flux density originates from this spot. Unfortunately, our
observations do not constrain the relative lightcurve around the time
of the peak precisely, the peak relative brightness may have been as
high as $\sim$30\% of the total, although it must have occurred
between our 1990 and 1999 epochs. The northeast bright spot appears
marginally resolved in our images. On 2005 Oct.\ 24 our FWHM resolution
was $1.65 \times 0.83$~mas, so the spot's angular diameter must have been
$\sim$1~mas,
corresponding to 14\% of the shell's outer diameter, or 1.5\ex{17}~cm.
Since the northeast bright spot appears to be part of the projected
shell, the most plausible interpretation is that it is due to the
ejecta interacting with a dense clump in the CSM\@. This would
account for the relative brightness of the spot first increasing and
then falling off. In fact, the relative lightcurve of the northeast
bright spot in Figure~\ref{ffspot} suggests that the shock has
traversed the bulk of the clump between ages 10 to 20~yr. Our
expansion curve suggests that, on average around the circumference,
the shock travelled a distance of 2\ex{17}~cm during this time, in
reasonable agreement with the clump diameter estimated above.
If we assume both the dense clump and the outer shock to be spherical,
with the clump having the dimensions estimated above, then, when the
shock is midway through the clump, $\sim$1\% of the outer shock front
area is interacting with the dense clump. Since the radio emission is
driven by the shock, we will consider the radio luminosity per unit
area of the shock, and we will further assume that the radio
luminosity scales as the 5-GHz flux density.
The northeast bright spot at its peak emitted about one quarter of the
5-GHz flux density of the remainder of the shell. we can therefore
calculate that the region in the dense clump must be $\sim$25 times
brighter than the remainder. Our proper motion measurements show that
the velocity of the shock in the clump is approximately the same as it
is for the remainder of the shell, so the radio luminosity per unit
area of the shock should scale with the CSM density\footnote{The radio
luminosity is thought to scale with the post-shock energy density
\citep[e.g.,][]{Chevalier1998, Chevalier1982b}. In the absence of
absorption, the radio luminosity will depend on the fractions of the
post-shock energy which go into the magnetic field and into
accelerated particles. There is no solid ground for thinking that
these fractions are constant, indeed they are likely to vary somewhat
both as a function of shock velocity and the density and chemical
composition of the CSM\@. It is likely, however, that these
variations are relatively small, whereas the variation in the CSM
density can be several orders of magnitude or more.}.
We found above (\S~\ref{smassloss}) that the present average CSM
density was $(5 \sim 13) \ex{-22}$~g~cm$^{-3}$, suggesting that the
densities in the clump are $(1.3 \sim 3.2) \ex{-20}$~g~cm$^{-3}$,
corresponding to hydrogen number densities of $5800 \sim
14,000$~cm$^{-3}$, giving the clump a total mass of $0.01 \sim
0.03$~\mbox{$M_\sun$}. We note, however, that these densities represent an
average over the clump, and much higher densities in parts of the
clump are not excluded by the radio data.
There are a number of other indicators pointing to the existence of
such dense clumps in the CSM of SN~1986J\@. First, the very high
Balmer decrement \citep{Rupen+1987}, and other features in the
late-time spectra \citep{Milisavljevic+2008}, suggest the presence of
very high number densities ($> 10^6$~cm$^{-3}$). Second, the slow
rise of the radio lightcurve is interpreted as also implying a very
clumpy CSM \citep{WeilerPS1990}. Third, because the H$\alpha$ line,
which has a width of only $\sim$1000~\mbox{km s$^{-1}$}\
\citep[e.g.,][]{Milisavljevic+2008, Leibundgut+1991}, is far too
narrow to arise from the ejecta near the forward shock ($v \sim
5700$~\mbox{km s$^{-1}$}; \S~\ref{sexp}), it has been suggested that most of the
H$\alpha$ emission is coming not from the ejecta, but rather from
shocked, dense clouds in the CSM resulting from a very clumpy wind
from the progenitor star \citep{ChugaiD1994, Chugai1993}.
Red supergiant stars of mass 20-40~\mbox{$M_\sun$}\ are known to have episodes of
strong mass loss, resulting in large density contrasts in the CSM\@.
For example, \citet{Smith+2001, SmithHR2009} show that the red
supergiant VY Canis Majoris (mass 20-40~\mbox{$M_\sun$}), which is likely
similar to the progenitor of SN~1986J, is surrounded by a nebula
produced by rapid mass loss ($2\sim4 \times 10^{-4}$~\mbox{$M_\sun$~yr$^{-1}$}). This
nebula, which extends out to a radius of $\gtrsim 10^{17}$~cm, similar
in scale to the CSM required for SN~1986J, is highly structured with
knots and filamentary arcs, which are distributed in an asymmetric
way. \citet{Smith+2001} argue that the structures in the wind were
produced by ejection events which were localized on the stellar
surface. For VY Canis Majoris then, and by implication also for
other red supergiants, the evidence seems to suggest that the winds
can be highly structured, with large density contrasts on small spatial
scales. Exactly such structure seems to be required to reproduce the
radio emission from SN~1986J.
\subsection{The Central Bright Spot: Is It Emission from the Compact Remnant of
the Supernova Explosion?}
\label{scompact}
What is the nature of the central bright spot --- is it the emission
associated with a neutron-star or black-hole compact remnant of the
supernova explosion? The first question would be whether the central
bright spot is positionally coincident with the explosion center. As
we showed in \S~\ref{svspot} above, the position of the central bright
spot is in fact compatible with being in the center of the shell, and
its proper motion is consistent with being stationary with respect to
the shell. Note that the expansion speed of the shell is much larger
than the peculiar velocities of stars and even of pulsars, which
latter have a velocity dispersion of $200 \sim 300$~\mbox{km s$^{-1}$}\ in the
Galaxy \citep[e.g.,][]{Hobbs+2005,LyneG1990},
so the effect of any peculiar motion of the progenitor or any ``kick''
received in the explosion would be small and the compact remnant would
be expected to remain essentially in the center of the shell over the
$\sim$25~yrs since the supernova explosion. In this respect, our
position and proper motion measurements for the central hot spot are
consistent with the expectations for a black-hole or neutron star
compact remnant.
We first noted the inversion in the integrated radio spectrum in
\citet{SN86J-1}, and we showed in \citet{SN86J-Sci} that the inversion
in the integrated spectrum was in fact associated with the central
bright spot, with the shell and the northeast bright spot discussed
above having a simple power-law spectrum as is usually seen in
supernovae. In \S~\ref{svla} and Figure~\ref{fspectra} above we show
that both the inflection and the high-frequency turnover points in the
radio spectrum are progressing downwards in frequency as SN~1986J
ages. The inverted spectrum of the central bright spot suggests
partly absorbed radio emission. An opacity decreasing due to the
expansion would cause the observed progression to lower frequencies of
both the inversion and high-frequency turnover points.
We argued in \citet{SN86J-COSPAR} that the absorption is probably due
to free-free absorption rather than to synchrotron self-absorption
(SSA)\@. In fact, our latest 22~GHz images, where the central bright
spot is somewhat resolved, largely rules out SSA on the following
grounds. Examination of the image (Fig.~\ref{fimage22}b) suggests a
size of $\gtrsim 1 \times 10^{17}$~cm for the central bright spot.
The radius below which SSA would be important, $R_p$, can be
calculated following \citet{Chevalier1998}:
\[
R_p = 8.8\times10^{15}\kappa^{-1/10}
\left( \frac{f}{0.5}\right)^{-1/19}
\left( \frac{F_p}{\rm Jy}\right)^{9/19} \]
\[
\times
\left( \frac{D}{\rm Mpc}\right)^{18/19}
\left( \frac{\nu}{\rm 5 GHz}\right)^{-1} {\rm cm}
\]
where $F_p$ is the flux density at the spectral peak, $\kappa$ is the
ratio of relativistic electron energy density to magnetic energy
density, $D$ is the distance, and $f$ is the filling factor. Assuming
equipartition ($\kappa = 1$) and substituting our values of $F_p$ =
3.8~mJy at $\nu_p$ = 14~GHz (2007.6), we obtain $R_p \simeq 1 \times
10^{16}$~cm,
which is far lower than the size determined from the image of $\gtrsim
1 \times 10^{17}$~cm, thus effectively ruling out significant SSA at
our observing frequencies.
The appearance of the central bright spot at a time long after the
supernova's integrated radio spectrum was optically thin at cm
wavelengths suggested that any remaining absorption due to the
circumstellar material was minimal. It was natural, therefore, to
assume that the absorption seen for the central bright spot was due to
un-shocked material interior to the shocked shell, rather than to the
CSM\@. With this assumption, the central bright spot's likeliest
physical location was in the center of the shell, and its most likely
interpretation was in terms of radio emission associated with the
neutron star or black hole expected to have been left behind after the
explosion \citep{SN86J-Sci, SN86J-COSPAR, SN86J-COSPAR-2}.
The turnover frequency is approximately equal to
the frequency at which the optical depth to free-free absorption,
$\tau$, is unity, is given by
$$\nu_{\tau = 1} = 0.3 \, (T_e^{-1.35} N_e^2 \, dl)^{1/2} \; {\rm
GHz,}$$ where $T_e$ is the electron temperature in K, $N_e$ is the
number density of electrons in cm$^{-3}$, assumed constant along
the path-length in pc, $dl$, with $N_e^2 \, dl$ being the emission
measure.
Our integrated spectra (Fig.~\ref{fspectra}) show that the spectral
peak due to the central bright spot occurs at $\sim$20~GHz in 2002.4.
If we assume\footnote{Although $T_e$ is not well
known, and can vary strongly with radius, this is probably a
reasonable value for radii $>1 \times 10^{17}$~cm \citep[see,
e.g.,][]{LundqvistF1988}.}, a $T_e$ of $10^4$~K then $N_e^2 \, dl$ is
$1.1\ex{9}$~cm$^{-6}$~pc. By 2007.6, the turnover frequency is
$\sim$14~GHz, so $N_e^2 \, dl = 0.6\ex{9}$~cm$^{-6}$~pc.
The temporal dependence of the absorption is consistent with that
expected from the gas within the expanding shell. We found that the
emission measure decreased by a factor of $\sim$2 between 2002.4 and
2007.6. If we assume the system is expanding homologously, then the
density within the shell is proportional to $r^{-3}$ and $N_e^2 \, dl
\propto r^{-5}$. If we further take $r \propto t^{0.69}$ as we found
for the shell, we would expect $N_e^2 \, dl$ to be $\propto
t^{-3.45}$, which would lead to a decrease by close to the observed
factor of 2 between 2002.4 and 2007.6.
We note, however, that the inverted part of the spectrum is much
flatter than is expected from free-free absorption, which produces
spectra with an exponential cutoff at low frequencies. Such flatter
spectra below the turnover are produced when a range of different
optical depths is present. This suggests that the absorbing material
is somewhat fragmented, with some lines of sight having much lower
optical depths than others. Indeed, fragmentation of the ejecta is
not surprising, since there are a number of instabilities operating in
the expanding shell of ejecta
\citep[e.g.,][]{Gull1973,JunN1996a,ChevalierB1995}, which are expected
to lead to fragmentation. If the central bright spot is due to a
pulsar, then further instabilities occur when the young pulsar's wind
nebula expands into the supernova ejecta
\citep[e.g.,][]{BandieraPS1983,ChevalierF1992}.
As we have shown, the absorption of the radio emission from the
central bright spot is consistent with what might be expected from the
intervening material in the shell and its observed expansion rate.
This suggests that the integrated spectrum above the turnover point
shows the intrinsic, un-absorbed spectrum of the central bright spot.
In fact, the integrated spectra in Figure~\ref{fspectra} show that the
spectral index above the turnover point is very similar to that below
the inversion point. So, the intrinsic spectral index of the central
bright spot is also similar to that of the shell.
If the central bright spot were in fact in the physical center of the
shell and represented radio emission associated with the compact
remnant of the supernova --- either a black hole or a neutron star ---
then it would be somewhat of a coincidence for its radio emission to
have a similar spectral index as that of the shell emission. In
particular, if the central bright spot were the wind nebula (PWN)
around a very young pulsar, then one might expect it to have a
somewhat flatter spectral index, as most PWNe have spectral indices in
the range $-0.3$ to 0.0 \citep{GaenslerS2006} and as there are no
filled-center remnants in the catalog of \citet{Green2004} which have
broadband radio spectra steeper\footnote{see {\tt
http://www.mrao.cam.ac.uk/surveys/snrs} for an updated version of the
catalog. The possible exception is Vela~X, the central source in the
Vela supernova remnant, which is often associated with the PWN\@.
\citet{Alvarez+2001} found $\alpha^{\rm 8.4 GHz}_{\rm 0.09 GHz} =
-0.39 \pm 0.03$ for Vela~X, indicating a relatively steep spectrum for
a PWN\@. More recently, however, \citet{Hales+2004} reported on 31~GHz
observations of a strong source near the Vela pulsar that they
identify with the PWN, and for which they find a slightly inverted
spectrum with $\alpha^{\rm 31 GHz}_{\rm 8.4 GHz} = +0.10 \pm 0.06$}
than $\alpha = -0.3$. In particular, the youngest known PWNe, the
Crab Nebula and G21.5$-$0.5, have rather flat radio spectra with
spectral indices of $-0.27$ \citep{Crab-1997} and +0.08
\citep{G21.5expand}, respectively. The observed steep spectrum of the
central bright spot, therefore, would not seem to favor the hypothesis
of a young PWN\@. However, an only 27-yr old PWN may not be comparable
to the much older known PWNe, so the observed steep spectrum does
not rule out a young PWN.
Alternatively, if the central bright spot represents emission from
jets emanating from a black hole environment, the radio spectrum is
consistent with the expectations but the radio luminosity is not. The
possible presence of both self-absorbed and optically thin components
in the jet can give rise to both flat and steep spectra, so the
observed spectrum can easily be accommodated. However, SN~1986J's
central bright spot is far more radio-luminous than any known
stellar-mass black hole system. In particular, a ``fundamental
plane'' relationship between the radio and X-ray luminosities and
black-hole mass has been observed for black-hole systems of a wide
range of masses \citep[see, e.g.,][]{FalckeKM2004, Ho2008}. The
unabsorbed X-ray flux from SN~1986J was measured in late 2003 by
\citet{Houck2005a} to be $\sim 1.6\times 10^{-13}$~erg~s$^{-1}$. Even
with the assumption that all this X-ray flux is due to the central
bright spot, the ``fundamental plane'' relation would suggest radio
luminosities for a black-hole system of $\sim 10^{-4}$ of that
observed for the central bright spot of $\sim 2 \times
10^{36}$~erg~s$^{-1}$ (\S~\ref{sfspot}). However, very little is
known about such young black-hole systems, so it is not impossible
that they would be far more luminous than expected from the
fundamental plane relationship. So again, we must conclude that the
radio properties of the central bright spot do not seem to argue
strongly for an accreting black-hole system, but neither can they
exclude it.
\subsection{The Central Bright Spot as a Shell Component?}
\label{scenterclump}
As we showed in \S~\ref{sshellspot} above, the northeast bright spot
can naturally be explained as the impact of the shock-front on a dense
condensation in the CSM\@. Could the central bright spot be a similar
phenomenon, with the dense condensation being coincidentally located
near the projected center of the shell? In this case, an approximate
equality of the intrinsic spectral indices of the central bright spot
with that of the remainder of the supernova is expected, since usually
not much variation of the spectral index around the shell is observed.
As we showed, there is considerable evidence that the CSM of red
supergiants and of SN~1986J in particular are quite clumpy, and the
northeast bright spot suggests that there is at least one dense
condensation in the CSM of SN~1986J\@. There is no particular reason
to only expect a single dense clump in the CSM, so the existence of
more than one dense clump seems not unlikely.
The expanding shell impacting on a dense CSM clump would cause a local
brightening of the radio emission. Although the bulk of the CSM is
optically thin at cm wavelengths, clumps with densities as high as
$N_e > 10^6$~cm$^{-3}$, and sizes small compared to the shell
diameter, would be expected to have large optical depths for cm-wave
radio emission. For example, a clump with a diameter of 5\ex{16}~cm,
corresponding to an angle of $\sim$0.3~mas on our images, and a
density of 1\ex{6}~cm$^{-3}$ would have
a free-free optical depth at 8.4~GHz of $\sim$30.
The shock hitting a dense clump will produce localized, bright radio
emission, as is seen in the northeast bright spot. However, if the
dense CSM clump lies on the near side of the expanding shell,
fortuitously near the center in projection, then the optically thick
clump itself will in fact block most of this bright radio emission.
With time, the clump will become optically thin as it either fragments
or as the shock eats through it. This scenario would produce a
sequence much like what is seen for the central bright spot: a delayed
turn on, dependent on the distance of the clump from the explosion
center and its density, with an inverted spectrum expected from
free-free absorption, followed eventually by an optically thin decay.
At 5~GHz, the brightness of the central bright spot is still
increasing, (Fig.~\ref{ffspot}) so we are still in the rising part of
its lightcurve.
If we interpret the central bright spot as emission from the shock
interacting with such a CSM clump, what can be deduced about the
physical conditions in this clump? In our latest image at 22~GHz, we
can estimate an angular diameter of the central bright spot of
$\sim$0.7~mas (FWHM), suggesting a linear diameter of $\sim 1 \times
10^{17}$~cm or 0.03~pc. The turnover frequency at this
epoch is $\sim$12~GHz. If we take the turnover frequency to be that
at which the free-free optical depth is unity, the density of the
absorbing clump can be estimated from the electron number density
which in turn can be calculated from the emission measure using the
equation given in \S~\ref{scompact} above. If we again assume $T_e =
10^4$~K
and a constant $N_e$, and further assume that the
clump's line-of-sight depth is equal to the central bright spot's
diameter in the plane of the sky, we can calculate that $N_e \sim
2.5\ex{5}$~cm$^{-3}$. Again, the presence of
significantly higher densities cannot be excluded, but the presence of
substantial material at the calculated density seems required. The
total mass of such an absorbing clump would be $\sim$0.1~\mbox{$M_\sun$}\
(assuming full ionization).
The swept-up mass in 2008.8 is $0.8 \sim 2.0$~\mbox{$M_\sun$}\ (out to a radius of
$6.3\times10^{17}$~cm; see \S~\ref{smassloss}), so the mass of the
clump represents 13\% to 5\% of the total swept-up mass.
By comparing the part of the spectrum above 20~GHz with that at low
frequencies (see Fig.~\ref{fspectra}), we can estimate that the
unabsorbed radio luminosity of the central bright spot is about twice
that
of the remainder of the supernova. Note that the unabsorbed luminosity
of the central bright spot is currently larger than that of the
northeast bright spot, since a substantial fraction of the central
bright spot's luminosity is still absorbed, with the turnover
frequency currently being $\sim$12~GHz. If we again assume that radio
luminosity per unit shock area scales only with the CSM density, we
can estimate the density of the clump relative to that of the average
CSM density as we did for the northeast bright spot
(\S~\ref{sshellspot}). Using the above clump diameter of 1\ex{17}~cm,
and a shell outer radius of $6\ex{17}$~cm (for 2006 Dec.)\ we find
that the spot covers
$\sim$0.2\% of the shell surface. The clump must therefore be $\sim
1000\times$ denser than the average corresponding CSM\@. As might be
expected, since the central bright spot has a much higher unabsorbed
luminosity than the northeast bright spot, the density required for
the relevant enhancement of the emissivity is also much higher.
We calculated a representative density of $N_e \sim
2.5\ex{5}$~cm$^{-3}$ for the central clump above, which suggests a
representative average density for the CSM at $r \simeq 6.2\ex{17}$~cm
of $\sim 250$~cm$^{-3}$, which
in turn implies a mass-loss rate of 6\ex{-5} \mbox{$M_\sun$~yr$^{-1}$}\ (for \mbox{$v_{\rm w}$} =
10~\mbox{km s$^{-1}$}) which is consistent with the values derived in
\S~\ref{smassloss} above.
As mentioned in \S~\ref{soptvel} above, the oxygen forbidden-line
emission seems to be dominated by two dense clumps, both on the near
side of the remnant. If the central radio bright spot is due to a CSM
clump on the near side of the SN, then also in the radio, there would
be evidence for two dense clumps. Could the clumps responsible for
radio bright spots be be the same as the two clumps seen in the
optical emission lines? The association seems unclear, since the
radio emission, as we have argued, would be due to dense clumps in the
CSM, exterior to the forward shock, while the forbidden-line emission
is plausibly attributed ejecta heated by the reverse shock, that is
interior to the reverse shock.
While it is conceivable that dense CSM clumps cause distortions first
in the forward shock but subsequently also in the reverse shock, which
latter could in turn give rise to the bright features seen in the
oxygen forbidden lines, any association between the radio bright spots
and the oxygen-line features must remain speculative.
All said, we think now that the spectral evolution seen in the new
observations make an interpretation of the central bright spot in
terms of a second shell component and an interpretation in terms of
emission from a PWN or a black-hole environment equally plausible.
\section{SUMMARY AND CONCLUSIONS}
\begin{trivlist}
\item{1.} We have obtained new multi-frequency VLA flux density
measurements and VLBI images of SN~1986J, showing the evolution of this
supernova in the radio.
\item{2.} The evolution of the integrated radio spectrum is complex.
The spectrum at the lowest frequencies has a spectral index in the
range of $-0.7 \sim -0.5$. The spectrum above the high-frequency
turnover, presumably the intrinsic spectrum of the central bright spot,
is equally steep within the uncertainties.
\item{3.} The shell continues to expand. The average expansion speed
of the shell between 1999 and 2009 was $5700 \pm 1000$~\mbox{km s$^{-1}$}. This
speed is compatible with continued power-law expansion, with the
radius increasing $\propto t^{0.69\pm0.03}$. The increase in the rate
of flux density decay at $t \sim 7$~yr is likely due to a flattening
in the profile of the ejecta profile rather than a steepening in the
one of the CSM, as no increase in deceleration is observed.
\item{4.} The equipartition magnetic field decreases as the supernova
expands, with an approximate value of $B \simeq 60 \; (t/{\rm
10 \; yr})^{-1.3}$~mG\@. The dependence on the outer shock front radius is
$B \propto r^{-1.8}$.
\item{5.} Various estimates of the mass-loss converge to values in the
range of $(4 \sim 10) \times 10^{-5}$ \mbox{$M_\sun$~yr$^{-1}$}\ (for \mbox{$v_{\rm w}$} = 10~\mbox{km s$^{-1}$}).
\item{6.} The VLBI images show two bright spots in addition to the
shell-like structure. The first is in the northeast of the shell,
while the second is near the projected center of the shell. The flux
densities of the spots relative to that of the SN as a whole vary with
time, with that of the northeast bright spot increasing till
$\sim$1999, and decreasing since, while that of the central bright
spot continues to increase.
\item{7.} The northeast bright spot is likely due to a dense clump in
the circumstellar material (CSM). The bright spot's proper motion is
consistent with homologous power-law expansion. Number densities of
$N_e \sim 10^4$~cm$^{-3}$ or higher are suggested for the clump.
\item{8.} The central bright spot has a partly absorbed radio
spectrum, with an intrinsic radio spectrum that is similar to that of
the shell. The amount of absorption is decreasing with time.
\item{9.} The central bright spot's original interpretation as
originating in the physical center of the shell and being emission due
to the neutron star or black hole remnant of the supernova is
supported by its central position, and its stationarity to within
$1\sigma$. The new observations, however, suggest an equally
plausible alternative explanation of the central bright spot being
radio emission due to the shell impacting upon a second dense CSM
clump, fortuitously located on the near side of the shell close to the
projected center. The amount of absorption suggests number densities
of $N_e \gtrsim 2.5$\ex{5}~cm$^{-3}, \sim 10^3$~higher than
the average density of the CSM.
\item{10.} From the VLBI images from 1987 to 2008, we produced a movie
showing the supernova's evolution.
\end{trivlist}
\acknowledgements
The European VLBI Network is a joint facility of European and Chinese
radio astronomy institutes funded by their national research
councils. We have made use of NASA's Astrophysics Data System Abstract
Service. We thank the anonymous referee for his suggestions.
\bibliographystyle{apj}
|
\section{Introduction}
Massive galaxy clusters provide several means for independently
examining any viable cosmological model. Cluster samples used for this
purpose are usually complete in terms of X-ray and redshift
measurements such as the RDCS (Rosati et al. 1998, Borgani et al. 1999) and the MACS
survey (Ebeling, Edge, \& Henry 2001, Ebeling et al. 2007) that we examine here with lensing. Ideally, controlled samples of clusters with reliable lensing masses would allow the most direct comparison with theory but these are presently still in their infancy with only a handful of clusters discovered in wide-field weak lensing searches (Taylor et al. 2004, Wittman et al. 2006, Massey et al. 2007, Miyazaki et al. 2007), and may be often subject to projection biases (e.g., Hennawi \& Spergel 2005). Weak lensing surveys should have the advantage of being approximately volume-limited over the redshift range 0.1<$z$<1.0, where the lensing distance ratio has a broad maximum.
Ongoing measurements of the Sunyaev-Zel'dovich (SZ) effect (which
has already been detected in many tens of clusters, including some at
relatively high redshifts $z\sim 0.9$, e.g., LaRoque et al. 2006) in a
large number of clusters over a wide redshift range should improve upon X-ray
selection, due to the redshift-independence of the effect (though clearly high
redshift clusters are relatively less well resolved). Also, since the insensitivity of the effect
to internal gas density variations will not be biased as with X-ray
flux selection towards systems with luminous shocked gas,
measurements should better correlate with cluster masses.
Strong gravitational lensing (SL) is nearly always seen in
sufficiently detailed observations of massive clusters. Einstein radii
derived from such observations provide a reliable projected mass
within the observed Einstein radius, which depends only on fundamental
constants, G and c, and knowledge of the lens and source distances.
Constraining the inner mass profile from SL is much harder, requiring
many sets of multiple images covering a wide range of source redshift,
to overcome inherent lensing degeneracy between the gradient of the
mass profile and the scaling of the bend angle with source distance,
so that SL based mass profiles have been usefully constrained for only
several clusters producing reliable mass maps (e.g., Abell 370, Kneib et al. 1993, Richard et al. 2009a; Abell 901, Deb et al. 2010; Abell 1689, Broadhurst et al. 2005, Coe et al. 2010; Abell 1703, Limousin et al. 2008; Cl0024+1654, Liesenborgs et al. 2008, Zitrin et al. 2009b; MS 2137.3-2353, Gavazzi et al. 2003, Merten et al. 2009; RXJ1347, Brada\v{c} et al. 2008a, Halkola et al. 2008; SDSS J1004+4112, Sharon et al. 2005, Liesenborgs et al. 2009; "The bullet cluster", Brada\v{c} et al. 2006).
In quite a few of these clusters there is a significant inconsistency between the Einstein radius directly measured from SL analyses and the Einstein radius predicted from model profiles fitted to weak-lensing measurements of the same clusters, often subject to background selection and dilution problems over a wide radial range (e.g., Medezinski et al. 2010), though more so towards the SL regime where the degree of contamination by cluster members is often poorly corrected. The importance of a SL analysis of a significant sample is clear, especially in the statistical aspect, where the modelling method is similar for all examined clusters and thus supplies a coherent view for objective comparison (e.g., Richard et al. 2009b).
The development of SL modelling-methods has increased in response to
the improvement in data quality since the discoveries of the first giant arcs (e.g., Lynds \& Petrosian 1986, Soucail et al. 1987, Kneib et al. 1996). Most methods can be classified as ``parametric'' if based on physical parameterisation, and as ``non-parametric'' if they are ``grid-based'' (see also \S4.4 in Coe et al. 2008, and references therein). Still, many of these methods include too many parameters to be
well-constrained by the number of initially known multiply-lensed systems. Here we use ACS imaging to identify new multiply-lensed systems in order to constrain the mass distributions and define the critical curves of the sample, motivated by the successful minimalistic approach of Broadhurst et al. (2005) to lens modelling. In following work (Zitrin et al. 2009b) we presented an improved modelling method which involved only 6 free parameters enabling easier constraint, since the number of constraints has to be equal or larger to the number of parameters in order to get a reliable fit. Two of these parameters are primarily set to reasonable values so only 4 of these parameters have to be constrained initially, which sets a very reliable starting-point using obvious systems. The mass distribution is therefore primarily well constrained, uncovering many multiple-images which can be then iteratively incorporated into the model, by using their redshift estimation and location in the image-plane.
The Massive Cluster Survey (MACS) has been aimed to create a complete sample of the very X-ray luminous clusters in the Universe, successfully
increasing the number
of known such clusters
to hundreds, particularly at $z>0.3$ (Ebeling, Edge, \& Henry 2001). From this a
complete sample of 12 high-$z$ MACS clusters ($z>0.5$) was defined by Ebeling
et al. (2007) which have proved very interesting in several follow-up studies
including deep X-ray, SZ and HST imaging. Detection of a large-scale filament has been reported in the case of MACS J0717.5+3745 by
Ebeling et al. (2004), for which many multiply-lensed images have been recently identified by Zitrin et al. (2009a), revealing this object to
be the largest known lens, with an Einstein radius equivalent to 55$\arcsec$ (for a source at $z\sim2.5$). In the case of MACS J1149.5+2223 (Zitrin \& Broadhurst 2009), a background spiral galaxy at $z=1.49$ (Smith et al. 2009) has been shown to be multiply-lensed into several very large images, requiring a very shallow, unrelaxed central mass distribution. Another large multiply-lensed sub-mm source at a redshift of $z\simeq2.9$ has been identified in MACS J0454.1-0300 (also referred to as MS 0451.6-0305; Takata et al. 2003, Borys et al. 2004, Berciano Alba et al. 2007, 2009), MACS J0025.4-1222 was found to be a ``bullet cluster''-like (Brada\v{c} et al. 2008b), and other MACS clusters have been recently used for an extensive arc statistics study (Horesh et al. 2010).
The X-ray data available for this sample (see Ebeling et al. 2007) along with the high-resolution HST/ACS imaging and
SZ imaging (e.g., Laroque et al. 2003) make these MACS targets particularly useful for understanding the nature of the most massive clusters. Here we complete a full SL analysis of this 12 cluster sample with the available deep HST/ACS imaging, principally to derive their Einstein radii and
also to help constrain the central mass distributions. The effective Einstein radii derived here are the corresponding radii of circles of equivalent areas to those enclosed within the critical curves, or similarly, the radii within which the averaged $\kappa$ is equal to 1. In addition, this work can also supply detailed magnification maps to help motivate deeper searches for high-$z$ galaxies behind these clusters.
Another major motivation for pursuing accurate lensing-maps
is the increased precision of model predictions for cluster-
size massive halos in the standard $\Lambda$CDM model and close variance (see Umetsu \& Broadhurst 2008, e.g., Bullock et
al. 2001, Spergel et al. 2003, 2007, Tegmark et al. 2004, Hennawi et al. 2007, Neto et al. 2007, Duffy et
al. 2008, Keselman, Nusser \& Peebles 2009, Meneghetti et al. 2010, Fedeli et al. 2010). In the standard hierarchical model, large massive bodies are due to collapse recently and thus are not expected to be found in large numbers at high redshifts. On the other hand, the larger volume available with increasing distance means that in
practice we cannot expect to reside next to the most massive
cluster (see also Zitrin et al. 2009a). The increasing number of accurately analysed clusters have revealed larger lenses than predicted by the standard $\Lambda CDM$ model (Broadhurst \& Barkana 2008, Puchwein \& Hilbert 2009). This discrepancy is empirically supported by the surprisingly concentrated mass profiles measured for such clusters, when combining the inner strong lensing with the outer weak lensing signal (Gavazzi et al. 2003, Broadhurst et al. 2005 \& 2008, Limousin et al. 2008, Donnarumma et al. 2009, Oguri et al. 2009, Umetsu et al. 2009, Zitrin et al. 2009b,2010), or independently, when using the internal dynamics of cluster members (Lemze et al. 2009) and deep X-ray data (Lemze et al. 2008). Geometrically, lensing is considered to be optimised at intermediate redshifts,
where for a given mass the critical density for lensing is minimal, but
this is partly offset by the late hierarchical growth of high-mass systems. This trade-off results in estimates
of the amplitude of strong lensing to favour the redshift range
$z=0.2-0.4$. Our recent analysis of two X-ray selected MACS clusters reveals very large lenses (e.g., MACS J1149.5+2223, Zitrin \& Broadhurst 2009; MACS 0717.5+3745, Zitrin et al. 2009a) at high redshift, increasing
the tension with $\Lambda$CDM predictions, since large Einstein radius clusters
at high-$z$ require earlier formation
than implied by the standard $\Lambda$CDM model (Sadeh \& Rephaeli 2008). The
largest lensing clusters have proven to be excellent targets for accessing the
faint early Universe due to their large magnification consistently providing
the highest redshift galaxies (Ebbels et al. 1996, Franx et al. 1997, Frye \&
Broadhurst 1998, Bouwens et al. 2004, Kneib et al. 2004, Bradley et al. 2008,
Zheng et al. 2009).
The statistics of large Einstein radii provide an important opportunity to test the standard $\Lambda$CDM paradigm, as it probes both the high-mass end of the cluster mass function and central mass distributions of massive clusters (Oguri \& Blandford 2009). Increasing numbers of theoretical and observational works have been done recently, deriving Einstein radius distributions of various samples. For example, Cypriano et al. (2003) have derived the mass distributions of 24 X-ray selected Abell clusters via weak-lensing. By fitting to SIS profiles they obtain a mean Einstein radius of $\sim17 \arcsec$. More recently, Hoekstra (2007) finds by weak-lensing analysis and SIS fitting a mean Einstein radius of $\sim14 \arcsec$ for a sample of 20 X-ray luminous clusters. Okabe et al. (2009) find the mass distribution and Einstein radii for a sample of $\sim30$ LoCuSS clusters by fitting the weak-lensing data to CIS profiles, obtaining a mean value of $\sim22 \arcsec$ (see Okabe et al. 2009 and references therein). Note that these values quoted here are calculated from the corresponding tables in these papers, and are not necessarily scaled or normalised to certain lens and source redshifts.
Other recent work more relevant to our study summarises SL analysis for a sample of 20, mostly relaxed \emph{undisturbed} clusters (Richard et al. 2009b). They find that the Einstein radii are distributed log-normally with a peak at $\theta_{e}=14.45 \arcsec$ and a corresponding $1.95 \times 10^{14} M_{\odot}$ enclosed mass (within R$<$250 kpc), and show that the predicted distribution of Einstein radii from $\Lambda$CDM cosmology falls short of the observed Einstein radii by a factor of 2. It has been claimed that much larger Einstein radii can be contemplated only with mass distributions which are highly prolate and aligned along the line of sight (Corless \& King 2007, Oguri \& Blandford 2009, see also Hennawi et al. 2007, Meneghetti et al. 2007, Sereno, Jetzer \& Lubini 2010). Analysing the 12 X-ray selected, high-$z$ MACS clusters is important since they are predicted to be massive and should form efficient lenses, whose properties can be then compared to such studies, playing an important role in probing the $\Lambda$CDM scenario.
The paper is organised as follows: in \S2 we describe the sample and observations; the lensing
analysis is described in \S3 detailing each cluster separately. Our
overall results are presented and discussed in \S4, followed by a Conclusion (\S5).
Throughout the paper we adopt the standard cosmology ($\Omega_{\rm m0}=0.3$,
$\Omega_{\Lambda0}=0.7$,$h=0.7$).
\section{The Sample and Observations}
The $z>0.5$ MACS clusters have been imaged with the Wide Field Channel
(WFC) of the ACS installed on HST, mainly in the framework of proposal
ID 9722 (PI: Ebeling). This mainly comprises two-filter observations
in the F555W and F814W bands, taken during 2003 and 2004 with typical
exposure times of $\sim 4500$s in each filter. We make use of some other
HST data available in the Hubble Legacy Archive (HLA), which we list
in Table \ref{sample}. In addition, available SExtractor (Bertin \& Arnouts
1996) photometry catalogues were also downloaded from the HLA, and were used
to construct colour-magnitude diagrams for identifying the red cluster sequence
galaxies belonging to each cluster, as a starting point for each
lensing model, as detailed in \S \ref{model}.
\begin{table*}
\caption{Properties of the MACS $ z>0.5$ sample. The following data (columns 1-7) are based on Ebeling et al. (2007): \emph{Column 1:} cluster name in the MACS survey; \emph{Columns 2 \& 3:} The RA and Declination of the X-ray centroids (as determined from Chandra ACIS-I data); \emph{Column 4:} redshifts; \emph{Column 5:} velocity dispersions, in km s$^{-1}$; \emph{Column 6:} Chandra X-ray luminosities in $10^{44}$ erg $s^{-1}$ quoted for the $0.1$-$2.4$ keV band. These luminosities are quoted within $r_{200}$ and exclude X-ray point sources; \emph{Column 7:} X-ray temperatures, measured from Chandra data within $r_{1000}$, but excluding a central region of 70 kpc radius around the listed X-ray centroid; \emph{Column 8:} Ebeling et al. (2007) morphology code - assessed visually based on the appearance of the X-ray contours and the goodness of the optical/X-ray alignment. The assigned codes (from apparently relaxed to extremely disturbed) are 1 (pronounced cool core, perfect alignment of X-ray peak and single cD galaxy), 2 (good optical/X-ray alignment, concentric contours), 3 (nonconcentric contours, obvious small-scale substructure), and 4 (poor optical/X-ray alignment, multiple peaks, no cD galaxy), errors are estimated as less than 1; \emph{Column 9:} ACS bands used here. Note, the clusters MACS J0018.5+1626 and MACS
J0454.1-0300 are better known from earlier work as Cl0016+1609 and
MS 0451.6-0305, respectively. For more information see Ebeling et
al. (2007).}
\label{sample}
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|c|c|c|}
\hline\hline
MACS & $\alpha$ (J2000.0) & $\delta$ (J2000.0) & $z$ & $\sigma$ & $L_{x, Chandra}$ & $KT (kev)$ & M.C.E. & ACS bands\\
\hline
J0018.5+1626 & 00 18 33.835 & +16 26 16.64 & 0.545 & $1420^{+140}_{-150}$ & $19.6\pm{0.3}$ & $9.4\pm{1.3}$ & 3 & F606W,F775W,F850LP\\
J0025.4-1222 & 00 25 29.381 & -12 22 37.06 & 0.584 & $740^{+90}_{-110}$ & $8.8\pm{0.2}$ & $7.1\pm{0.7}$ & 3 & F450W,F555W,F814W\\
J0257.1-2325 & 02 57 09.151 & -23 26 05.83 & 0.505 & $970^{+200}_{-250}$ & $ 13.7\pm{0.3}$ & $10.5\pm{1.0}$& 2 & F555W,F814W\\
J0454.1-0300 & 04 54 11.125 & -03 00 53.77 & 0.538 & $1250^{+130}_{-180}$ & $16.8\pm{0.6}$ & $7.5\pm{1.0}$ & 2 & F555W,F775W,F814W\\
J0647.7+7015 & 06 47 50.469 & +70 14 54.95 & 0.591 & $900^{+120}_{-180}$ & $15.9\pm{0.4}$ & $11.5\pm{1.0}$& 2 & F555W,F814W\\
J0717.5+3745 & 07 17 30.927 & +37 45 29.74 & 0.546 & $1660^{+120 }_{-130}$ & $24.6\pm{0.3}$ & $11.6\pm{0.5}$& 4 & F555W,F606W,F814W\\
J0744.8+3927 & 07 44 52.470 & +39 27 27.34 & 0.698 & $1110^{+130 }_{-150}$ & $22.9\pm{0.6}$ & $8.1\pm{0.6}$& 2 & F555W,F814W\\
J0911.2+1746 & 09 11 11.277 & +17 46 31.94 & 0.505 & $1150^{+260 }_{-340}$ & $7.8\pm{0.3}$ & $8.8\pm{0.7}$& 4 & F555W,F814W\\
J1149.5+2223 & 11 49 35.093 & +22 24 10.94 & 0.544 & $1840^{+120 }_{-170}$ & $17.6\pm{0.4}$ & $9.1\pm{0.7}$ & 4 & F555W,F814W\\
J1423.8+2404 & 14 23 47.663 & +24 04 40.14 & 0.543 & $1300^{+120}_{-170}$ & $16.5\pm{0.7}$ & $7.0\pm{0.8}$ & 1 & F555W,F814W\\
J2129.4-0741 & 21 29 26.214 & -07 41 26.22 & 0.589 & $1400^{+170}_{-200}$ & $15.7\pm{0.4}$ & $8.1\pm{0.7}$ & 3 & F555W,F814W\\
J2214.9-1359 & 22 14 57.415 & -14 00 10.78 & 0.503 & $1300^{+90}_{-100}$ & $14.1\pm{0.3}$ & $8.8\pm{0.7}$ & 2 & F555W,F814W\\
\hline
\end{tabular}
\end{center}
\end{table*}
\section{Strong Lensing Modelling and Analysis}\label{model}
We apply our well tested approach to lens modelling, which we have
applied successfully before to various clusters uncovering unprecedentedly
large numbers of multiply-lensed images in A1689, Cl0024, MACS J1149.5+2223 and MACS J0717.5+3745 (respectively, Broadhurst et al. 2005, Zitrin
et al. 2009b, Zitrin \& Broadhurst 2009, Zitrin
et al. 2009a). The full details of this approach can be found in these
earlier papers. Briefly, the basic assumption adopted is that mass
approximately traces light, so that the photometry of the red cluster
member galaxies is the starting point for our model. In a recent paper (Zitrin et al. 2010) we analyse Abell 1703 and show that this assumption is well based, by comparison to an assumption-free non-parametric technique (Liesenborgs et al. 2006) which yields a very similar overall mass distribution. In addition, we show that similar to other clusters we have analysed to date, our modelling method has the inherent flexibility to find and reproduce many multiple-images even if initially constrained by only a few obvious systems.
As mentioned, the starting point of the model are cluster member galaxies, which are identified as lying close to the cluster sequence by the photometry provided in the Hubble Legacy Archive (in this process sometimes massive foreground galaxies are included as well, scaled down by the relative distance ratio, since they can locally affect nearby images). We then approximate the large scale distribution of matter by assigning a
power-law mass profile to each galaxy, the sum of which is then
smoothed. The degree of smoothing ($S$) and the index of the power-law ($q$) are
the most important free parameters determining the mass profile. A worthwhile improvement in
fitting the location of the lensed images is generally found by
expanding to first order the gravitational potential of the smooth
component, equivalent to a coherent shear describing the overall matter ellipticity, where the direction of the
shear and its amplitude are free, allowing for some flexibility in
the relation between the distribution of DM and the distribution of
galaxies which cannot be expected to trace each other in detail. This freedom also allows the effective centre to slightly shift, as was the case in our analysis of Cl0024 (Zitrin et al. 2009b). The
total deflection field $\vec\alpha_T(\vec\theta)$, consists of the
galaxy component, $\vec{\alpha}_{gal}(\vec\theta)$, scaled by a
factor $K_{gal}$, the cluster DM component
$\vec\alpha_{DM}(\vec\theta)$, scaled by (1-$K_{gal}$), and the
external shear component $\vec\alpha_{ex}(\vec\theta)$:
\begin{equation}
\label{defTotAdd}
\vec\alpha_T(\vec\theta)= K_{gal} \vec{\alpha}_{gal}(\vec\theta)
+(1-K_{gal}) \vec\alpha_{DM}(\vec\theta)
+\vec\alpha_{ex}(\vec\theta),
\end{equation}
where the deflection field at position $\vec\theta_m$
due to the external shear,
$\vec{\alpha}_{ex}(\vec\theta_m)=(\alpha_{ex,x},\alpha_{ex,y})$,
is given by:
\begin{equation}
\label{shearsx}
\alpha_{ex,x}(\vec\theta_m)
= |\gamma| \cos(2\phi_{\gamma})\Delta x_m
+ |\gamma| \sin(2\phi_{\gamma})\Delta y_m,
\end{equation}
\begin{equation}
\label{shearsy}
\alpha_{ex,y}(\vec\theta_m)
= |\gamma| \sin(2\phi_{\gamma})\Delta x_m -
|\gamma| \cos(2\phi_{\gamma})\Delta y_m,
\end{equation}
and $(\Delta x_m,\Delta y_m)$ is the displacement vector of the
position $\vec\theta_m$ with respect to a fiducial reference position,
which we take as the lower-left pixel position $(1,1)$, and
$\phi_{\gamma}$ is the position angle of the spin-2 external
gravitational shear measured anti-clockwise from the $x$-axis.
The normalisation of the model and the relative scaling of the
smooth DM component versus the galaxy contribution brings the
total number of free parameters in the model to 6.
Note that since
our goal is to find the Einstein radius distribution and the masses within the critical curves, and since many clusters lack multiple-images redshift information, we do not attempt to accurately constrain here the mass profiles of the sample clusters and thus the parameters $q$ and $S$ are relatively irrelevant. Any reasonable values for these parameters yield similar critical curves, Einstein radii and
thus also the corresponding masses enclosed within them.
This effectively leaves us with 4 parameters per model which can be easily constrained by the multiple-images incorporated and found here. Still, we have explored also the $q$ and $S$ parameter space for each cluster to verify this. In addition, we showed that in cases where there are too few bands in order to obtain reliable photometric redshifts and where no spectroscopic redshifts are available, a very efficient assumption is that the outer, blue multiple-images are at a redshift of $z\sim2-2.5$. This assumption is based on previous analysis of many clusters for which the vast majority of the outer blue images are at this redshift range due to the \emph{nesting} effect by which the critical curve expands for higher redshift objects, yet the redshift of these blue images is limited by the Lyman-alpha break to be lower than $z\sim3$. For example, this assumption was proven to be very accurate in MACS J1149.5+2223, where the intermediate-distance blue images we assumed at $z\sim2$ (system 3 in Zitrin \& Broadhurst 2009) were later spectroscopically found to be at $z=1.89$, and our corresponding estimation of the spiral galaxy at $z\simeq1.5$ (system 1 in Zitrin \& Broadhurst 2009), was later verified spectroscopically to be at $z=1.49$ (Smith et al. 2009). In addition, in our works on A1689 and Cl0024 we showed that the same minimum is obtained both when minimising according to the images location in the image-plane, and when minimising according to the photometric redshifts. This means that in practice also the slope which generally can be constrained only using redshifts information, can be well approximated by minimising solely the image-plane RMS of the reproduced images, especially if a clear minimum is seen (see Zitrin et al. 2009b). Note, here we aim to constrain the critical curves and the mass enclosed within them which are both relatively independent and indifferent to the profile, enabling an accurate measurement also without redshift information, which in turn can be used to approximate the magnification of the lens, since this quantity is profile dependent.
Firstly we use our preliminary models to lens various well detected candidate lensed galaxies back to the
source plane using the derived deflection field, and then relens this
source plane to predict the detailed appearance and location of
additional counter images, which may then be identified in the data by
morphology, internal structure and colour. We stress that multiple images found this way
must be accurately reproduced by our model and are not simply eyeball
``candidates'' requiring redshift verification. Note, due to the volume of this work we do not attempt to exhaust each cluster field finding great numbers of multiple-images, and concentrate instead on finding clear examples of multiply-lensed systems to constrain the models and obtain the Einstein radius. The fit is assessed by the RMS uncertainty in the image plane:
\begin{equation} \label{RMS}
RMS_{images}^{2}=\sum_{i} ((x_{i}^{'}-x_{i})^2 + (y_{i}^{'}-y_{i})^2) ~/ ~N_{ima
ges},
\end{equation}
where $x_{i}^{'}$ and $y_{i}^{'}$ are the locations given by the
model, and $x_{i}$ and $y_{i}$ are the real images location, and the
sum is over all $N_{images}$ images. The best-fit solution is obtained by
the minimum RMS, and the uncertainties are determined by the
location of predicted images in the image plane.
Importantly, this image-plane minimisation does not suffer from the
well known bias involved with source plane minimisation, where
solutions are biased by minimal scatter towards shallow mass profiles
with correspondingly higher magnification. The model is successively
refined as additional sets of multiple images are identified and then
incorporated to improve the fit, using also their redshifts measurements or estimates for better constraining the mass slope through the cosmological relation of the $D_{ls}/D_{s}$ growth (see Zitrin et al. 2009b). We detail this procedure for each cluster in the corresponding subsection below, each followed by corresponding figures of the multiple-images and the critical curves, and the mass distribution (Figures \ref{curves0018} to \ref{contours2214}).
\subsection{MACS J0018.5+1626}
The galaxy cluster MACS J0018.5+1626 ($z=0.546$, also known as Cl 0016+1609, or MS 0016) has been subject to extensive study, mainly due to its high X-ray luminosity, strong SZ effect, and high redshift (e.g., Dressler \& Gunn 1992, Dressler et al. 1999, Luppino et al. 1999, Ellis \& Jones 2002, Clowe et al. 2000, 2003, Laroque et al. 2003), setting an example of a rich distant cluster (Worrall \& Birkinshaw, 2003, and references therein). It has been established that this cluster is the major part of a large-scale filamentary structure (Worrall \& Birkinshaw, 2003, and references therein, Tanaka, 2007). This fact and the prominent X-ray emission imply a high cluster mass, which should be manifested in multiply-lensed systems spread throughout the field. However, due to lower quality optical data and the complexity of its structure, no multiply-lensed systems were found in MACS J0018.5+1626 up to date. Only one strongly-lensed arc was referred to before in this cluster (Lavery 1996), yet no spectroscopic redshift nor counter-images were introduced.
We begin our analysis by constructing an initial model in the method described above (\S \ref{model}), where as a starting point we insert averaged values to the various parameters, since no system is known a-priori to constrain the fit. We then notice the similar colours and symmetry of the images of systems 1, 2, and 3. After verifying that the reproduction of these systems is physically likely in the context of our model, we use these systems to fully constrain the model and derive the critical curves (see Figure \ref{curves0018}), their corresponding Einstein radius ($\theta_{e}=28\pm2 \arcsec$ for a source at $z_{s}\sim2$; systems 2 \& 3) and the mass enclosed within them ($1.46 \pm 0.1 \times 10^{14} M_{\odot}$), see Figures \ref{curves0018} and \ref{contours0018} for more information. Accordingly, the estimated redshift of system 1 is $z_{s}\sim1.5$.
There are various other mass estimates
for this cluster, mainly from weak-lensing analyses. Smail et al. (1997) quote values of $\sim 2.9 \times 10^{14} M_{\odot}$ and $\sim 1.87 \times 10^{14} M_{\odot}$ within 200 kpc (depending on the method, see also Smail et al. 1995), and Hoekstra (2007) quotes a value of $\sim 7.9 \times 10^{14} M_{\odot}$ within 500 kpc, all in general accordance with our measurement. The Einstein radius for this cluster was also mentioned before. Hoekstra (2007) quotes $\sim 9 \arcsec$ from a SIS model fit to the tangential
distortion from 0.25 to 1.5 $h^{-1}$ Mpc, while Williams, Navarro \& Bartelmann (1999) quote $25 \arcsec$ according to the arc discussed by Lavery (1996), in accordance with our estimation.
The reference centre of our analysis is fixed near the centre of the F775W (program ID 10493) ACS frame at: RA = 00:18:33.41, Dec = +16:26:14.95 (J2000.0), where one arcsecond corresponds to 6.42 kpc at the redshift of this cluster.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{0018curves2.jpg}
\end{center}
\caption{Galaxy cluster MACS J0018.5+1626 ($z=0.546$, also known as Cl 0016+1609, or MS 0016) imaged
with Hubble/ACS F606W, F775W, and F850LP bands. The blue curve
overlaid shows the tangential critical curve corresponding to the
distance of system~1 at an estimated redshift of $z\sim1.5$, and which
passes through the close triplet of lensed arcs in this system. The larger
critical curve overlaid in white corresponds to the larger source
distance estimated as $z\sim2-2.5$, passing through the close pairs of the candidate images of systems 2 and 3. This critical curve encloses a
large lensed region, with an equivalent Einstein radius of $\sim150~kpc$ at the
redshift of the cluster. Note, our model predicts tiny images of pairs 2.1/2.2 and 3.1/3.2 at the other side of the cluster, which we are not able to securely detect due to poor depth and lack of colour range.}
\label{curves0018}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm, trim=5mm 0mm 5mm 5mm,clip]{contours0018scale.jpg}
\end{center}
\caption{2D surface mass distribution ($\kappa$), in units of the
critical density, of MACS J0018.5+1626. Contours are shown in linear units, derived from
the mass model constrained using the multiply-lensed images seen in
Figure \ref{curves0018}. Axes are in ACS pixels ($0.05 \arcsec /pixel$), and a $20\arcsec$ scale bar is overplotted.}
\label{contours0018}
\end{figure}
\subsection{MACS J0025.4-1222}
The galaxy cluster MACS J0025.4-1222 ($z=0.584$) and its merging properties were recently analysed by Brada\v{c} et al. (2008b), which found multiply-lensed images of 4 background galaxies, and obtained spectroscopic redshifts for a pair of them and
photometric redshifts for the other two systems: system A+B
(system 1 here) at $z_{spec}=2.38$, system C (system 2 here) at
$z_{phot}=1^{+0.5}_{-0.2}$, and system D (system 3 here) at
$z_{phot}=2.8^{+0.4}_{-1.8}$. We use systems 1 and 2 (see Figure \ref{curves0025}) and one more
uncovered arc (system 4 in Figure \ref{curves0025}; $z_{s}\sim1.9$), to derive the corresponding Einstein radius ($\theta_{e}= 30\pm2 \arcsec$ for $z_{s}\sim2.38$) and the mass enclosed within it ($2.42^{+0.10}_{-0.13}\times 10^{14} M_{\odot}$), in excellent agreement with the result of Brada\v{c} et al (2008b; $\simeq 2.5 \times 10^{14} M_{\odot}$, within 300 kpc centred on the SE BCG). In addition we find that system A in Brada\v{c} et al (2008b; system 1 here) includes also an additional arc on the other side of the BCG, as seen in Figure \ref{curves0025} (arc 1.2), which our model reproduces very accurately (see Figure \ref{rep0025}). The reference centre of our analysis is fixed near the centre of the ACS frame at: RA = 00:25:29.61, Dec = -12:22:52.89 (J2000.0), where one arcsecond corresponds to 6.64 kpc at the redshift of this cluster.
\begin{figure}
\begin{center}
\includegraphics[width=90mm,trim=0mm 0mm 0mm 0mm,clip]{0025curves2.jpg}
\end{center}
\caption{Galaxy cluster MACS J0025.4-1222 ($z=0.584$) imaged
with Hubble/ACS F555W and F814W bands. The overlaid critical curve (white) corresponds to system 1 ($z_{s}=2.38$), enclosing a critical area of an effective Einstein radius of $\sim 200$ kpc at the redshift of this cluster.}
\label{curves0025}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm, trim=5mm 0mm 5mm 5mm,clip]{contours0025scale.jpg}
\end{center}
\caption{2D surface mass distribution ($\kappa$), in units of the
critical density, of MACS J0025.4-1222. Contours are shown in linear units, derived from
the mass model constrained using the multiply-lensed images seen in
Figure \ref{curves0025}. Axes are in ACS pixels ($0.05 \arcsec /pixel$), and a $20\arcsec$ scale bar is overplotted.}
\label{contours0025}
\end{figure}
\subsection{MACS J0257.1-2325}
In the galaxy cluster MACS J0257.1-2325 ($z=0.505$) various arcs are seen, from which we use 13 multiply-lensed images, belonging to $\sim7$ background distant galaxies, in the redshift range $z\sim1$ to $z\sim4$ to fully constrain the model. The reference centre of our analysis is fixed near the centre of the ACS frame at: RA = 02:57:07.24, Dec = -23:26:02.88 (J2000.0), where one arcsecond corresponds to 6.17 kpc at the redshift of this cluster.
This cluster was not analysed before and we did not find any arc redshift information. We assume a redshift of $z_{s}\sim2-2.5$ for the faint distant blue images (system 4 in Figure \ref{curves0257}) according to which the spectacular images next to the cD galaxy (system 1; see also Figure \ref{rep0257}) are at $z_{s}\sim1$, the blue images of systems 2 and 3 are at $z_{s}\sim1.5-2$, and the red drop-out candidates are at $z_{s}\sim3.5-4$ (system 5) helping to constrain the fit. We find that the critical curves (Figure \ref{curves0257}) have an equivalent Einstein radius of $39\pm2 \arcsec$ (for $z_{s}\sim2$) and enclose a mass of $3.35^{+0.58}_{-0.10} \times 10^{14} M_{\odot}$.
.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{0257curves2.jpg}
\end{center}
\caption{Galaxy cluster MACS J0257.1-2325 ($z=0.505$) imaged
with Hubble/ACS F555W and F814W bands. The tangential critical curves overlaid in blue correspond to a source redshift of $z_{s}\sim1$, passing through the close pair of images 1.2 and 1.3. The larger critical curves overlaid in white correspond to a source redshift of $z_{s}\sim2-2.5$, passing through the close pair of images 4.1 and 4.2.}
\label{curves0257}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=85mm,trim=5mm 0mm 5mm 5mm,clip]{contours0257scale.jpg}
\end{center}
\caption{2D surface mass distribution ($\kappa$), in units of the
critical density, of MACS J0257.1-2325. Contours are shown in linear units, derived from
the mass model constrained using the multiply-lensed images seen in
Figure \ref{curves0257}. Axes are in ACS pixels ($0.05 \arcsec /pixel$), and a $20\arcsec$ scale bar is overplotted.}
\label{contours0257}
\end{figure}
\subsection{MACS J0454.1-0300}
The galaxy cluster MACS J0454.1-0300 ($z=0.538$, also known as MS 0451.6-0305) has been subject to extensive study due to its high X-ray luminosity, strong SZ effect, and high redshift (e.g., Ellingson et al. 1998, Molnar et al. 2002, Laroque et al. 2003, Geach et al. 2006), and is known to host various lensed sub-mm sources with radio counterparts (Takata et al. 2003, Borys et al. 2004, Berciano Alba et al. 2007, 2009), which we incorporate in our analysis below.
Previous strong-lensing models of this cluster were introduced by Borys et al. (2004) and Berciano Alba et al. (2009) with similar symmetry as our model.
We fully constrain our model with the arcs and their redshift information listed in these papers (see also Figure \ref{curves0454} here), from which we derive an effective Einstein radius of $19\pm2 \arcsec$ for a source redshift of $z_{s}=2.9$ (system 1 in Figure \ref{curves0454}) and a mass of $0.82^{+0.03}_{-0.01} \times 10^{14} M_{\odot}$ enclosed within this critical curve. For a source redshift of $z_{s}\sim2$ we correspondingly get an effective Einstein radius of $13\pm2 \arcsec$, and a mass of $0.41^{+0.03}_{-0.01} \times 10^{14} M_{\odot}$ enclosed within the critical curve (see Figure \ref{curves0454}). Berciano Alba et al. (2009) quote a mass of $1.73 \times 10^{14} M_{\odot}$ within 30 $\arcsec$ of the cluster centre. Our model yields $1.8 \times 10^{14} M_{\odot}$ within 30 $\arcsec$ of the cluster centre, in full agreement. In our modelling process we find two other systems of multiply-lensed images. The first (system 3) is a triplet of faint arcs next to the core, at an estimated redshift of $z_{s}\sim1.5-2$. The second system (number 4) consists of 4 bright images, all have prominent emission in the K'-band (see Takata et al. 2003), corresponding to a redshift of $z_{s}\simeq 2.9$ similar to system 1. See Figure \ref{curves0454} for more details.
The reference centre of our analysis is fixed near the centre of the F555W (proposal ID 9722) ACS frame at: RA = 04:54:10.36, Dec = -03:01:03.02 (J2000.0), where one arcsecond corresponds to 6.37 kpc at the redshift of this cluster.
\begin{figure}
\begin{center}
\includegraphics[width=85mm,trim=0mm 0mm 0mm 0mm,clip]{0454curves2.jpg}
\end{center}
\caption{Galaxy cluster MACS J0454.1-0300 ($z=0.538$, also known as MS 0451.6-0305) imaged
with Hubble/ACS F555W, F775W, and F814W bands. The tangential critical curves overlaid in white correspond to a source redshift of $z_{s}=2.9$ (system 1, see also Borys et al. 2004), passing through the close pair of images 1.1/1.2, and 4.1/4.2. The smaller critical curves overlaid in blue correspond to a source redshift of $z_{s}\sim2$.}
\label{curves0454}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=5mm 0mm 5mm 5mm,clip]{contours0454scale.jpg}
\end{center}
\caption{2D surface mass distribution ($\kappa$), in units of the
critical density, of MACS J0454.1-0300. Contours are shown in linear units, derived from
the mass model constrained using the multiply-lensed images seen in
Figure \ref{curves0454}. Axes are in ACS pixels ($0.05 \arcsec /pixel$), and a $20\arcsec$ scale bar is overplotted.}
\label{contours0454}
\end{figure}
\subsection{MACS J0647.7+7015}
In the galaxy cluster MACS J0647.7+7015 ($z=0.591$) we note a remarkable blue background system lensed 6 times, where each lensed image is a quartet of arcs, spread across the image and between the BCGs. Additional blue faint images between images 1.5 and 1.6 (see Figure \ref{curves0647}) may also be related to this system. We did not find any record of past analysis nor redshift information and we assume the redshift of this system to be $z_{s}\sim2$ corresponding to an equivalent Einstein radius of $28\pm2 \arcsec$ enclosing a mass of $2.07\pm0.10 \times 10^{14} M_{\odot}$. This also corresponds to the higher redshift assumed for system 2, which is estimated to be a drop-out galaxy at $z_{s}\sim3.5$ helping us to constrain the fit. See Figure \ref{curves0647} for more details. The reference centre of our analysis is fixed between the two cD galaxies at: RA = 06:47:50.23, Dec = +70:14:55.37 (J2000.0), where one arcsecond corresponds to 6.67 kpc at the redshift of this cluster.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{0647curves2.jpg}
\end{center}
\caption{Galaxy cluster MACS J0647.7+7015 ($z=0.591$) imaged
with Hubble/ACS F555W and F814W bands. The critical curves overlaid in white correspond to the six images we identified as belonging to system 1, at an estimated source redshift of $z_{s}\sim2$, enclosing a critical area with an effective Einstein radius of $\sim 190$ kpc, at the redshift of this cluster. The second system consists of 3 drop-out candidate images, corresponding to a redshift of $z_{s}\sim3.5$.}
\label{curves0647}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=5mm 0mm 5mm 5mm,clip]{contours0647scale.jpg}
\end{center}
\caption{2D surface mass distribution ($\kappa$), in units of the
critical density, of MACS J0647.7+7015. Contours are shown in linear units, derived from
the mass model constrained using the multiply-lensed images seen in
Figure \ref{curves0647}. Axes are in ACS pixels ($0.05 \arcsec /pixel$), and a $20\arcsec$ scale bar is overplotted.}
\label{contours0647}
\end{figure}
\subsection{MACS J0717.5+3745}
The galaxy cluster MACS J0717.5+3745 ($z=0.55$) was analysed recently by Zitrin et al. (2009a). This very X-ray luminous cluster is the denser north-western region of the large-scale filament found by Ebeling et al. (2004) and it is the largest known lens, with an equivalent Einstein radius of 55 \arcsec and a mass of $7.4\pm0.5 \times 10^{14} M_{\odot}$. 13 multiply-lensed systems were used to constrain the fit as seen in Figure \ref{curves0717} taken from Zitrin et al. (2009a). Due to its very large critical area, this cluster is a great target for finding high-$z$ objects, and has been recently proposed for uncovering early stars (or ``dark stars'', Zackrisson et al. 2010).
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{0717curves2.jpg}
\end{center}
\caption{Galaxy cluster MACS J0717.5+3745 ($z=0.546$) imaged
with Hubble/ACS F555W, F606W, and F814W bands. The 34 multiply lensed
images identified by our model are numbered here. The white curve
overlaid shows the tangential critical curve corresponding to the
distance of system~1 at an estimated redshift of $z\sim2-2.5$, and which
passes through several close pairs of lensed images in this system. The larger
critical curve overlaid in red corresponds to the larger source
distance for the red dropout galaxy number ~5, at the estimated photometric
redshift of $z\sim3.5-4$. This large tangential critical curve encloses a
very large lensed region equivalent to $\sim400~kpc$ in radius at the
redshift of the cluster. Figure was originally published in Zitrin et al. 2009a.}
\label{curves0717}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=5mm 0mm 5mm 5mm,clip]{contours0717scale.jpg}
\end{center}
\caption{2D surface mass distribution ($\kappa$), in units of the
critical density, of the inner central region of MACS J0717.5+3745. Contours are shown in linear units, derived from
the mass model constrained using 34 multiply-lensed images seen in
Figure \ref{curves0717}. Axes are in ACS pixels ($0.05 \arcsec /pixel$), and a $20\arcsec$ scale bar is overplotted. Note that the central mass distribution of
is rather flat reflecting the unrelaxed appearance
of this cluster. Figure was originally published in Zitrin et al. 2009a.}
\label{contours0717}
\end{figure}
\subsection{MACS J0744.8+3927}
The luminous X-ray cluster MACS J0744.8+3927 ($z=0.698$) is the highest-$z$ cluster of this sample, comprising several obvious close pairs of
multiply-lensed galaxies which are immediately visible throughout the frame, with which we begin our modelling process (see Figure \ref{curves0744}). Systems 2 and 4 are likely drop-outs and we estimate their redshift as $z_{s}\sim3.5$ which helps to constrain our model. These systems correspond to similar lensing distance ratios thus basing this assumption. We did not find record of past analysis yet recently a spectroscopic redshift of a resolved galaxy behind this cluster was published (Jones et al. 2009; $z_{s}=2.21$, marked as ``S1'' in Figure \ref{curves0744} here). Unfortunately, no counter images were reported. As can be seen in Figure \ref{curves0744}, this arc lies within the critical curve for a source at $z_{s}\simeq 2.2$ according to our model, and therefore should be multiply-lensed and we mark other possible counter images of this galaxy. Another option, which would favour a steeper model, is that the three clumps seen in this arc are counter images of the same galaxy, though we find this option to be less probable according to the velocity map presented in Jones et al. (2009), which shows that the three clumps have different velocities. In addition, another less probable case is that the critical curve for a source at $z_{s}\simeq 2.2$ is much smaller, though this does not agree with the other assumed redshifts for the outer blue arcs (system 1 for example), which are not likely be at a higher redshift than $z_{s}\sim3$ since they are closer to the cluster center than system 2 but following the same symmetry, yet clearly seen in the F555W band. Deeper imaging in more passbands will help to uniquely solve the strong-lensing in this cluster. We derive an equivalent Einstein radius of $31\pm2 \arcsec$ for a source redshift of $z_{s}\simeq 2.2$, enclosing a mass of $3.1\pm0.1 \times 10^{14} M_{\odot}$. The reference centre of our analysis is fixed on the cD galaxy at: RA = 07:44:52.80, Dec = +39:27:26.50 (J2000.0). One arcsecond corresponds to 7.17 kpc at the redshift of this cluster.
\begin{figure}
\begin{center}
\includegraphics[width=85mm,trim=0mm 0mm 0mm 0mm,clip]{0744curves2.jpg}
\end{center}
\caption{Galaxy cluster MACS J0744.8+3927 ($z=0.698$) imaged
with Hubble/ACS F555W and F814W bands. The critical curves overlaid in red correspond to systems 2 and 4, at an estimated source redshift of $z_{s}\sim3.5$. The inner critical curve overlaid in white corresponds to a source redshift of $z_{s}\simeq 2.2$ (arc ``S1''; measured spectroscopically to be at this redshift by Jones et al. 2009). We mark possible counter images of this arc and other multiply-lensed images we uncovered throughout the frame.}
\label{curves0744}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=5mm 0mm 5mm 5mm,clip]{contours0744scale.jpg}
\end{center}
\caption{2D surface mass distribution ($\kappa$), in units of the
critical density, of MACS J0744.8+3927. Contours are shown in linear units, derived from
the mass model constrained using the multiply-lensed images seen in
Figure \ref{curves0744}. Axes are in ACS pixels ($0.05 \arcsec /pixel$), and a $20\arcsec$ scale bar is overplotted.}
\label{contours0744}
\end{figure}
\subsection{MACS J0911.2+1746}
In the galaxy cluster MACS J0911.2+1746 ($z=0.505$) not many prominent arcs are seen, in agreement with the small critical area we derived ($\theta_{e}=11^{+3}_{-1} \arcsec$, for a source redshift of $z_{s}\sim2$). We found no record of previous SL analysis of this cluster nor arcs redshift information. Only a few arcs are seen throughout the frame, from which we are able to match 7 multiply-lensed images which are iteratively incorporated into the model, belonging to 3 distant galaxies. Two of these galaxies (systems 1 and 2) are to our estimation at $z_{s}\sim2$. We find that the mass enclosed within the critical curves (for a source redshift of $z_{s}\sim2$) is $0.28^{+0.02}_{-0.01} \times 10^{14} M_{\odot}$. This critical curve and the enclosed mass are both the smallest of the current sample. See Figures \ref{curves0911} and \ref{contours0911} for more details.
The reference centre of our analysis is fixed near the centre of the ACS frame at: RA = 09:11:10.30, Dec = +17:46:39.33 (J2000.0), where one arcsecond corresponds to 6.17 kpc at the redshift of this cluster.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{0911curves2.jpg}
\end{center}
\caption{Galaxy cluster MACS J0911.2+1746 ($z=0.505$) imaged
with Hubble/ACS F555W and F814W bands. The 7 multiply-lensed
images identified by our model are numbered here. The white curve
overlaid shows the tangential critical curve corresponding to the
distance of system~1 at an estimated redshift of $z\sim2$. This cluster comprises the smallest Einstein radius and mass of the analysed sample. The area enclosed within the marked critical curve corresponds to an equivalent Einstein radius of 68 kpc at the cluster redshift.}
\label{curves0911}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=5mm 0mm 5mm 5mm,clip]{contours0911scale.jpg}
\end{center}
\caption{2D surface mass distribution ($\kappa$), in units of the
critical density, of MACS J0911.2+1746. Contours are shown in linear units, derived from
the mass model constrained using the multiply-lensed images seen in
Figure \ref{curves0911}. Axes are in ACS pixels ($0.05 \arcsec /pixel$), and a $20\arcsec$ scale bar is overplotted.}
\label{contours0911}
\end{figure}
\subsection{MACS J1149.5+2223}
This cluster was first analysed by us (see Zitrin \& Broadhurst 2009) noting several large blue spiral galaxy
images which are clearly visible near the central brightest cluster galaxy
(Figure \ref{curves11495}). Shortly after, spectroscopic redshifts were published (Smith et al. 2009) in full agreement with the predictions of our analysis: we assumed $z\simeq1.5$ for the spiral-galaxy (system 1), and $z\simeq2$ for the outer blue images (system 3 in Figure \ref{curves11495}), which were later verified to be at $z=1.49$ and $z=1.89$, respectively. Many other faint lensed galaxies are also visible, most of which we have been able to securely identify as belonging to 10 sets of multiply-lensed background galaxies (Figure
\ref{curves11495}); see
Zitrin \& Broadhurst (2009) for more details. We find that the critical curves for a source redshift of $z\simeq2$ enclose a mass of $1.71\pm0.20 \times 10^{14} M_{\odot}$, and have an equivalent Einstein radius of $27\pm3 \arcsec$.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{11495curves2.jpg}
\end{center}
\caption{Large scale view of the multiply-lensed galaxies identified by our
model in MACS J1149.5+2223 ($z=0.544$). In addition to the large spiral galaxy system~1, many other
fainter sets of multiply lensed galaxies are uncovered by our
model. The white curve overlaid shows the tangential critical curve
corresponding to the lensing distance of system~1. The larger critical curve
overlaid in blue corresponds to the average distance of the fainter
systems, passing through close pairs of lensed images in systems 2 and
3. This large scale elongated ``Einstein ring'' encloses
a very large critically lensed region equivalent to $170~kpc$ in radius. For this cluster one
arcsecond corresponds to 6.4 kpc, with the standard cosmology. Figure was originally published in Zitrin \& Broadhurst 2009.}
\label{curves11495}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=5mm 0mm 5mm 5mm,clip]{contours11495scale.jpg}
\end{center}
\caption{2D surface mass distribution ($\kappa$), in units of the critical density, of MACS J1149.5+2223. Contours are shown in linear units, derived from the mass model constrained using 33 multiply lensed images seen in Figure \ref{curves11495}. Note, the central mass distribution is shallow, and rounder in shape than the critical curves. Figure was originally published in Zitrin \& Broadhurst 2009. Axes are in ACS pixels ($0.05 \arcsec /pixel$), and a $20\arcsec$ scale bar is overplotted.}
\label{contours11495}
\end{figure}
\subsection{MACS J1423.8+2404}
The galaxy cluster MACS J1423.8+2404 ($z=0.543$) was recently analysed by Limousin et al. (2009), finding 3 sets of multiply-lensed galaxies. System 1 here was spectroscopically measured by them to be at $z_{s}=2.84$ and system 2 here was spectroscopically measured by them to be at $z_{s}=1.78$. We incorporate these images in order to fully constrain our model, with an additional locally-lensed arc found here (system 3). The corresponding critical curves for a source redshift of $z_{s}\sim2$ are overlaid in Figure \ref{curves1423}, enclosing a mass of $1.3\pm0.40 \times 10^{14} M_{\odot}$ and yielding an effective Einstein radius of $\theta_{e}=20\pm2 \arcsec$. This Einstein radius is in agreement with that quoted by Limousin et al. (2009), yet the mass is lower than quoted by them, but fully agrees with the Mass/Critical-area relation seen in Figure \ref{reme}, which supports our measurement.
The reference centre of our analysis is fixed near the centre of the ACS frame at: RA = 14:23:48.05, Dec = +24:05:00.23 (J2000.0), where one arcsecond corresponds to 6.40 kpc at the redshift of this cluster.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{1423curves2.jpg}
\end{center}
\caption{Galaxy cluster MACS J1423.8+2404 ($z=0.543$) imaged
with Hubble/ACS F555W and F814W bands. The multiple-images used in our model are marked on the image. System 1 was spectroscopically measured by Limousin et al. (2009) to be at $z_{s}=2.84$ and system 2 was spectroscopically measured by them to be at $z_{s}=1.78$. On the image we overlay the critical curves for a source redshift of $z_{s}\sim2$, enclosing a critical area with equivalent Einstein radius of $\sim130$ kpc at the cluster redshift.}
\label{curves1423}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=5mm 0mm 5mm 5mm,clip]{contours1423scale.jpg}
\end{center}
\caption{2D surface mass distribution ($\kappa$), in units of the
critical density, of MACS J1423.8+2404. Contours are shown in linear units, derived from
the mass model constrained using the multiply-lensed images seen in
Figure \ref{curves1423}. Axes are in ACS pixels ($0.05 \arcsec /pixel$), and a $20\arcsec$ scale bar is overplotted.}
\label{contours1423}
\end{figure}
\subsection{MACS J2129.4-0741}
In the galaxy cluster MACS J2129.4-0741 ($z=0.589$) various arcs are seen throughout the image. One spectacular system consists of 6 lensed images in similar colours as the cluster members (system 1; see Figure \ref{curves2129}). We use this system and the large arc (system 2) further away from the centre to fully constrain the model. Estimating that the outer blue arc (system 2) is at $z_{s}\sim2-2.5$, correspondingly the six-times lensed galaxy of system 1 is at a redshift of $z_{s}\sim1-1.5$. The outer critical curve, corresponding to system 2 encloses an area of an effective radius of $\theta_{e}=37\pm2 \arcsec$ and a mass of $3.4^{+0.6}_{-0.3} \times 10^{14} M_{\odot}$. The reference centre of our analysis is fixed at: RA = 21:29:26.123, Dec = -07:41:27.28 (J2000.0), where one arcsecond corresponds to 6.66 kpc at the redshift of this cluster.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{2129curves2.jpg}
\end{center}
\caption{Galaxy cluster MACS J2129.4-0741 ($z=0.589$) imaged
with Hubble/ACS F555W and F814W bands. The large white critical curve corresponds to system 2, at an estimated redshift of $z_{s}\sim2$, enclosing a critical area of an effective Einstein radius of $\sim250$ kpc at the redshift of this cluster. Comprising 6 remarkable images in the centre of the image, system 1 is at a redshift of $z_{s}\sim1$, whose critical curve is overlaid in blue.}
\label{curves2129}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=5mm 0mm 5mm 5mm,clip]{contours2129scale.jpg}
\end{center}
\caption{2D surface mass distribution ($\kappa$), in units of the
critical density, of MACS J2129.4-0741. Contours are shown in linear units, derived from
the mass model constrained using the multiply-lensed images seen in
Figure \ref{curves2129}. Axes are in ACS pixels ($0.05 \arcsec /pixel$), and a $20\arcsec$ scale bar is overplotted.}
\label{contours2129}
\end{figure}
\subsection{MACS J2214.9-1359}
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{2214curves2.jpg}
\end{center}
\caption{Galaxy cluster MACS J2214.9-1359 ($z=0.503$) imaged
with Hubble/ACS F555W and F814W bands. The critical curve overlaid in white corresponds to system 1, enclosing a critical area of an effective Einstein radius of $\sim140$ kpc at the redshift of the cluster, for an estimated source redshift of $z_{s}\sim2$.}
\label{curves2214}
\end{figure}
The galaxy cluster MACS J2214.9-1359 ($z=0.503$) comprises several prominent large blue arcs, 4 of which we match as system 1 (see Figures \ref{curves2214} and \ref{rep2214}). This system might include an additional image, with mirror symmetry to 1.4, if the nearby foreground object is prominently included, yet we do not use this image for constraining our model. We estimate the redshift of systems 1 and 2 as $z_{s}\sim2$, which encloses a mass of $1.25\pm0.10 \times 10^{14} M_{\odot}$ in a critical area with an effective Einstein radius of $\theta_{e}=23\pm2 \arcsec$.
The reference centre of our analysis is fixed at: RA = 22:14:56.59, Dec = -14:00:17.23 (J2000.0), where one arcsecond corresponds to 6.16 kpc at the redshift of this cluster.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=5mm 0mm 5mm 5mm,clip]{contours2214scale.jpg}
\end{center}
\caption{2D surface mass distribution ($\kappa$), in units of the
critical density, of MACS J2214.9-1359. Contours are shown in linear units, derived from
the mass model constrained using the multiply-lensed images seen in
Figure \ref{curves2214}. Axes are in ACS pixels ($0.05 \arcsec /pixel$), and a $20\arcsec$ scale bar is overplotted.}
\label{contours2214}
\end{figure}
\section{Discussion}
Here we examine the SL properties of the whole sample. The Einstein radii, enclosed mass, and high-$z$ magnifications are listed in Table \ref{results_table}, and discussed in the following subsections.
\begin{table*}
\caption{Summary and results of the SL analysis. Part of the following data are based on Ebeling et al. (2007; see also Table \ref{sample}). \emph{Column 3:} Effective Einstein radius, in arcseconds. Simply the square root of the area enclosed within the critical curves divided by $\pi$; \emph{Column 4:} Effective Einstein radius, in kpc, for $z_{s}\sim2$. An uncertainty of $\Delta z\pm0.5$ in the estimated source redshift results in a typical $\sim10\%$ deviation in the lensing distance ratio, and usually up to such deviation also in the Einstein radii. \emph{Column 5:} Mass enclosed within the critical curves, in $10^{14} M_{\odot}$. The quoted errors correspond to the different models. They do not include the source redshift uncertainty where exists. An uncertainty of $\Delta z\sim$0.5 can cause a typical mass uncertainty of $^{+50}_{-10} \%$; \emph{Column 6:} $M/L_{B}$, in $(M/L)_{\odot}$, fluxes were converted to luminosities using the LRG template described
in Ben\'itez et al. (2009); \emph{Column 7:} Lower limit of the highly magnified ($> \times10$) area for a high-$z$ source at $z_{s}\sim8$, in square arcminutes; \emph{Column 8:} image-plane RMS, in arcseconds; for more details see also Table \ref{sample}.}
\label{results_table}
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|c|c|}
\hline\hline
MACS & $z$ & $\theta_{e}$'' & $\simeq r_{e}$ & Mass & $(M/L_{B})_{\odot}$ & $\mu >10$ & $RMS$ & $\sigma$ & $L_{x}$ & $KT (kev)$ & M.C.E.\\
& & &(kpc)&($10^{14} M_{\odot}$)&&&image & km s$^{-1}$&(Chandra)&&\\
\hline
J0018.5+1626 & 0.545 & $24\pm2$& 154& $1.46\pm0.1$ & 278& $3.1\sq\arcmin$&$1.5\arcsec$ &$1420^{+140}_{-150}$ & $19.6\pm{0.3}$ & $9.4\pm{1.3}$ & 3\\
J0025.4-1222 & 0.584 & $30\pm2$& 199& $2.42^{+0.10}_{-0.13}$ & 171&$3.1\sq\arcmin$&$2\arcsec$ &$740^{+90}_{-110}$ & $8.8\pm{0.2}$ & $7.1\pm{0.7}$ & 3\\
J0257.1-2325 & 0.505 & $39\pm2$& 241 & $3.35^{+0.58}_{-0.10}$ & 435&$2.9\sq\arcmin$&$1\arcsec$ &$970^{+200}_{-250}$ & $ 13.7\pm{0.3}$ & $10.5\pm{1.0}$& 2\\
J0454.1-0300 & 0.538 & $13^{+3}_{-2}$& 83 & $0.41^{+0.03}_{-0.01}$ & 159&$2.6\sq\arcmin$&$2.7\arcsec$ &$1250^{+130}_{-180}$ & $16.8\pm{0.6}$ & $7.5\pm{1.0}$ & 2\\
J0647.7+7015 & 0.591 & $28\pm2$& 187 & $2.07\pm0.1$ & 256&$2.8\sq\arcmin$&$3\arcsec$ &$900^{+120}_{-180}$ & $15.9\pm{0.4}$ & $11.5\pm{1.0}$& 2\\
J0717.5+3745 & 0.546 & $55\pm3$& 353 & $7.4\pm0.5$ & 370&$3.3\sq\arcmin$&$2.2\arcsec$ &$1660^{+120 }_{-130}$ & $24.6\pm{0.3}$ & $11.6\pm{0.5}$& 4\\
J0744.8+3927 & 0.698 & $31\pm2$&222 & $3.1\pm0.1$ & 380& $2.5\sq\arcmin$&$1\arcsec$ & $1110^{+130 }_{-150}$ & $22.9\pm{0.6}$ & $8.1\pm{0.6}$& 2\\
J0911.2+1746 & 0.505 & $11^{+3}_{-1}$&68 & $0.28^{+0.02}_{-0.01}$ & 101 &$1.4\sq\arcmin$& $1.5\arcsec$ & $1150^{+260 }_{-340}$ & $7.8\pm{0.3}$ & $8.8\pm{0.7}$& 4\\
J1149.5+2223 & 0.544 & $27\pm3$ & 173& $1.71\pm0.20$ & 190& $2.6\sq\arcmin$ & $1.8\arcsec$&$1840^{+120 }_{-170}$ & $17.6\pm{0.4}$ & $9.1\pm{0.7}$ & 4\\
J1423.8+2404 & 0.543 & $20\pm2$& 128 & $1.3\pm0.40$ & 194& $0.7\sq\arcmin$& $3\arcsec$ & $1300^{+120}_{-170}$ & $16.5\pm{0.7}$ & $7.0\pm{0.8}$ & 1\\
J2129.4-0741 & 0.589 & $37\pm2$& 246& $3.4^{+0.6}_{-0.3}$& 314&$2.6\sq\arcmin$& $3.4\arcsec$ &$1400^{+170}_{-200}$ & $15.7\pm{0.4}$ & $8.1\pm{0.7}$ & 3\\
J2214.9-1359 & 0.503 & $23\pm2$ &142 & $1.25\pm0.10$ & 153&$1.4\sq\arcmin$& $1\arcsec$ & $1300^{+90}_{-100}$ & $14.1\pm{0.3}$ & $8.8\pm{0.7}$ & 2\\
\hline
\end{tabular}
\end{center}
\end{table*}
\subsection{The Einstein Radius Distribution}\label{er}
The 12 clusters have effective Einstein radii in the range, 11$\arcsec$ to 55$\arcsec$, with a median value of $27.5\arcsec$, (and a mean value of $28 \pm 3.6\arcsec$) assuming a fixed source redshift of $z_s\simeq2$, see Figure \ref{hist}. This corresponds to a range of physical
radii, 68 kpc to 353 kpc, with a median value of 180 kpc, when transforming to the redshift of each lens. In the hierarchical model large-scale perturbations collapse recently and thus should be found relatively locally.
This sample includes very large Einstein radii, exceeding even the
most impressive lenses previously studied at lower redshifts as analysed by Broadhurst \& Barkana (2008). More recently, Richard et al. (2009b) carefully studied the lensing properties of 20 mainly undisturbed clusters at lower-$z$, selected to have SL features in Hubble data and measured X-ray data, for which their mean Einstein radius is $\theta_{e}=14.45 \arcsec$, a factor of $\sim$two smaller than our sample. This difference reflects
the larger masses of the MACS sub-sample that we studied, which is purely X-ray flux selected and restricted to $z>0.5$. This MACS sample therefore does not suffer from a lensing-selection bias, but the effect of projection bias must be taken into account when evaluating 2D lensing observations with 3D mass profiles predicted by theory (e.g., Hennawi et al. 2007, Oguri \& Blandford 2009, Sereno, Jetzer \& Lubini 2010), as discussed in \S \ref{cdm} below.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{histf.jpg}
\end{center}
\caption{Einstein Radius distribution. This histogram shows the number of clusters per bin, where 6 equally-spaced bins were used to divide the range $11 \arcsec$ to $55 \arcsec$. The solid curve is the expected distribution of Einstein radii as calculate in \S \ref{cdm} for the $\Lambda$CDM model, incorporating both the scatter on the $L_x-M$ relation and the projection bias from lensing. We multiply the resulting theoretical $dN/d\theta_e$ curve by the width of the bins to normalise it. The observed distribution is skewed to larger Einstein radii than predicted. The median values of both the observed and the theoretical distributions are plotted on the histogram in a blue dashed line, and a black dash-dotted line, respectively.}
\label{hist}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{histRennan1f.jpg}
\end{center}
\caption{Comparison of the cumulative distribution of observed Einstein Radii (blue solid stairs) and the theoretical distribution predicted by the $\Lambda$CDM model (black solid curve) described in \S \ref{cdm}.The median values of both the observed and the theoretical cumulative distributions are plotted on the histogram in a blue dashed line, and a black dash-dotted line, respectively.}
\label{histR}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{MassVsRe.jpg}
\end{center}
\caption{Mass enclosed within the critical curves as a function of the effective Einstein radii. The theoretical relation for symmetric mass distribution, $M \propto r_{e}^{2}$, is tightly followed by the data, indicating that in general the central mass distributions of these clusters are well behaved.}
\label{reme}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{MoverLVsthethae.jpg}
\end{center}
\caption{Mass-to-light ratio enclosed within the critical curves as a function of the Einstein radius (in arcseconds).
A general trend of increasing $M/L$ with the system scale size is apparent.
}
\label{ml}
\end{figure}
Most of these clusters have resisted analysis by strong lensing despite the long availability of the Hubble data. Our success in ``cracking'' these irregular lenses and defining their critical curves is encouraging for the application of our approach to
complex unrelaxed clusters more generally. For such clusters, where the deflection fields are not symmetric, the identification of multiple images must be aided by
models that allow for flexibility in describing the mass distribution, as described in section \ref{model}, free of the
strictures imposed by the use of idealised elliptical potentials in conventional models.
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{arc0018.jpg}
\end{center}
\caption{Reproduction of system 1 in MACS J0018.5+1626 by our model. Since the arc is too faint to be noticed here we lens the middle arclet and paint in red the triplet outcome.}
\label{rep0018}
\end{figure}
\subsection{Central Mass Distributions}
The masses enclosed by the critical curves range from $2.8^{+0.2}_{-0.1}\times10^{13} M_{\odot}$ to $7.4\pm0.5\times10^{14} M_{\odot}$ with a median value of $1.9\times10^{14} M_{\odot}$, and a mean value of $2.4^{+0.8}_{-0.2}\times10^{14} M_{\odot}$ as shown in table \ref{results_table}. These
are calculated for each cluster by integrating the surface mass density distribution within the 2D model critical curve, scaled to a source redshift of $z_s\sim2$. We examine the relation between these Einstein masses, $M_{ein}$ enclosed within the critical curves, and the effective Einstein radii derived above. Theoretically this should of course simply scale as $M_{ein} \propto \theta_{e}^{2}$ for symmetric lenses. Here the lenses are not symmetric, but as can be seen in Figure \ref{reme} the quadratic relation is tightly followed, indicating that asymmetry is not very significant in terms of the mass distribution, and furthermore the measured or assumed redshifts of $z\simeq2$ for the relevant systems is a reliable estimate. In fact, we see clearly in that the 2D mass distributions are in general noticeably rounder than the critical curves which are very sensitive to substructure, as can be seen by comparing the 2D mass distributions with the critical curves, shown in Figures \ref{curves0018} - \ref{contours2214}. Further encouragement for the accuracy of SL mass estimates is presented in recent work by Meneghetti et al. (2010), showing that SL methods based on parametric modelling are accurate at the level of few percent at predicting the projected inner mass.
Richard et al. (2009b) found a mean mass of $1.95 \times 10^{14} M_{\odot}$ enclosed mass (within R$<$250 kpc) for their sample of 20 clusters, while Smith et al. (2005) examined the mass distribution of 10 X-ray selected clusters, and found a
similar mean
mass of $1.86 \times 10^{14} M_{\odot}$ (see also Richard et al. 2009b). These values are in good agreement with our sample median mass ($1.9\times10^{14} M_{\odot}$), and slightly lower than our sample mean mass which is boosted by few extremely massive clusters (e.g., MACS J0717.5+3745).
We examine the behaviour of the central mass-to-light
ratio, $M/L_{B}$, with the Einstein radius and shown in Figure \ref{ml}. The
luminosity is summed over all cluster sequence galaxies identified as described in section \ref{model}. A clear correlation is obtained such that $M/L_{B}$ increases with the Einstein radius, and with values quite typical of other massive clusters (Sarazin 1988, Rines et al. 2004, Medezinski et al. 2010).
This scaling towards higher $M/L_B$ provides confidence in our method. At the high mass end our result is
similar to the peak M/L found in the detailed weak lensing profile studies of Medezinski
et al. (2010). Measuring $M/L_{B}$ in high-$z$ clusters is of interest also for the added insight
on the amount of DM initially associated with individual galaxies as compared
with the overall cluster DM component (see also Sarazin 1988).
\begin{figure}
\begin{center}
\includegraphics[width=80mm,trim=0mm 0mm 0mm 0mm,clip]{sys1Rep0025.jpg}
\end{center}
\caption{Reproduction of system 1 in MACS J0025.4-1222 by our model, by delensing image 1.1 into the source plane, and relensing the source-plane pixels onto the image plane to accurately form the other images.}
\label{rep0025}
\end{figure}
\subsection{The Magnification Distribution}
As can be seen in Table \ref{results_table}, most of the sample clusters provide large regions ($\geq2.5 \sq\arcmin$) of highly magnified ($>\times 10$) sky, useful for the search of the first stars and galaxies (e.g., Zackrisson et al. 2010). One has to be cautious when making such a statement since unlike the critical curves and the enclosed mass, the magnification is sensitive to the mass profile which requires source redshift estimation for several sources, which we have here for only a few clusters (section \ref{model}). Still, as explained in \S 2, reasonable constraints are put on the mass profile slope by the image-plane minimisation, which generally has a broad minimum at the same point in the parameter space where the slope is also approximately correct. In addition, we are able to put a lower boundary on the high magnification area, by choosing the steeper profiles which naturally generate lower magnifications, or by assuming a boosted redshift for the main multiply-lensed system decreasing the lensing distance ratio for high-$z$ galaxies. Thus we constrain the area of high magnification for a high-$z$ source at the current limit of $z_{s}\sim8$, showing that most of this sample clusters are excellent targets for high-$z$ galaxy searches due to these large high-magnification areas and lens power. We estimate, based on previously analysed clusters (e.g., Cl 0024+1654, Zitrin et al. 2009b; MACS J1149.5+2223, Zitrin \& Broadhurst 2009; MACS J0717.5+3745, Zitrin et al. 2009a) that in practice the areas of high magnification are about 10-20\% higher than the lower limits presented in Table \ref{results_table}, with an upper limit of about 40-50\%.
\begin{figure}
\begin{center}
\includegraphics[width=60mm,trim=0mm 0mm 0mm 0mm,clip]{sys1_0257.jpg}
\end{center}
\caption{Reproduction of system 1 in MACS J0257.1-2325 by our model, by delensing image 1.1 into the source plane, and relensing the source-plane pixels onto the image plane to accurately form the other images. Note, our model predicts an extra tiny image near the core of the cD covered by its light.}
\label{rep0257}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=85mm,trim=0mm 0mm 0mm 0mm,clip]{sys1_0647_final.jpg}
\end{center}
\caption{Reproduction of system 1 in MACS J0647.7+7015 by our model, by delensing image 1.1 into the source plane, and relensing the source-plane pixels onto the image plane to accurately form the other images.}
\label{rep0647}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=70mm,trim=0mm 0mm 0mm 0mm,clip]{sys10717.jpg}
\end{center}
\caption{Reproduction of system 1 in MACS J0717.5+3745 by our model. Published originally in Zitrin et al. 2009a.}
\label{rep0717}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=70mm,trim=0mm 0mm 0mm 0mm,clip]{11495_reproduction.jpg}
\end{center}
\caption{Reproduction of system 1 in MACS J1149.5+2223 by our model. Published originally in Zitrin \& Broadhurst 2009.}
\label{rep11495}
\end{figure}
\subsection{Comparison with $\Lambda$CDM}\label{cdm}
We generally follow Broadhurst \& Barkana (2008; see also Sadeh \& Rephaeli 2008)
in constructing the theoretically predicted distribution of Einstein radii.
In particular, we adopt the NFW parameters measured by Neto et al. (2007) for
simulated halos, although we reduce $c_N$ by $10\%$ according to the results of
Duffy et al. (2008) who used the more recently measured values of the
cosmological parameters. Studies based on large numerical simulations
(Zhao et al. 2003,2009, Gao et al. 2008) have found for massive halos a rather weak
decline of $c_N$ with increasing redshift; we neglect this decline, which renders
our results (in terms of the disagreement with observations) conservative.
We then correct the concentration parameter distribution
using Hennawi et al. (2007) to obtain the effective parameters for the
population of clusters observed in projection. Note though that here
we are considering a sample selected by X-ray flux (not lensing
cross-section), so there is no lensing bias, but there still is the
projection bias; this refers to the fact that when halos are seen in
projection (at randomly distributed angles), a different effective
projected $c_N$ is measured than would be obtained from a spherical
analysis of the halo density profile. In particular, projection
increases the mean $c_N$ and adds substantially to the scatter
compared to the distribution of the spherically-measured
$c_N$. Neto et al. (2007) split their halo sample into two groups
(``relaxed'' and ``unrelaxed'') and analysed the statistics of each
group separately, so we do the same, and in the end combine the two
groups according to their relative numbers in the simulation by
Neto et al. (2007), obtaining our predicted results for the total, combined
population of halos. This can then be compared with the observed
sample, which does not select for relaxed clusters but includes both
relaxed and unrelaxed ones.
In order to approximately simulate a flux-selected sample, we need to convert $L_X$ to halo mass.
We use the power-law relation
from Reiprich \& B\"{o}hringer (1999; with an updated Hubble constant), along
with the observed scatter of $L_X$ for a given mass, which we model as
a lognormal distribution with a typical scatter that corresponds to a
factor of 1.5 . Note that this observed relation is based on
$M_{500}$, while the results from the above simulations refer to
$M_{200}$, so we must include the relation between $M_{500}$ and
$M_{200}$ which is itself a function of $c_N$. An important caveat is
that the observed $L_X-M$ relation was measured for relaxed clusters
(whose mass could be estimated from the X-ray emission assuming
hydrostatic equilibrium). We expect that unrelaxed clusters would tend
to have an unusually high $L_X$ for a given mass (e.g., in a
post-merger phase), which would insert more low-mass clusters within
the sample than we assume, leading to smaller predicted Einstein
angles, increasing the discrepancy between the theory and the
observations (again making our results conservative).
On the other hand, it should be noted that much of the scatter in the $L_X-M$ relation probably arises from cool cores in the center (since it is significantly reduced if the inner core is excised), implying a possible bias towards relaxed and more concentrated systems. However, such a bias seems less probable in our case due to the high X-ray luminosity, the relatively high redshift, and the clear spread-out appearance of most of the clusters analysed here.
The observed properties of each halo (flux, Einstein radius
distribution, etc.) change slowly with redshift, so to simplify the
calculation we approximate the clusters as all lying at a typical
redshift $z = 0.55$. However, the halo abundance changes rapidly with
redshift, so we calculate the integrated halo mass function for all
halos observed at $z>0.5$, in the fraction of the sky corresponding to
the MACS survey (around a quarter of the sky). We use the
Sheth \& Tormen (1999) formula, which accurately fits the halo mass function
(i.e., the distribution of $M_{200}$) measured in simulations.
This halo mass function was convolved with
with the $c_N$ distribution to obtain
the distribution of $M_{500}$, and then add the $L_X-M_{500}$ scatter
in order to obtain the halo mass function corresponding to selection
by a given minimum flux. Since the $L_X-M_{500}$ relation is not
strictly valid for unrelaxed clusters, we only use the power-law and
the scatter from this relation but allow some flexibility in the
normalisation, choosing an effective minimum flux that corresponds to
a predicted abundance of 12 clusters (as in the observed sample). This
is reasonable also since we are interested here in testing the
predicted $\theta_e$ distribution, not in testing the predicted
cluster abundance. Note that the flux selection does not affect the
distribution of the very highest $\theta_e$ values, since all
clusters within the survey
solid angle at $z>0.5$ that can produce the largest $\theta_e$'s will
be well above the flux threshold.
Given our predicted, X-ray selected mass function in terms of
$M_{500}$, we must convolve it with the $\theta_e$ distribution of
each halo of mass $M_{500}$. From the simulations we have the $c_N$
distribution (and thus also the $\theta_e$ and $M_{500}$ distribution)
for a given $M_{200}$, but we can invert this conditional probability
using Bayes' theorem. Note that the $M_{200}$ to $M_{500}$ conversion
should not really include the projection bias that is included in
$\theta_e$, but having two different $c_N$ distributions for each
$M_{200}$ would makes things far more complicated. Our simplification
of using a single $c_N$ distribution is reasonable since the
projection bias typically affects $M_{500}$ (for a given $M_{200}$)
only by $15\%$. Also, since $c_N$ affects both $M_{500}$ and
$\theta_e$ in the same direction, the effect on our final result
(which derives $\theta_e$ for a given $M_{500}$) is much smaller.
Our predicted $\theta_e$ distribution is compared to the observations
in Figure \ref{histR}. We compare the two cumulative distributions
$N(>\theta_e)$, i.e., the total number of clusters expected above
$\theta_e$, where the theoretical distribution has been normalised (as
noted above) to agree with the total of 12 clusters in the sample. The
observed distribution clearly lies at higher Einstein radii compared
with the predicted distribution. One indication of this is the most
discrepant cluster, J0717.5+3745 with $\theta_e = 55 \arcsec$.
The theoretical calculation yields a probability of only $3.4\%$ of
finding such a large $\theta_e$ in this $z>0.5$ cluster sample, giving
a significant (though not extreme) 2-$\sigma$ discrepancy. More
generally, the observed distribution resembles the predicted one
except offset towards higher angles by about a factor of 1.4
(comparing, for example, the two medians). The standard K-S statistic
for comparing two distributions gives a probability of $7.4\%$ that
the observed distribution is drawn from the predicted one. This is
also around a 2-$\sigma$ discrepancy, largely independent of the above
number since the K-S statistic focuses on the middle portion of the
probability distributions and is insensitive to the edges. The
relatively low significance of the K-S discrepancy is an inevitable result of the
Poisson fluctuations expected with a sample of only 12 clusters.
The discrepancy that we find, of Einstein radii that are too high by
about a factor of 1.4, appears smaller than the roughly factor of 2
difference found by Broadhurst \& Barkana (2008). This may be explained at least
partially by the difference between a lensing selected and an X-ray
selected sample. A more robust comparison would be possible with an
X-ray selected cluster sample in which the clusters have reliable
virial masses measured using weak lensing. Another caveat is that we
have compared the observed Einstein radii with the predictions based
on pure dark matter simulations. The possible effect of baryons on the
halo density profile has been difficult to study with simulations, since
hydrodynamic simulations that produce a large effect do so along with
a central baryon concentration that disagrees with observations and is
due to the ``overcooling'' problem in simulations. Analytical models
suggest that even without producing a central baryon concentration,
the complex coupled history of baryon and dark matter accretion may
allow the baryons to significantly affect the final central density
profile, potentially alleviating the discrepancy at least partially
(Barkana \& Loeb 2010). On the other hand, recent simulations suggest that the discrepancy actually increases when AGN feedback is taken into account to overcome the overcooling problem (Duffy et al. 2010).
\subsection{Effects of Uncertainties on the Results}
In this work we determine the Einstein radii and projected masses for
the 12 clusters of the $z>0.5$ MACS sample, and compare these
to simulations. We make use of our well-established modelling
method to identify many sets of multiply-lensed images which
are in turn used to constrain the models, and determine the
Einstein radii and projected masses for $z_{s}\simeq2$.
Typical uncertainties in the source redshift estimates of $\Delta
z\pm0.5$ impose only minor uncertainties on the comparison with $\Lambda$CDM.
Firstly, such an uncertainty does not enable a unique
determination of the profile slope, which we do not attempt to
fully constrain here since both the Einstein radius and the
projected mass are indifferent to the mass profile slope and
are determined by the data. Secondly, overestimating a source
redshift would in practice increase the projected mass and the observed Einstein
radius for a source at $z_s=2$, while
maintaining the relation seen in Figure \ref{reme} and thus
resulting only in growth of the discrepancy from $\Lambda$CDM
simulations which are not dependent on the observed mass. On the other hand, underestimating a source
redshift would in practice decrease the observed Einstein
radius and projected mass for a source at $z_s=2$ by less than
10\%, thus insignificantly influencing the results.
\section{Summary}
The MACS $z>0.5$ sample has proven to be of great value for many very different
studies spanning the X-ray through radio spectrum. We have extracted additional useful information from
this sample by solving the strong lensing for these clusters. Previous
discoveries from this sample include the largest lensed images of a
highly-magnified distant spiral galaxy (MACS J1149.5+2223; Zitrin \&
Broadhurst 2009, Smith et al. 2009), the largest known lens (MACS J0717.5+3745; Zitrin et al. 2009a), lensed sub-mm sources (Takata et al. 2003, Borys et al. 2004, Berciano Alba et al. 2007, 2009), and another ``bullet cluster'' (MACS J0025.4-1222; Brada\v{c} et al. 2008b).
We presented mass models and the corresponding critical curves for the 12 high-$z$ X-ray luminous clusters of the sample via strong-lensing analysis in HST/ACS images.
Several of these clusters have only a few
strongly-lensed images and corresponding low masses and Einstein radii, while
most are very massive and rich
with multiply-lensed arcs due to large critical area and high lens power. The equivalent Einstein radii of this sample range from 11$\arcsec$ to 55$\arcsec$ with a median value of $\simeq28\arcsec$ (similar to the mean value), or 180 kpc. The corresponding masses enclosed within these curves range from $2.8^{+0.2}_{-0.1}\times10^{13} M_{\odot}$ to $7.4\pm0.5\times10^{14} M_{\odot}$ with a median value of $1.9\times10^{14} M_{\odot}$, and a mean value of $2.4^{+0.8}_{-0.2}\times10^{14} M_{\odot}$. We find that the enclosed mass follows tightly the quadratic relation with the equivalent Einstein radius, indicating that the deviation from symmetry is not prominent in these clusters and that our measurements are highly accurate given the measured or assumed redshifts. In addition, we have shown that $M/L_{B}$ increases as expected with the scale size of the system, or Einstein radius.
We compare these results to the predictions of $\Lambda$CDM taking into account
projection biases and find that these predictions fall short of our measured Einstein radii
by a factor of $\simeq1.4$. In addition, the standard K-S statistic gives a
probability of only $7.4\%$ that the observed distribution is drawn from the
predicted one, corresponding to a $\sim2\sigma$ discrepancy.
It is apparent that the observed Einstein radii and implied high masses are not
likely at high redshifts in the context of the $\Lambda$CDM model
when no lensing-selection biases are involved,
showing again
possible
tension with the prediction of the standard model. Finally, as
a result of their unrelaxed mass distributions, most of these clusters cover a very large
area ($>2.5\sq \arcmin$) of high magnification ($>\times 10$) making them
primary targets for high-$z$ galaxy searches, for which substantial HST time
will shortly be forth-coming.
It should be acknowledged that theoretical predictions may clearly entail further uncertainties than those taken into account here, and a revised and more thorough analysis is needed in order to obtain higher significance results once source redshifts are available. Such comparisons should also be independently applied to other large samples, so that a significant number of clusters are compared to the $\Lambda$CDM model in order to statistically support any claimed discrepancy.
\begin{figure}
\begin{center}
\includegraphics[width=85mm,trim=0mm 0mm 0mm 0mm,clip]{sys1_2214.jpg}
\end{center}
\caption{Reproduction of system 1 in MACS J2214.9-1359 by our model, by delensing image 1.1 into the source plane, and relensing the source-plane pixels onto the image plane to accurately form the other images.}
\label{rep2214}
\end{figure}
\section*{acknowledgments}
We thank the anonymous reviewer of this work for very useful comments. AZ acknowledges Eran Ofek and Salman Rogers for their publicly available Matlab
scripts used in part of this work. AZ thanks Sharon Sadeh for useful discussions,
and Elinor Medezinski for sharing some Subaru data. This research is being
supported by the Israel Science Foundation (ISF) grant 1400/10. RB is supported by the ISF grant 823/09. ACS was developed under
NASA contract NAS 5-32865. This research is based on observations provided in the
Hubble Legacy Archive which is a collaboration between the Space Telescope Science
Institute (STScI/NASA), the Space Telescope European Coordinating Facility
(ST-ECF/ESA) and the Canadian Astronomy Data Centre (CADC/NRC/CSA).
Part of this work is based on data collected at the Subaru
Telescope, which is operated by the National Astronomical
Society of Japan.
|
\section{Introduction}
A large amount of experimental data from the Relativistic Heavy Ion
Collider (RHIC) \cite{rhic1,rhic2,rhic3,rhic4} strongly suggest that
a novel form of matter, a strongly coupled quark gluon plasma
(sQGP), may have been formed in the central region of high-energy
heavy-ion collisions. Among the experimental evidences for the
formation of sQGP are the jet quenching phenomena that include the
strong suppression of hadron spectra with large transverse momentum
in central $Au+Au$ collisions as compared to $p+p$ collisions
\cite{star-suppression,phenix-suppression} and the disappearance of
back-to-back correlations of large transverse momentum
hadrons~\cite{star-dihadron}. These jet quenching patterns observed
at RHIC are in good agreement with the theoretical predictions for
jet quenching~\cite{quenching-1992,theory1,theory2,Wang:1998ww,theory3,theory4}
or suppression of large transverse momentum hadrons as a consequence
of parton energy loss and medium modified parton fragmentation
functions induced by multiple scattering as partons propagate
through the dense medium after their initial production. The parton
energy loss and medium modification of the fragmentation functions
due to multiple parton scattering and induced gluon bremsstrahlung
are controlled by the jet transport parameter \cite{Baier:1996sk},
\begin{equation}
\hat q_R=\rho \int dq_T^2 \frac{d\sigma_R}{dq_T^2} q_T^2.
\label{eq:qhat0}
\end{equation}
or the average squared transverse momentum broadening per unit
length for a jet in color representation $R$, which is also related
to the gluon distribution density of the
medium\cite{Baier:1996sk,CasalderreySolana:2007sw} and therefore
characterizes the medium property as probed by an energetic jet.
Here we consider a picture in which the jet parton interacts with
medium particles or quasi-particles with cross section $\sigma_{R}$
and $\rho$ is the local particle density of the medium. Study of the
jet quenching phenomena therefore can provides important information
on the space-time profile of the jet transport parameter and
consequently properties of the sQGP in heavy-ion collisions.
Extensive phenomenological studies on the suppression of single
hadron spectra \cite{Vitev:2002pf,Wang:2003mm,theory1,Eskola:2004cr,Renk:2006nd} and dihadron
correlations \cite{Owens,zhanghz} at large transverse momentum have
been carried out since the first observation of the strong jet
quenching phenomena. Recent emphasis of the phenomenological studies
has been shifted to systematic analyses of the experimental data with different
jet quenching models and qualitative extraction of the jet
transport parameter
\cite{Majumder:2007ae,Bass:2008rv,Bass:2008ch,Qin:2009gw,Armesto:2009zi}.
In this paper, we will carry out phenomenological analysis of the
suppression of single hadron spectra within the higher-twist (HT)
formalism for medium modification of the parton fragmentation
functions and the next-to-leading order perturbative QCD parton
model \cite{Owens} for initial jet production. We will calculate the
single hadron suppression within three different models for the
dynamical evolution of the bulk matter: (1+1)d Bjorken model, (1+3)d
ideal hydrodynamic model and a parton cascade model. We will study
in particular the effect of collective expansion, transverse flow,
non-equilibrium and most importantly jet quenching in the hadronic phase of
the bulk matter. Since the jet transport
parameter is related to the gluon distribution density of the
bulk medium, one expects it to be different in a QGP and hadronic
matter. We will use the information on jet transport parameter in cold
nuclei extracted from phenomenological studies of deeply inelastic
scattering (DIS) off large nuclei \cite{Deng:2009qb} and extrapolate
to a hot hadronic gas assuming that the gluon distribution in a
nucleon in cold nuclei is the same as in a hot hadronic gas. This
allows us to focus on the values of the jet transport parameter in a
QGP that can be accommodated by the experimental data on single
hadron suppression and the effect of dynamic evolution of the bulk
medium.
The rest of the paper is organized as follows. In the next section, we will review the energy
loss and modified fragmentation functions in hot nuclear medium within the HT approach and their
dependence on the space-time profile of the jet transport parameter $\hat q_{R}$. We then
describe the next-to-leading order (NLO)
pQCD parton model for single hadron production in heavy-ion collisions in Sec.~\ref{sec:b}.
In Sec .~\ref{sec:c}, we will introduce three different dynamic evolutions for the bulk medium. The numerical
calculations and phenomenological studies of the experimental data on single hadron suppression
and extraction of the jet transport parameters within each dynamic model are carried out in Sec .~\ref{sec:d}.
Finally, we conclude in Sec.~\ref{sec:e} with a summary.
\section{Energy loss and modified fragmentation functions}\label{sec:a}
\begin{figure}
\centerline{\includegraphics[width=6cm]{fig1.eps}} \caption{A
typical Feymann diagram for quark-gluon re-scattering processes in
DIS with three possible cuts, central(C), left(L), and right(R).}
\label{fig:scatter-in-nucleon}
\end{figure}
Within the generalized factorization of twist-four processes,
one can calculate the nuclear modification of fragmentation functions and the energy loss of a
quark propagating through nuclear matter after it is produced via a hard
process in DIS off a nuclear target \cite{guoxiaofeng,benwei-nuleon}. Within such HT approach,
the medium modification to the quark fragmentation functions in DIS off a nuclear target is caused by
multiple scattering between the struck quark and the nuclear medium on the quark's propagation path
and the induced gluon bremsstrahlung as illustrated in Fig.~(\ref{fig:scatter-in-nucleon}).
The medium modified quark fragmentation function,
\begin{eqnarray}
\tilde{D}_{q}^{h}(z_h,Q^2) &=&
D_{q}^{h}(z_h,Q^2)+\frac{\alpha_s(Q^2)}{2\pi}
\int_0^{Q^2}\frac{d\ell_T^2}{\ell_T^2} \nonumber\\
&&\hspace{-0.7in}\times \int_{z_h}^{1}\frac{dz}{z} \left[ \Delta\gamma_{q\rightarrow qg}(z,x,x_L,\ell_T^2)D_{q}^h(\frac{z_h}{z})\right.
\nonumber\\
&&\hspace{-0.2 in}+ \left. \Delta\gamma_{q\rightarrow
gq}(z,x,x_L,\ell_T^2)D_{g}^h(\frac{z_h}{z}) \right] ,
\label{eq:mo-fragment}
\end{eqnarray}
takes a form very similar to the vacuum bremsstrahlung corrections that leads to the
Dokshitzer-Gribov-Lipatov-Altarelli-Parisi (DGLAP) \cite{dglap} evolution equations in pQCD for fragmentation
functions, except that the medium modified splitting functions,
\begin{eqnarray}
\Delta\gamma_{q\rightarrow qg}(z,x,x_L,\ell_T^2)&=&
[\frac{1+z^2}{(1-z)_{+}}T_{qg}^{A}(x,x_L) \nonumber\\
&&\hspace{-0.7in} +\delta(1-z)\Delta
T_{qg}^{A}(x,x_L)]\frac{2\pi\alpha_sC_A}{\ell_T^2N_cf_q^A(x)};
\label{eq:delta-gamma} \\
\Delta\gamma_{q\rightarrow gq}(z,x,x_L,\ell_T^2)&=&\Delta\gamma_{q
\rightarrow qg}(1-z,x,x_L,\ell_T^2),
\label{eq:equal}
\end{eqnarray}
depend on the properties of the medium through the twist-four quark-gluon correlations inside the nucleus,
\begin{eqnarray}
T^{A}_{qg}(x,x_L) &=& \int \frac{d^2q_T}{(2\pi)^2}\int
\frac{dy^-}{2\pi}d\xi^{-}dy_2^- d^2\xi_T
e^{i(x+x_{L})p^{+}y^{-}} \nonumber \\
&&\hspace{-1.0in}\times e^{ix_Tp^{+}\xi^{-}-i\vec{q}_T\vec{\xi}_T}
\left\{ (1-e^{-ix_Lp^+y_2^-}) (1-e^{-ix_Lp^+(y^{-}-y_1^-)}) \right.\nonumber \\
&&\hspace{-0.9in}+\frac{1-z}{2} \left[ e^{-ix_Lp^+y_{2}^{-}} (1-e^{-ix_Lp^+(y^{-}-y^-_1)})\right. \nonumber \\
&&\hspace{-0.9in}+\left.\left. e^{-ix_{L}p^{+}(y^{-}-y_{1}^{-})}(1-e^{ix_Lp^{+}y_{2}^{-}}) \right] \right\}\theta(-y_2^-)\theta(y^{-}-y_1^{-})\nonumber \\
&&\hspace{-0.9in} \times\langle A | \bar{\psi}_q(0)\,
\frac{\gamma^+}{2}F_{\sigma}^{\ +}(y_{2}^{-})\,
F^{+\sigma}(y_{1}^-,\xi_T)\,\psi_q(y^-) | A\rangle ,
\label{eq:qgmatrix}
\end{eqnarray}
where $\xi=y_1 - y_2$, $y$, $y_1$ and $y_2$ are space-time
coordinates associated with the quark and gluon fields as
illustrated in Fig.~\ref{fig:scatter-in-nucleon}. Here we include contributions beyond the helicity
approximation \cite{benwei-nuleon} and consider only the twist-four matrices that are enhanced
by the large nuclear size. The relative transverse coordinate $\xi_T$ is the Fourier conjugate of the
transverse momentum $q_T$ in the gluon distribution function. The momentum
fraction
\begin{eqnarray}
x_L=\frac{\ell_T^2}{2z(1-z)p^{+}q^{-}}, \label{eq:x_L}
\end{eqnarray}
is the total (+component) longitudinal momentum transfer from the target
to the propagating quark and radiated gluon with longitudinal momentum $zq^{-}$ and
transverse momentum $\ell_T$. We impose kinematic
constraints $x_{L}\leq 1$ and $\ell_{T}\leq \min [zE,(1-z)E]$
in the integration in Eq.~(\ref{eq:mo-fragment}).
A similar (+) longitudinal momentum transfer
\begin{eqnarray}
x_T=\frac{q_T^2-2q_T\cdot \ell_T}{2zp^{+}q^{-}}, \label{eq:x_T}
\end{eqnarray}
is also provided by the initial gluon with transverse momentum $q_{T}$.
In the modified splitting function,
\begin{eqnarray}
f_{q}^{A}(x)=\int\frac{dy^-}{2\pi}e^{ix p^+y^-}\langle
A|\bar{\psi}_q(0)\frac{\gamma^+}{2}\psi_q(y^-)|A\rangle
\end{eqnarray}
is the quark distribution function which represents the production rate of the initial quark
carrying $x=Q^2/2p^{+}q^{-}$ (the Bjorken variable) fraction of the nucleon (+) longitudinal
momentum in DIS. The quark-gluon matrix element,
\begin{eqnarray}
\Delta T_{qg}^{A}(x,x_L) &=& \int_0^{1}dz\frac{1}{1-z}\left[2T_{qg}^{A}(x,x_L)|_{z=1}\right. \nonumber\\
&&\hspace{-0.7in}\left.-(1+z^2)T_{qg}^{A}(x,x_L)\right]
\end{eqnarray}
comes from the virtual correction to the induced gluon bremsstrahlung process.
If we define the quark energy loss as the energy carried by the radiative gluon in the multiple
scattering processes, the total energy loss for a propagating quark in a deeply inelastic scattering (DIS)
off a large nucleus is
\begin{eqnarray}
\frac{\Delta E}{E} = \frac{C_A\alpha_s^{2}}{N_c}\int_0^{Q^2}
\frac{d\ell_T^2}{{\ell_T^4}}
\int_0^1 dz (1+z)^2
\frac{T_{qg}^{A}(x,x_L)}{f_q^A(x)}.
\label{eq:dis0}
\end{eqnarray}
\begin{figure}
\centerline{\includegraphics[width=6cm]{fig2.eps}} \caption{Feymann
diagram for induced gluon radiation in hot QGP medium that
contributes the quark energy loss.} \label{fig:scatter-in-QGP}
\end{figure}
One can extend the results for medium modified parton fragmentation
functions in DIS to the case of quark propagation in a hot medium
(either QGP or hot hadronic matter) after it is produced through
initial hard processes before the formation of the QGP
\cite{CasalderreySolana:2007sw,Majumder:2007ae} in high-energy
heavy-ion collisions. In this case, we can assume the quark is
produced through a hard process such as the virtual-photon-nucleus
collisions in Fig.~\ref{fig:scatter-in-QGP} and will interact with
partons in the hot medium which is independent of the initial quark
production process. Therefore, we can replace the nucleus state as a
product of the initial nucleus state and a thermal ensemble of
quasi-particle states in the hot medium,
\begin{equation}
|A\rangle \rightarrow \int\frac{d^3k}{(2\pi)^32k^+}\psi_{k}(y)e^{ik\cdot y}|k\rangle\otimes|A\rangle,
\end{equation}
and consider only the final state interaction between
the produced quark and the bulk medium as illustrated in Fig.~\ref{fig:scatter-in-QGP}.
Here $f(k,y)=|\psi_{k}(y)|^{2}$ is the
local phase-space density of the quasi-particle distribution in the medium. We also
neglect multiple particle correlations inside the hot medium which can be included through
an effective gluon distribution density \cite{CasalderreySolana:2007sw}.
Under these assumptions, the quark-gluon correlation function in the HT approach to
multiple scattering will take a factorized form,
\begin{eqnarray}
\frac{T^{A}_{qg}(x,x_L)}{f_q^A(x)} &=&\frac{N_{c}^{2}-1}{4\pi\alpha_sC_{R}}\frac{1+z}{2} \int dy^{-}
2 \sin^{2}\left[\frac{y^{-}\ell_{T}^{2}}{4Ez(1-z)}\right]
\nonumber \\
&&\hspace{0.0 in}\times\left[\hat{q}_R(E,x_L,y)+c(x,x_{L}) \hat{q}_R(E,0,y)\right] \, , \label{eq:corr2}
\end{eqnarray}
where $c(x,x_{L})=f_{q}^{A}(x+x_{L})/f_{q}^{A}(x)$ and $\hat{q}_R$ is the generalized jet transport parameter,
\begin{eqnarray}
\hat q_R(E,x_L,y)&=&\frac{4\pi^2\alpha_sC_{R}}{N_{c}^{2}-1}
\int\frac{d^3k}{(2\pi)^3} f(k,y)\nonumber \\
&&\hspace{-0.2in}\times\int \frac{d^2q_T}{(2\pi)^2}
\phi_k(x_T+x_L,q_T), \label{eq:qhat-phi}
\end{eqnarray}
which depends on the fractional (+) longitudinal momentum transfer
$x_L$ from medium gluons \cite{note}. Here $\phi_k(x,q_T)$ is the transverse momentum dependent gluon
distribution function from the quasi-particle in the medium,
\begin{eqnarray}
\phi_k(x,q_T) &=& \int \frac{d\xi^-}{2\pi k^+} d^2\xi_T\,
e^{ix k^+\xi^- - i{\bf q}_T\cdot {\bf \xi}_T} \nonumber \\
&&\hspace{-0.1in}\times\langle k| F_{\sigma
}^+(0)F^{+\sigma}(\xi^-,{\bf \xi}_T)|k\rangle\, .
\end{eqnarray}
The two terms in the square brackets of Eq.~(\ref{eq:corr2}) correspond to two
different processes in the multiple scattering. The first term involves (+) longitudinal momentum
transfer $x_{L}$ between the propagating parton and the medium due to final gluon production.
It contains what is normally defined as pure elastic energy
loss~\cite{elastic}. The second term that is proportional to the normal (or special) transport parameter
$\hat{q}_{R}(E,y)=\hat{q}_{R}(E,x_L=0,y)$ corresponds to pure
radiative processes. In this paper, we assume $x \gg x_L, x_{T}$ and only focus on
the contribution of radiative energy loss. Therefore, $c(x,x_{L})\approx 1$,
$\hat{q}_{R}(E,x_L,y)\approx \hat{q}_{R}(E,0,y) \equiv \hat{q}_{R}(E,y)$. Given the space-time
profile of the jet transport parameter $\hat{q}_{R}(E,y)$, one should be able to calculate the modified
fragmentation function according to Eq.~(\ref{eq:mo-fragment}). The corresponding quark energy loss
in Eq.~(\ref{eq:dis0}) can be expressed as
\begin{eqnarray}
\frac{\Delta E}{E} &=& \frac{N_{c}\alpha_s}{\pi} \int dy^-dz
{d\ell_\perp^2}
\frac{(1+z)^3}{\ell_T^4} \nonumber \\
&& \hspace{-0.5in}\times \hat q_R(E,y)
\sin^2\left[\frac{y^-l_T^2}{4Ez(1-z)}\right], \label{eq:de-twist}
\end{eqnarray}
in terms of the jet transport parameter. The jet transport parameter for a gluon is $9/4$
times of a quark and therefore the radiative energy loss of a gluon jet is also $9/4$ times
larger than that of a quark jet. In our following calculations of the hadron spectra in heavy-ion
collisions, we will use the same formalism for both quark and gluon modified fragmentation
functions but with jet transport parameters that differ by a factor of 9/4 for quark and gluon jets.
\section{Single hadron spectra within NLO pQCD parton model}\label{sec:b}
In this paper, we employ the NLO pQCD parton model for the initial
jet production spectra which has been shown to work well for large
$p_T$ hadron production in high energy nucleon-nucleon
collisions~\cite{owens1987}. In a factorized form, the inclusive
particle production cross section in $p+p$ collisions can be
calculated as a convolution of parton distribution functions inside
the proton, elementary parton-parton scattering cross sections and
parton fragmentation functions,
\begin{eqnarray}
\frac{d\sigma^h_{pp}}{dyd^2p_T}&=&\sum_{abcd}\int
dx_adx_bf_{a/p}(x_a,\mu^2)f_{b/p}(x_b,\mu^2) \nonumber \\
&&\hspace{-0.1in}\times\frac{d\sigma}{d\hat{t}}(ab\rightarrow cd)
\frac{D_{h/c}^0(z_{c},\mu^2)}{\pi z_{c}}+\mathcal {O}(\alpha_s^3),
\label{eq:pp}
\end{eqnarray}
where $d\sigma/d\hat{t}(ab\rightarrow cd)$ are elementary parton
scattering cross sections at leading order (LO) $\alpha_s^2$. The
next-to-leading order (NLO) contributions including
$2\rightarrow3$ tree level contributions and 1-loop virtual corrections
to $2\rightarrow2$ tree processes. We refer to
Ref.~\cite{Owens} for a detailed description.
The proton parton distribution functions (PDF) $f_{a/p}(x_a,\mu^2)$ are
given by the CTEQ6M parametrization \cite{distribution} where $x_a$ is the fractional
momentum carried by the beam partons. The fragmentation function (FF) $D^0_{h/c}(z_c,\mu^2)$
of parton $c$ into hadron $h$, with $z_c$ the momentum fraction of a parton jet carried by a produced
hadron are given by the updated AKK parametrization~\cite{akk08}, which is recently improved
with new input from the RHIC data, especially for $\pi^\pm$, $K^\pm$,
$p/\bar{p}$, $K_S^0$ and $\Lambda/\bar{\Lambda}$ particles.
We use a NLO Monte Carlo based program~\cite{Owens} to calculate the
single hadron spectra in our study. In this NLO program, the
factorization scale and the renormalization scale are chosen to be the
same (denoted as $\mu$) and are all proportional to the transverse momentum of the final hadron $p_T$.
As pointed out in Ref.~\cite{zhanghz}, the calculated single inclusive pion spectra
within the NLO pQCD parton model for $p+p$ collisions agree well with the
experimental data at the RHIC energy using the scale in the range
$\mu=0.9 \sim 1.5 p_T$. We find that the calculated $\pi^{0}$ spectra in
$p+p$ collisions with the scale $\mu=1.2p_T$ can fit RHIC data
very well and will use the same scale in $A+A$ collisions in our
following calculations.
We further assume that the factorized form for large transverse momentum single
hadron spectra in NLO pQCD parton model can be applied to heavy-ion collisions,
\begin{eqnarray}
\frac{1}{N_{\rm bin}^{AB}(b)}\frac{d\sigma_{AB}^h}{dyd^2p_T} &=&\sum_{abcd}\int
dx_adx_b f_{a/A}(x_a,\mu^2)f_{b/B}(x_b,\mu^2) \nonumber \\
&&\hspace{-0.5in}\times \frac{d\sigma}{d\hat{t}}(ab\rightarrow
cd)\frac{\langle \tilde{D}_{c}^{h}(z_{h},Q^2,E,b)\rangle}{\pi z_{c}}+\mathcal {O}(\alpha_s^3), \nonumber \\
\label{eq:AA}
\end{eqnarray}
where $N_{\rm bin}^{AB}(b)$ is the number of binary nucleon-nucleon collisions at a fixed impact $b$ of
$A+B$ nuclear collisions, $\langle \tilde{D}_{c}^{h}(z_{h},Q^2,E,b)\rangle$ is the medium modified
parton fragmentation function averaged over the initial production position and jet propagation direction,
and $f_{a/A}(x_{a},\mu^{2})$ is the effective parton distributions per nucleon inside a nucleus. For
large transverse momentum hadron production, the effect of nuclear modification of the parton
distributions is small. We will neglect such nuclear effect in our study here.
According to the definition of the jet transport parameter in
Eq.~(\ref{eq:qhat-phi}), it should be proportional to the local
particle density at $\vec r_{j}=\vec r +(\tau-\tau_{0})\vec n$ in
either QGP or hadronic phase of the evolving bulk medium
along the path of a propagating jet which is a straight line in the
eikonal approximation. Here $\tau_{0}$ is the formation time of the
medium and the direction of jet propagation is defined by its
azimuthal angle $\phi$ in the transverse plane. The quark-gluon
correlation in Eq.~(\ref{eq:corr2}) that enters the modified
fragmentation function in Eq.~(\ref{eq:mo-fragment}) will involve an
integration over the duration of the jet propagation and depend on
the initial production position $\vec r$ and propagation direction
$\vec n$. In heavy-ion collisions at a fixed impact parameter $\vec
b$, the initial jet production cross section is proportional to the
overlap function of two colliding nuclei $\int d^{2}r
t_{A}(r)t_{B}(|\vec b -\vec r|)$. One therefore has to average over
the initial jet production position and propagation direction to
obtain the effective medium modified parton fragmentation function,
\begin{eqnarray}
\langle \tilde D_{a}^{h}(z_h,Q^2,E,b)
\rangle &=& \frac{1}{\int
d^2{r}t_{A}(|\vec{r}|)t_{B}(|\vec{b}-\vec{r}|)} \nonumber \\
&&\hspace{-1.0in}\times \int\frac{d\phi}{2\pi}d^2{r}
t_{A}(|\vec{r}|)t_{B}(|\vec{b}-\vec{r}|)\tilde{D}_{a}^{h}(z_h,Q^2,E,r,\phi,b). \nonumber \\
\label{eq:frag}
\end{eqnarray}
With a given space-time profile of the jet transport parameter, one can therefore use the above
effective medium modified fragmentation functions to calculate the large transverse
momentum hadron spectra and the the suppression factor (or nuclear modification factor),
\begin{eqnarray}
R_{AB}(b)=\frac{d\sigma_{AB}^h/dyd^2p_T}{N_{bin}^{AB}(b)
d\sigma_{pp}^h/dyd^2p_T}, \label{eq:rab}
\end{eqnarray}
for characterization of the effect of jet quenching on hadron spectra in heavy-ion collisions \cite{modification-factor}.
Phenomenological study of the experimental data on the above hadron suppression factor will
in turn provide us constraints on the space-time profile of the jet transport parameter $\hat q_{R}(E,y)$.
The computation of the effective medium-modified jet fragmentation function in Eq.~(\ref{eq:frag}) involves
a 6-dimensional integration which is in addition to the Monte Carlo integration in the NLO pQCD calculation of
the single hadron spectra~\cite{Owens}. Numerically, we will compute and tabulate the medium modified fragmentation functions
as functions of the initial jet energy $E$, fractional momentum $z$ and the factorization scale $Q^{2}$ for fixed value of
impact parameter $b$. We will then
use numerical interpolation of the table for the medium modified fragmentation functions in the NLO pQCD calculation of the single
hadron spectra.
\section{Bulk matter evolution}\label{sec:c}
Recent phenomenological studies
\cite{CasalderreySolana:2007sw,Majumder:2007ae,Bass:2008rv,Bass:2008ch,Qin:2009gw}
have focused on the initial values of the jet transport parameters
as constrained by the experimental data. We study in this paper
the uncertainty of the extracted initial jet transport parameter due
to different models of the dynamical evolution of the bulk
medium and in particular the effect of transverse expansion, flow
effect, non-equilibrium and phase transition by considering three
different models for the bulk evolution. Assuming fast
thermalization of the partonic matter in the initial stage of
high-energy heavy-ion collisions, we will consider first (1+1)d Bjorken
\cite{Bjorken:1982qr} and (1+3)d ~\cite{hirano1, hirano2} ideal
hydrodynamic evolution of the bulk matter for the calculation of
medium modified parton fragmentation functions. Inclusion of shear
and bulk viscosity in the viscous hydrodynamics can affect
significantly the elliptic flow or azimuthal asymmetry of the final
hadron spectra at freeze-out. Their effect on the space-time profile
of the bulk medium such as entropy density, however, is much smaller
(on the order of a few percent)
\cite{Teaney:2003kp,Romatschke:2007mq,Song:2007ux,Dusling:2007gi}
which we will neglect here in the hydrodynamic evolution of the bulk
matter for the study of jet quenching. To consider non-equilibrium
evolution of the bulk matter and its effect on jet quenching, we
will also calculate medium modified hadron spectra with the
space-time profile of parton density as given by a parton cascade
model~\cite{Xu:2004mz,Xu:2007aa}.
\subsection{ (1+1)d Bjorken Expansion}\label{sec:c-a}
In a (1+1)d Bjorken model \cite{Bjorken:1982qr} for the bulk
evolution, the system is assumed to undergo a longitudinally
boost-invariant expansion with the (1+1)d ideal hydrodynamic equation,
\begin{equation}
\frac{d\epsilon}{d\tau}=-\frac{\epsilon+P}{\tau},
\end{equation}
where $\epsilon$ is the energy density, $P$ is the pressure and
$\tau=\sqrt{t^{2}-z^{2}}$ is invariant time. For a massless ideal
gas equation of state (EOS), the solution of the above ideal
hydrodynamic equation leads to a time evolution of the entropy
density,
\begin{equation}
s =s_{0}\frac{\tau_{0}}{\tau},
\end{equation}
with $s_{0}$ the entropy density at the initial time $\tau_{0}$.
Since the jet transport parameter $\hat q_{R}$ is directly
proportional to gluon distribution density and therefore
proportional to the particle number or entropy density, we assume
that it will have a similar time dependence due to longitudinal
expansion. In the (1+1)d Bjorken model, we further neglects the
transverse expansion and phase transition. Therefore the transverse
profile of jet transport parameter will be given by the initial
transverse profile of the parton density which we assume to be
proportional
to the transverse density of participant nucleons
according to the wounded nucleon model of nucleus-nucleus collision~\cite{Bialas:1976ed},
\begin{eqnarray}
\hat{q}= \hat{q}_0
\frac{n_{\rm{part}}(\vec{b},\vec{r})}{n_{\rm{part}}(\vec{0},0)}
\frac{\tau_0}{\tau}\,,
\label{q-hat-npart}
\end{eqnarray}
where $n_{\rm{part}}(\vec{b},\vec{r})$ is the transverse density of participant nucleons in nucleus-nucleus
collisions with impact parameter $b$,
\begin{eqnarray}
n_{\rm{part}}(\vec{b},\vec{r})&=& t_A(|\vec{r}|)
\left(1-e^{-\sigma_{NN}t_{A}(|\vec{b}-\vec{r}|)}\right ) \nonumber\\
&& +\ t_A(|\vec{b}-\vec{r}|)\left(1-e^{-\sigma_{NN}t_{A}(|\vec{r}|)}\right)\,,
\label{npart}
\end{eqnarray}
and $\sigma_{NN}$ is the nucleon-nucleon total inelastic cross section, which is set to be
$\sigma_{NN}=42$ mb at the RHIC energy. Here $t_A(\vec{r})$ is the nuclear thickness function,
\begin{equation}
t_A(\vec{r})=\int_{-\infty}^{\infty} dz\, \rho_A(\vec{r}, z)\,,
\end{equation}
and $\rho_A(\vec{r}, z)$ is the single nucleon density in the
nucleus normalized to $\int d^{2}r t_{A}(\vec r)=A$. The total
number of participant nucleons in nucleus-nucleus collisions at
impact-parameter $b$ is then $N_{\rm part}(b)=\int d^{2}r n_{\rm
part}(\vec b,\vec r)$. In the space-time profile of the jet
transport parameter in Eq.~(\ref{q-hat-npart}), $\hat q_{0}$ is
defined as the jet transport parameter at the center of the bulk
medium in the most central nucleus-nucleus collisions ($b=0$).
We will consider two different nuclear distributions for
$\rho_A(\vec{r},z)$ in the (1+1)d Bjorken model. One is the
Woods-Saxon distribution:
\begin{equation}
\label{wsf}
\rho_A(\vec{r},z)=\frac{n_0}{1+e^{\frac{\sqrt{|\vec{r}|^2+z^2}-R_A}{d}}}\,,
\end{equation}
where $d=0.662$ fm, $R_A=1.26 A^{1/3}$ fm, and $n_0$ is the
normalization factor given by $\int d^2r dz \,
\rho_A(\vec{r},z)=A$~\cite{Schwierz:2007ve}. We also consider
the simple hard-sphere distribution,
\begin{equation}
\label{hsf} \rho_A(\vec{r},z)=\frac{3A}{4\pi
R_A^3}\theta(R_{A}-\sqrt{|\vec{r}|^{2}+z^{2}}),
\end{equation}
with $R_{A}=1.12A^{1/3}$ for comparison.
In the present work we focus on jet quenching in the $10\%$
most central collisions. By comparing the averaged number of
participating nucleons,
\begin{eqnarray}
\label{av_npart} && \langle N_{\rm part} \rangle = \frac{\int db \,
b N_{\rm part}(b)} {\int db\, b},
\end{eqnarray}
which is given by experiments for the most
$10\%$ central collisions~\cite{Adams:2004bi}, we determine the
averaged value of impact-parameter $b=3.15$ fm for Woods-Saxon
nuclear distribution and $b=2.2$ fm for the hard-sphere
distribution.
\subsection{(1+3)d Ideal Hydrodynamical Expansion}\label{sec:c-b}
To take into account of both longitudinal and transverse expansion in the ideal hydrodynamic
description of the bulk matter evolution, we use the output of calculations
by Hirano {\it et al.}~\cite{hirano1,hirano2} for Au+Au collision at the
RHIC energy with Glauber collisions geometry and the average impact parameter $b=3.2$ fm, corresponding
to $10\%$ most central events. The initial condition for the (1+3)d ideal hydrodynamic equations is fixed such that
the final bulk hadron spectra from the RHIC experiments are reproduced \cite{hirano1,hirano2}.
We use the code provided by Hirano \cite{hirano-p} that interpolates the
numerical data from the hydrodynamic solution on a space-time grid for energy density, temperature,
flow velocity and fraction of QGP phase.
For our purpose, we have to infer parton and hadron number density
from the energy density or temperature provided by the hydrodynamic
solution, which uses a Bag model for the equation of state (EOS)
\cite{Nonaka:2000ek}. In this Bag model, the parton number and energy
density in the QGP phase are,
\begin{eqnarray}
\label{rhoqgp}
\rho^{QGP}&=&\left ( 16 \frac{\zeta{(3)}}{\pi^2}+
36 \frac{3\zeta{(3)}}{4\pi^2}\right ) T^3 \equiv a_1 T^ 3, \\
\epsilon^{QGP}&=&\left (16 \frac{\pi^2}{30}+36 \frac{7\pi^2}{240}\right )
T^4 +B\equiv b_1 T^4 +B, \label{eq:eos2}
\end{eqnarray}
for three quark flavors. The Bag constant $B=(247 \mbox{ Mev})^4$ gives rise to a first-order
phase transition from the QGP to a hadron resonance gas at the critical temperature $T_c=170$ MeV.
In our calculation of medium modified parton fragmentation functions
we consider jet-medium interaction in both partonic and hadronic
phase throughout the evolution of the bulk matter. Neglecting hadron
correlation in the medium, the jet transport parameter will be
proportional to hadron density and the soft gluon distribution
within each hadron. There have been several phenomenological studies
\cite{Wang:2002ri,Majumder:2004pt,Majumder:2009zu} of jet quenching in deeply inelastic scattering
off large cold nucleus as measured by the HERMES
\cite{Airapetian:2007vu} experiment. Within the same high-twist
approach, the jet transport parameter is found \cite{Deng:2009qb} to
be $\hat q_{N}\approx 0.02$ GeV$^{2}$/fm at the center of the cold
nucleonic matter in a large nucleus. In the (1+3)d ideal
hydrodynamic model, the hadronic phase of the medium evolution is
considered as a hadron resonance gas, in which the jet
transport parameter can be approximate as,
\begin{equation}
\hat q_{h}=\frac{\hat q_{N}}{\rho_{N}}\left[ \frac{2}{3}\sum_{M}\rho_{M}(T)+\sum_{B}\rho_{B}(T)\right]
\label{eq:qhath}
\end{equation}
where $\rho_{N}=n_{0}\approx 0.17$ fm$^{-3}$ is the nucleon density in the center of a large nucleus and the
factor $2/3$ accounts for the ratio of constituent quark numbers in mesons and baryons. The
hadron density at a given temperature $T$ and zero chemical potential is
\begin{eqnarray}
\sum_{h}\rho_{h}(T)=\frac{T^{3}}{2\pi^{2}}
\sum_{h}\left(\frac{m_{h}}{T}\right)^{2}
\sum_{n=1}^{\infty}\frac{\eta_{h}^{n+1}}{n}K_{2}\left(n\frac{m_{h}}{T}\right),
\end{eqnarray}
where $\eta_{h}=\pm$ for meson (M)/baryon (B). In the paper, we will include all hadron resonances with
mass below 1 GeV. Other hadron resonance gas models normally include hadrons with mass up to 2 GeV which
will only make some differences to our assumed $\hat q_{h}$ close to the phase transition temperature.
\begin{figure}
\centerline{\includegraphics[width=9cm]{fig3.eps}}
\caption{The temperature dependence of $\hat q/T^{3}$.}
\label{fig:qhat-t3}
\end{figure}
With the above approximation for hadron density and jet transport
parameter in the hadron phase of bulk medium evolution, the jet transport parameter throughout
the evolution of the medium can be expressed as,
\begin{equation}
\label{q-hat-qgph}
\hat{q} (\tau,r)= \hat{q}_0\frac{\rho^{QGP}(\tau,r)}{\rho^{QGP}(\tau_{0},0)} (1-f) + \hat q_{h}(\tau,r) f \,,
\end{equation}
where $f(\tau,r)$ is the fraction of the hadronic phase at any given
time and local position and $\hat q_{0}$ denotes the jet transport
parameter at the center of the bulk medium in the QGP phase at the
initial time $\tau_{0}$. To illustrate the behavior of the jet
transport parameter in dense medium, we plot $\hat q/T^{3}$ in
Fig.~\ref{fig:qhat-t3} as a function of the temperature $T$
according to the above assumptions, where we have used the initial
value of $\hat q_{0}=0.9^{+0.05}_{-0.04} $ GeV$^{2}$/fm, as
extracted by phenomenological study of experimental data later in
this paper, at an initial temperature $T_{0}=$ 369 MeV in the most
central $Au+Au$ collisions at RHIC. Apparently, we have neglected
any additional temperature dependence beyond that in
Eq.~(\ref{q-hat-qgph}). Perturbative calculations
\cite{CasalderreySolana:2007sw,Wang:2000uj} in finite temperature
QCD with resummed hard-thermal loops give rise to a logarithmic
temperature dependence of $\hat q/T^{3}$ for a fixed value of the
strong coupling constant $\alpha_{s}$. The same perturbative calculations also lead
to somewhat weak jet-energy dependence of the jet transport parameter which
we will neglect in this phenomenological study. Therefore, one can consider the exacted
jet transport parameter as an averaged value over the range of jet energies in the
experimental data.
In order to study the effect of phase transition in the (1+3)d hydrodynamic evolution of the
bulk matter in jet quenching, we set $f=0$ in Eq.~(\ref{q-hat-qgph}) and determine the parton density
$\rho= a_{1}[(\epsilon -B)/b_{1}]^{3/4}$ according to Eq.~(\ref{rhoqgp}) throughout the bulk evolution,
with $\epsilon$ given by the (1+3)d hydrodynamic evolution at given space and time. Calculations with
this prescription of space-time profile of the jet transport parameter will be labelled as no phase
transition in this paper. We can also set $\hat{q}_0^h=0$ to quantify the hadronic contribution to jet quenching.
The jet transport parameter $\hat{q}$ in Eq.~(\ref{q-hat-qgph}) is
given in the frame where the medium is at rest. To take into account
of the radial flow, one can simply make the following substitute
\cite{flow1,flow2}
\begin{equation}
\label{q-hat-flow}
\hat{q} \to \hat{q} \frac{p^\mu u_\mu}{p_0} \,,
\end{equation}
where $p^\mu$ is the four momentum of the jet and $u^\mu$ is the
four flow velocity in the collision frame, which is provided by the solution of the (1+3)d
hydrodynamical equations \cite{hirano1,hirano2}. Such flow dependence effectively
decreases (increases) the parton energy loss when the jet propagates
along the same (opposite) direction as the flow.
\subsection{(1+3)d Parton Cascade}\label{sec:c-c}
To study the effect of non-equilibrium evolution of the bulk matter on jet quenching in this paper, we
also consider a parton cascade model. The Boltzmann Approach of MultiParton Scatterings (BAMPS)
model \cite{Xu:2004mz,Xu:2007aa} solves the Boltzmann equation for on-shell gluons with pQCD based
interactions, which include elastic scattering, bremsstrahlung and its back reaction. The thermal equilibration
process of gluons has been thoroughly investigated within this parton transport model \cite{Xu:2007aa,El:2007vg}.
BAMPS can in principle describe collective flow behavior \cite{Xu:2007jv,Xu:2008av,Bouras:2009nn} and
the nuclear modification factor $R_{AA}$ of jet quenching \cite{Fochler:2008ts} within a common framework.
In this work we use the output from BAMPS \cite{Xu:2008av}
model for central $Au+Au$ collisions with $b=3.4$ fm at the RHIC
energy. Initial gluons or minijets are produced by binary
nucleon-nucleon collisions with a Glauber geometry. The Lorentz
contracted nuclei have a longitudinal extent of $0.2$ fm. Thus, most
minijets are produced at $t=0.1$ fm/c when two nuclei fully overlap.
The strong QCD coupling constant is set to be a constant,
$\alpha_s=0.3$. Gluons freeze out when the local energy density
goes below a critical value, which is chosen as $\epsilon_c=0.6
\mbox{ GeV/}\mbox{fm}^3$. This corresponds to a critical temperature
of $T_c=175$ MeV in a fully equilibrated gluon gas. In the present
implementation of BAMPS there are no hadronization and hadron
cascade. Gluons propagate freely (free-streaming) after freeze-out.
During this stage we regard one gluon as one hadron according to a
parton-hadron duality picture. With these setups the final
transverse energy and the elliptic flow $v_2$ from BAMPS
calculations agree with the experimental data at RHIC
\cite{Xu:2007jv}.
We use Eqs.~(\ref{q-hat-qgph}) and (\ref{q-hat-flow}) to evaluate the jet transport parameter
in the evolving bulk medium as described by the BAMPS model. The local gluon density
and flow velocity are obtained by integrating the local gluon phase space distribution given
by BAMPS. The hadron fraction $f$ is either $0$ before gluon freeze-out or $1$ after gluon freeze-out
as a simplification for the hadronization process. By setting $f=0$ all the time, one effectively neglects the
effect of phase transition and the difference in the jet transport parameter in QGP and hadronic phase.
The hadron density will be estimated from a free-streaming hadron gas according to an assumption of
hadron-parton duality.
\begin{figure*}[t]
\begin{center}
\includegraphics[width=155mm]{raa.eps}
\caption{(color online) (a) Nuclear modification factor at
midrapidity for the most $10\%$ central Au+Au collisions at the maximum
RHIC energy $\sqrt{s}=200$ GeV/n. The symbols are the PHENIX data taken from
Ref.~\cite{raa}. The curves are NLO pQCD calculations with three different values of
$\hat{q}_0 \tau_0$ in the (1+1)d Bjorken
model with the Hard-Sphere (upper) and Woods-Saxon (lower) distribution function.
(b) The corresponding $\chi^2$ of the fit as a function of $\hat{q}_0\tau_0$.} \label{fig:raa}
\end{center}
\end{figure*}
\section{Numerical results and comparisons}\label{sec:d}
With the three different models for the bulk matter evolution and
the prescriptions for the evaluation of the jet transport parameter
in the evolving medium as described in the last section, we can
conduct a systematic analysis of the experimental data on
suppression of high $p_{T}$ hadron spectra within the higher-twist
approach to the medium modified jet fragmentation functions. Our aim
in this work is to study the theoretical uncertainties associated
with the dynamical evolution of the bulk medium and effects of jet hadron
interaction in the hadronic medium on jet quenching. Since we assume that the jet
transport parameter in the hadronic phase can be estimated from the
value as determined in the cold nuclear matter in the DIS off
nuclear targets [Eq.~(\ref{eq:qhath})], our specific problem in this
case is to determine the value of the jet transport parameter $\hat
q_{0}$ at the center of the overlapped region in the central $Au+Au$
collisions in the QGP phase at the initial time $\tau_{0}$ [see
Eqs.~(\ref{q-hat-npart}) and (\ref{q-hat-qgph})].
Shown in Fig.~\ref{fig:raa}(a) are comparisons between the PHENIX
experimental data on $R_{AA}$ for the neutral pion spectra in the
$10\%$ most central $Au+Au$ collisions \cite{raa} at the highest
RHIC energy $\sqrt{s}=200$ GeV/$n$ and our NLO pQCD calculations
with the medium modified fragmentation functions as given in the
higher-twist approach, for three different values of $\hat{q}_0$ at
the initial time $\tau_0=0.6$ fm/$c$, using (1+1)d Bjorken model of
ideal hydrodynamical evolution without a phase transition. The upper
panel of Fig.~\ref{fig:raa} (a) shows the results with the
hard-sphere nuclear geometry while the lower panel is for
Woods-Saxon nuclear distribution. The difference between the
calculated suppression factors with hard-sphere and Woods-Saxon
nuclear distributions is about 10 percent, due to the small difference in
the thickness and overlap functions with these two nuclear distributions.
In the remainder of this paper, we will discuss the determination of initial jet transport parameter and its
uncertainties associated with the modeling of the bulk medium evolution.
We will carry out systematical $\chi^{2}$ fits to the experimental data on the suppression factor
$R_{AA}$ of neutral pion spectra from the PHENIX experiment \cite{raa}
with our NLO pQCD calculations that include jet quenching in
an evolving bulk medium. We will follow Ref.~\cite{Adare:2008cg} to carry out a modified $\chi^{2}$ analysis
and obtain the best fit parameters $\hat{q}_0 \tau_0$ for each model of bulk
medium evolution by minimizing
\begin{equation}
\chi^2=\sum_{i=1}^{N}\frac{[y_i+\epsilon_b\sigma_{b_i}+\epsilon_cy_i\sigma_c-y_i^{th}]^2}
{\tilde{\sigma}_i^2}+\epsilon_b^2+\epsilon_c^2\,,
\end{equation}
\begin{equation}
\tilde{\sigma}_i=\sigma_i(\frac{y_i+\epsilon_b\sigma_{b_i}+\epsilon_cy_i\sigma_c}{y_i})\,,
\end{equation}
of the fits with respect to two parameters $\epsilon_b$ and
$\epsilon_c$, where $y_i$ is the central value of experimental data
with three types of uncertainties, $\sigma_i$ (statistical plus
uncorrelated systematics), $\sigma_{b_i}$ (correlated systematics)
and $\sigma_{c}$ (normalization uncertainties) for $p_T$ larger than
$6.5$ GeV/$c$, and $y_i^{th}$ is the calculated results at each
$p_T$ point. Shown in Fig.~\ref{fig:raa} (b) as two examples are the
$\chi^2$ distributions as functions of $\hat{q}_0 \tau_0$
($\tau_{0}$=0.6 fm/$c$) from the fits with the NLO pQCD calculation
with (1+1)d Bjorken model for the bulk medium evolution. The best
fit value is $\hat{q}_0\tau_0=0.36 \mbox{GeV}^2$ and 0.39
$\mbox{GeV}^2$
for Woods-Saxon and hard-sphere nuclear distribution function, respectively.
In the following calculations, we will always use the Woods-Saxon nuclear distribution function.
Listed in Table~\ref{table1} are the best fit values of $\hat{q}_0
\tau_0$ and the corresponding $\chi^{2}$ for jet quenching in
various models for the bulk medium evolution. We will discuss and
compare them in detail in the remainder of this section.
\begin{table}[htbp]
\begin{center}
\begin{tabular}{l|ccccc}
\hline \hline Model & flow
&PT &$\hat{q}_0\tau_0 (\mbox{GeV}^2)$ & $\chi^{2}$ \\
\hline (1+1)d Bjorken [W-S] & &
& 0.36 & 78 \\
\hline (1+3)d Hydro & & & 0.42 & 116 \\
& $\surd$ & & 0.48 & 80 \\
& $\surd$ &$\surd$ & 0.54 & 106 \\
($\hat{q}_0^h=0$) & $\surd$ &$\surd$ & 0.78 & 71 \\
\hline BAMPS & & & 0.72 & 95 \\
& $\surd$ & & 0.84 & 84 \\
& $\surd$ & $\surd$ & 0.96 & 96 \\
($\hat{q}_0^h=0$) & $\surd$ & $\surd$ & 1.17 & 86 \\
\hline \hline
\end{tabular}
\caption{Initial jet transport parameter $\hat{q}_0\tau_0$ and the
corresponding $\chi^{2}$. The check sign ($\surd$) denotes the
inclusion of flow or phase transition (PT) in the calculations,
respectively. The initial $\tau_0$ is 0.6 fm/c for hydrodynamic and
0.3 fm/$c$ for BAMPS model.} \label{table1}
\end{center}
\end{table}
\subsection{(1+1)d Bjorken versus (1+3)d Ideal Hydrodynamical Model}\label{sec:d-b}
We first compare the results from (1+1)d Bjorken and (1+3)d ideal
hydrodynamic model with different options and study the effects of
the transverse expansion, radial flow and
contributions from jet quenching in the hadronic phase. Comparing
the (1+1)d Bjorken and (1+3)d ideal hydrodynamical evolution without
radial flow and phase transition it is clear that the transverse
expansion cools the system considerably faster in the later stage of
the evolution. This faster cooling due to the transverse expansion
leads to a faster decrease of the jet transport parameter in the
(1+3)d hydrodynamic model than in the (1+1)d Bjorken model as shown
in Fig.~\ref{fig:qhat} by the dot-dashed [(1+3)d hydrodynamics] and
solid lines [(1+1)d Bjorken hydrodynamics]. To compensate for this
faster cooling, one has to increase the value of the initial jet
transport parameter $\hat q_{0}$ by 17\% in the case of (1+3)d
hydrodynamic evolution relative to the case of (1+1)d Bjorken
expansion in order to fit the measured suppression factor
$R_{AA}(p_{T})$ for neutral pions in the central $Au+Au$ collisions
at RHIC.
\begin{figure*}[t]
\begin{center}
\includegraphics[width=140mm]{qhat.eps}
\caption{(color online). Time evolution of the scaled jet transport
parameter at position of radius of $2$ fm (left panel) and $4$ fm
(right panel) for various models of the bulk matter evolution. The
flow effect in Eq.(\ref{q-hat-flow}) is not included.}
\label{fig:qhat}
\end{center}
\end{figure*}
\begin{figure*}[t]
\begin{center}
\includegraphics[width=140mm]{flow.eps}
\caption{(color online). The radial flow velocity in the (1+3)d
hydrodynamical model and the parton cascade at position of radius of
$2$ fm (left panel) and $4$ fm (right panel). The solid curves show
the fraction of the hadron phase in the hydrodynamical model.}
\label{fig:flow}
\end{center}
\end{figure*}
Inclusion of the radial flow in the calculation of the jet transport parameter as in Eq.~(\ref{q-hat-flow}) will also
change the effective jet quenching throughout the evolution of the bulk medium. Shown in Fig.~\ref{fig:flow}
are the time evolution of the radial flow velocity in (1+3)d hydrodynamics (dashed line). The radial
flow velocity in the center of the dense medium (left panel), where jet quenching is the strongest, remains
small, at about 0.2, throughout hydrodynamic expansion. In the same figure, we also plot the fraction of the
hadron phase (solid lines) as a function of time in the (1+3)d hydrodynamics. The period
between $f=0$ and $f=1$ indicate the duration of the phase transition or the mixed phase.
We notice that at the beginning of the mixed phase the radial flow
velocity starts to saturate and even decrease a little. This is mainly due to the mass effect.
The energy, approximately $m_T v$, where $m_T$ is the transverse mass,
should be the same during the phase transition. Therefore, the radial flow velocity will decrease
as particles become massive during the phase transition. Such decrease will be partially compensated
by the continued collective expansion which tend to increase the flow velocity. As the phase transition
is completed, the radial flow velocity should increase again by further transverse expansion in the
hadronic phase and the resonances decays.
The radial flow will reduce the effective jet transport parameter
$\hat q$ if jets propagate along with it while increase the
effective value of $\hat q$ if jets propagate against it. After
averaging over the direction and initial position of jet production,
the effective jet transport parameter is merely reduced by less than
14\% \cite{flow1}. As shown in Table \ref{table1}, this leads to an
increase of $\hat q_{0}$ about 14\% when one includes the effect of
radial flow in the (1+3)d hydrodynamic evolution in fitting
experimental data on $R_{AA}(p_{T})$.
In both (1+1)d Bjorken and two options of the (1+3)d hydrodynamic
expansion as discussed above, we have neglected the difference of
the jet transport parameters in QGP and hadronic phases. These
options are labelled as no phase-transition (PT) by setting $f=0$ in
Eq.~(\ref{q-hat-qgph}) and using
$\rho^{QGP}=a_{1}[(\epsilon-B)/b_{1}]^{3/4}$ in Eq.~(\ref{rhoqgp})
throughout the evolution even in the mixed and hadronic phase.
In the realistic scenario, jet transport
parameter $\hat q_{h}$ in the hadronic phase should be different
from that in the QGP phase and we assume that $\hat q_{h}$ is
proportional to $\hat q_{N}$ in the cold nuclear medium and the
hadron number density as in Eq.~(\ref{eq:qhath}). With this
assumption, there is a discontinuity in $\hat q$ as a function of
time (or temperature as in Fig.~\ref{fig:qhat-t3}) at the end of the mixed phase,
as show by the dotted lines in Fig.~\ref{fig:qhat}, and the jet transport parameter
in the hadronic phase in the later stage of the bulk medium
evolution will be smaller than the scenario of no PT in the (1+3)d
hydrodynamics. Therefore, one needs a larger value of the initial
jet transport parameter $\hat q_{0}$, about 13\%, to account for the
measured suppression factor $R_{AA}(p_{T})$ as compared to an
evolving bulk medium with the same jet transport parameter
throughout difference phases.
To determine the contribution of jet quenching in the hadronic
phase, we set $\hat{q}^h_0=0$ in Eq.~(\ref{q-hat-qgph}), assuming
parton energy loss only happens in the QGP phase. In this case, the
extracted $\hat{q}_0$ will be about 44\% larger (see
Table~\ref{table1}). This implies that the hadronic phase
contributes to about 44\% of the total suppression of the high
$p_{T}$ hadron spectra. Such a large contribution is due to the
lifetime of the mixed and hadronic phase until the kinetic freeze-out which is
longer than the lifetime of the QGP phase, though the hadron density
is much smaller than the initial parton density in the QGP phase. A
better understanding of the hadronic contribution to the jet
quenching is, therefore, important for an accurate extraction of the
jet transport parameter in the initial phase of strongly interacting
QGP.
\subsection{(1+3)d Parton Cascade versus (1+3)d Ideal Hydrodynamical Model}
\label{sec:d-c}
In a parton cascade model such as BAMPS, the system takes times to
reach thermal or partial thermal equilibrium after the initial
parton production time, $\tau_{0}=0.3$ fm/$c$. During this period of
parton equilibration, there is entropy production accompanied by
rapid expansion. The net effect is a much faster decreasing of the
effective temperature \cite{Xu:2004mz,Xu:2007aa,Biro:1993qt} and the
parton density during the early stage of the bulk medium evolution.
The corresponding jet transport parameter in the equilibrating
gluonic matter in BAMPS calculation therefore decreases much faster
than that in both (1+1)d and (1+3)d ideal hydrodynamic models in the
early stage of evolution as shown by the dashed lines in
Fig.~\ref{fig:qhat}. It is smaller by more than a factor of 2 before
the hadronization. This leads to about 77\% increase in the initial
value of $\hat q_{0}$ that is needed to account for the measured
suppression of hadron spectra (see Table~\ref{table1}). Note that in
our calculations jet quenching starts at the thermalization
$\tau_0=0.6$ fm/$c$ in the ideal hydrodynamic model. However, it
starts at $\tau_0=0.3$ fm/$c$ in the BAMPS model immediately after
the initial gluons are produced and before thermal or partial
thermal equilibrium is reached. Such jet medium interaction and
parton energy loss during the thermalization stage in the BAMPS
model will contribute to some additional, though a small fraction,
suppression of the final hadron spectra.
In BAMPS, the hadronization of the gluonic matter happens when the
local gluon energy density is below $\epsilon_{c}=0.6 \mbox{GeV/
fm}^3$ and parton-hadron duality is used for the phase-space density
of hadrons, which are assumed to freeze out immediately. Such a
sudden hadronization and freeze-out scenario implies the change of the hadron
fraction $f$ from 0 to 1 at this critical density in the evaluation
of the jet transport parameter in Eq.~(\ref{q-hat-qgph}). The lack
of the mixed phase in BAMPS, which lasts for about $5-6$ fm/$c$ in
the (1+3)d ideal hydrodynamic evolution as indicated by the duration
for the hadron fraction to reach $f=1$ in Fig.~\ref{fig:flow}, leads
to smaller value of the jet transport parameter after the
hadronization. Comparing the extracted values of $\hat q_{0}$ from
the phenomenological fit with and without ($\hat q_{h}=0$) hadronic
phase as shown in Table~\ref{table1}, the hadronic interaction in
BAMPS model contributes to about 22\% of the jet quenching effect,
which is about a factor 2 smaller than that in the (1+3)d ideal
hydrodynamical model.
The large viscous effect in the parton cascade model
\cite{Xu:2007jv} generally leads to larger radial flow velocity
than that generated by an ideal hydrodynamic expansion as shown in
Fig.~\ref{fig:flow}. The free-streaming of hadrons immediately after
the hadronization also leads to the further increase of the flow
velocity. Such larger radial flow velocity in the BAMPS model
reduces the effective jet quenching parameter and therefore leads an
increase of about 16\% in the extracted initial $\hat q_{0}$ .
The decrease of the radial flow velocity at late times seen in Fig.~\ref{fig:flow}
is a numerical artifact, which comes from the chosen geometry of
spatial elements, where particles are collected to calculate the
radial flow velocity. Longitudinal length of each spatial element
corresponds to a space time rapidity window $|\eta| < 0.25$, while
the transverse length is fixed to be 0.3 fm. At late times the
longitudinal length is larger than the transverse one, and
particles with smaller transverse velocity (larger longitudinal
velocity) make up the main part of remaining particles in the spatial
elements. This is the reason for the observed decrease of the radial
flow velocity. If we choose the longitudinal length being equal as
the transverse one at the late times, we would see a continuous
increase of the flow velocity due to the free-streaming, before all
particles go out of the spatial element. This is, however, a
negligible effect on the extraction of $\hat q_{0}$, because the
hadron density is tiny at late times.
\section{Conclusions}\label{sec:e}
Using medium-modified fragmentation functions from the high-twist
approach to multiple parton scattering we have calculated single
inclusive hadron spectra in NLO perturbative QCD parton model. The
medium modification to the fragmentation function is very similar in
form to the DGLAP correction due to vacuum gluon bremsstrahlung,
except that the medium modified splitting function contains
information about the properties of the medium through a
path-integrated jet transport parameter $\hat q$. Therefore, the
medium modified fragmentation function will explicitly depend on the
space-time evolution of the bulk medium. We have focused our
attention on the dependence of the medium modified fragmentation
function on the space-time evolution of the bulk medium and their
influence on the suppression of the single hadron spectra at large
$p_{T}$.
We specifically considered the effects of transverse expansion,
radial flow, jet quenching in hadronic phase and non-equilibrium
effect with three different models for the bulk matter evolution in
central $Au+Au$ collisions at the RHIC energy: (1+1)d Bjorken
hydrodynamic model, (1+3)d ideal hydrodynamical model by Hirano
\cite{hirano1, hirano2} and (1+3)d BAMPS \cite{Xu:2004mz,Xu:2007aa}
parton cascade model. Since the jet transport parameter is
proportional to the particle gluon distribution density of the
medium, we have assumed that it will be proportional to the local
particle density in both QGP and hadronic phase during the bulk
matter evolution. The coefficient of the particle density
dependence, which is related to soft gluon distribution per particle,
will be different in the QGP and hadronic
phase. Assuming a similar density dependence of the jet transport
parameter in the hadronic phase as that in a large cold
nuclei, but rescaled by the local hadron density and the number of
constituent quarks, the only free parameter in the space-time profile is
the initial value of the jet transport parameter $\hat q_{0}$ at
initial time $\tau_{0}$ at the center of the QGP medium in central
heavy-ion collisions. We have extracted values of $\hat{q}_0$ at
$\tau_0$ in each model of bulk evolution with different options by
fitting the experimental data on the suppression factor
$R_{AA}(p_{T})$ for high $p_{T}$ neutral pions and studied the
uncertainties in the extracted value of $\hat q_{0}$.
By comparing the values of $\hat q_{0}$ extracted with different
options in models for the bulk evolution, we also found that the
transverse expansion leads to fast cooling of the hot medium and
therefore reduces jet quenching in the later stage of the evolution
by about 17\%. Radial flow also reduces the effective jet transport
parameter by 15\%. The phase transition from QGP to hadronic medium
further reduces the effective jet transport parameter in the
hadronic phase and leads to about 13\% reduction of jet quenching
effect. Within our model for the form of jet transport parameter due
to jet-hadron interaction, the overall contribution to jet quenching
from hadronic phase is about 22-44\%, depending on the model for
evolution of the hadronic phase.
By comparing (1+3)d ideal hydrodynamic and the BAMPS parton cascade
model, we found that the non-equilibrium evolution in the early
stage of parton cascade leads to faster decrease of the jet
transport parameter with time and therefore affects the overall jet
quenching the most by 77\%, part of which is caused by the lack of
mixed phase in the parton cascade model. Such uncertainties can be
further reduced by combining parton cascade and hadron cascade with
a model of of hadronization that can reproduce the EOS during the
phase transition as given by lattice QCD calculation.
\section*{Acknowledgements}
We would like to thank T. Hirano for providing numerical results
of (1+3)d ideal hydrodynamic evolution of the bulk matter in
heavy-ion collisions at RHIC and discussions. This work is supported
by the NSFC of China under Projects No. 10825523, No. 10635020, by
MOE of China under Projects No. IRT0624, by MOST of China under
Project No. 2008CB317106 and by MOE and SAFEA of China under Project
No. PITDU-B08033, and by the Director, Office of Energy Research,
Office of High Energy and Nuclear Physics, Divisions of Nuclear
Physics, of the U.S. Department of Energy under Contract No.
DE-AC02-05CH11231. The numerical calculations were performed at the
Center for Scientific Computing of Goethe University. This work was
financially supported by the Helmholtz International Center for FAIR
within the framework of the LOEWE program (Landes-Offensive zur
Entwicklung Wissenschaftlich-\"okonomischer Exzellenz) launched by
the State of Hesse, Germany. X.-N. Wang thanks the hospitality of
the Institut f\"ur Theoretische Physik, Johann Wolfgang
Goethe-Universit\"at and support by the ExtreMe Matter Institute
EMMI in the framework of the Helmholtz Alliance Program of the
Helmholtz Association (HA216/EMMI) during the completion of this
work.
|
\section{Introduction}
Propositional dynamic logic (PDL) is an important logic for reasoning about programs.
Its formulae consist of traditional Boolean formulae
plus ``action modalities'' built from a finite set of atomic programs
using sequential composition~$(\psp{}{})$, non-deterministic choice~$(\pup{}{})$,
repetition~$(\prp{})$, and test~$(\pip{})$.
The logic CPDL is obtained by adding converse ($\pcv{}$),
which allows us to reason about previous actions.
The satisfiability problem for CPDL is \textsc{exptime}{}-complete~\cite{vardi85:tamin_conver}.
De Giacomo and Massacci~\cite{de-giacomo-massacci-tableaux-converse-pdl}
give an \textsc{nexptime}{} tableau algorithm for deciding CPDL-satisfiability,
and discuss ways to obtain optimality,
but do not give an actual \textsc{exptime}{} algorithm.
The tableau method of Nguyen and Sza{\l}as~\cite{nguyen-szalas-cpdl} is optimal.
Neither method has been implemented,
and since both require an explicit analytic cut rule,
it is not at all obvious that they can be implemented efficiently.
Optimal game-theoretic methods for fix-point logics~\cite{lange-stirling-focus-games}
can be adapted to handle CPDL~\cite{lange03:satis_compl_conver_pdl_replay}
but involve significant non-determinism.
Optimal automata-based methods~\cite{vardi-wolper-automata}
for fix-point logics are still in their infancy
because good optimisations are not known.
We know of no resolution methods for CPDL.
We give an optimal tableau method for deciding CPDL-satisfiability
which does not rely on a cut rule.
Our main contribution is a sound method to combine our method
for tracking and detecting unfulfilled eventualities
as early as possible in PDL~\cite{gore-widmann-pdl-cade09}
with our method for handling converse for $ALCI$~\cite{gore-widmann-alci-tableaux09}.
The extension is non-trivial as the two methods cannot be combined naively.
We present a mixture of pseudo code and tableau rules
rather than a set of traditional tableau rules to enable easy implementation by others.
Our unoptimised OCaml implementation appears to be the first automated theorem prover for CPDL
(\url{http://rsise.anu.edu.au/~rpg/CPDLTabProver/}).
The proofs can be found in the appendix.
\section{Syntactic Preliminaries}
\label{secd:syntax-semantics}
\begin{definition}
Let~$\mathrm{AFml}$ and~$\mathrm{APrg}$ be two disjoint and countably infinite sets
of propositional variables and \emph{atomic programs}, respectively.
The set~$\mathrm{LPrg}$ of \emph{literal programs} is defined
as $\mathrm{LPrg} := \mathrm{APrg} \cup \{ \pcv{a} \mid a \in \mathrm{APrg} \}$.
The set~$\mathrm{Fml}$ of all formulae and the set~$\mathrm{Prg}$ of all \emph{programs}
are defined mutually inductively as follows where~$p \in \mathrm{AFml}$ and~$l \in \mathrm{LPrg}$:
\begin{tabular}[c]{ccccccccccccccc}
$\mathrm{Fml}$ \ \ \ & $\varphi$ & ::=
& $p$ & $\mid$
& $\pnot{\varphi}$ & $\mid$
& $\pand{\varphi}{\varphi}$ & $\mid$
& $\por{\varphi}{\varphi}$ & $\mid$
& $\pea{\gamma}{\varphi}$ & $\mid$
& $\paa{\gamma}{\varphi}$
\\
$\mathrm{Prg}$ \ \ \ & $\gamma$ & ::=
& $l$ & $\mid$
& $\psp{\gamma}{\gamma}$ & $\mid$
& $\pup{\gamma}{\gamma}$ & $\mid$
& $\prp{\gamma}$ & $\mid$
& $\pip{\varphi} \enspace.$
\end{tabular}
\noindent{} A $\pea{\mathrm{lp}}{}$-formula is a formula~$\pea{\gamma}{\varphi}$
where~$\gamma \in \mathrm{LPrg}$ is a \underline{\emph{l}}iteral \underline{\emph{p}}rogram.
\end{definition}
Implication~($\pimp{}{}$) and equivalence~($\peq{}{}$)
are not part of the core language but can be defined as usual.
In the rest of the paper, let~$p \in \mathrm{AFml}$ and~$l \in \mathrm{LPrg}$.
We omit the semantics as it is a straightforward extension of PDL~\cite{gore-widmann-pdl-cade09}
and write~$\psr{M, w}{\varphi}$ if~$\varphi \in \mathrm{Fml}$ holds in the world~$w \in W$ of the model~$M$.
\begin{definition}
For a literal program~$l \in \mathrm{LPrg}$, we define~$\pnnp{l}$ as~$a$
if~$l$ is of the form~$\pcv{a}$, and as~$\pcv{l}$ otherwise.
A formula~$\varphi \in \mathrm{Fml}$ is in \emph{negation normal form}
if the symbol~$\pnot{}$ appears only directly before propositional variables.
For every~$\varphi \in \mathrm{Fml}$, we can obtain a formula~$\mathop{\mathrm{nnf}}(\varphi)$ in negation normal form
by pushing negations inward such that~$\peq{\varphi}{\mathop{\mathrm{nnf}}{\varphi}}$ is valid.
We define~$\pneg{\varphi} := \mathop{\mathrm{nnf}}(\pnot{\varphi})$.
\end{definition}
We categorise formulae as $\alpha$- or $\beta$-formulae as shown in Table~\ref{tabd:alphabeta}
so that the formulae of the form~$\peq{\alpha}{\pand{\alpha_1}{\alpha_2}}$ and~$\peq{\beta}{\por{\beta_1}{\beta_2}}$ are valid.
\begin{table}[t]
\caption{Smullyan's $\alpha$- and $\beta$-notation to classify formulae}
\label{tabd:alphabeta}
\centering
\begin{tabular}{|c|c|c|c|c|c|c|}
\hline
$\alpha$
& $\pand{\varphi}{\psi}$
& $\paa{\pup{\gamma}{\delta}}{\varphi}$
& $\paa{\prp{\gamma}}{\varphi}$
& $\pea{\pip{\psi}}{\varphi}$
& $\pea{\psp{\gamma}{\delta}}{\varphi}$
& $\paa{\psp{\gamma}{\delta}}{\varphi}$
\\ \hline
$\alpha_1$
& $\varphi$
& $\paa{\gamma}{\varphi}$
& $\varphi$
& $\varphi$
& $\pea{\gamma}{\pea{\delta}{\varphi}}$
& $\paa{\gamma}{\paa{\delta}{\varphi}}$
\\ \hline
$\alpha_2$
& $\psi$
& $\paa{\delta}{\varphi}$
& $\paa{\gamma}{\paa{\prp{\gamma}}{\varphi}}$
& $\psi$
&
&
\\ \hline
\end{tabular}
\begin{tabular}{|c|c|c|c|c|}
\hline
$\beta$
& $\por{\varphi}{\psi}$
& $\pea{\pup{\gamma}{\delta}}{\varphi}$
& $\pea{\prp{\gamma}}{\varphi}$
& $\paa{\pip{\psi}}{\varphi}$
\\ \hline
$\beta_1$
& $\varphi$
& $\pea{\gamma}{\varphi}$
& $\varphi$
& $\varphi$
\\ \hline
$\beta_2$
& $\psi$
& $\pea{\delta}{\varphi}$
& $\pea{\gamma}{\pea{\prp{\gamma}}{\varphi}}$
& $\pneg{\psi}$
\\ \hline
\end{tabular}
\end{table}
An \emph{eventuality} is a formula of the form~$\pea{\gamma_1}{\dotsc \pea{\gamma_k}{\pea{\prp{\gamma}}{\varphi}}}$,
and~$\mathrm{Ev}$ is the set of all eventualities.
Using Table~\ref{tabd:alphabeta}, the binary relation~``$\mathrel{\rightsquigarrow}$''
relates a $\pea{}{}$-formulae~$\alpha$ (respectively~$\beta$),
to its reduction~$\alpha_1$ (respectively~$\beta_1$ and~$\beta_2$).
See~\cite[Def.~7]{gore-widmann-pdl-cade09} for their formal definitions.
\section{An Overview of our Algorithm}
\label{sec:algorithm}
Our algorithm builds an and-or graph~$G$
by repeatedly applying four rules (see Table~\ref{tabd:rules})
to try to build a model for a given~$\phi$ in negation normal form.
Each node~$x$ carries a formula set~$\Gamma_x$, a status~$\mathrm{sts}_x$,
and other fields to be described shortly.
Rule~1 applies the usual expansion rules to a node to create its children.
These expansion rules capture the semantics of CPDL.
We use Smullyan's $\alpha/\beta$-rule notation for classifying rules and nodes.
As usual, a node~$x$ is a (``saturated'') \emph{state}
if no $\alpha/\beta$-rule can be applied to it.
If~$x$ is a state then for each~$\pea{l}{\xi}$ in~$\Gamma_x$,
we create a node~$y$ with~$\Gamma_y = \{ \xi \} \cup \Delta$, where~$\Delta = \{\psi \mid \paa{l}{\psi} \in \Gamma_x\}$,
and add an edge from~$x$ to~$y$ labelled with~$\pea{l}{\xi}$
to record that~$y$ is an $l$-successor of~$x$.
If~$\Gamma_x$ contains an obvious contradiction during
expansion, its status becomes ``closed'', which is irrevocable.
Else, at some later stage, Rule~2 determines its status as either ``closed'' or ``open''.
``Open'' nodes contain additional information
which depends on the status of other nodes.
Hence, if a node changes its status, it might affect the status of another (``open'') node.
If the stored status of a node does not match its current status,
the node is no longer \emph{up-to-date}.
Rule~3, which may be applied multiple times to the same node,
ensures that ``open'' nodes are kept up-to-date
by recomputing their status if necessary.
Finally, Rule~4 detects eventualities which are impossible to fulfil
and closes nodes which contain them.
We first describe the various important components of our algorithm separately.
\paragraph{Global State Caching.}
For optimality, the graph~$G$ never contains two state nodes
which carry the same set of formulae~\cite{gore-widmann-alci-tableaux09}.
However, there may be multiple non-states which carry the same set of formulae.
That is, a non-state node~$x$ carrying~$\Gamma$
which appears while saturating a child~$y$ of a state~$z$ is unique to~$y$.
If a node carrying~$\Gamma$ is required in some other saturation phase,
a new node carrying~$\Gamma$ is created.
Hence the nodes of two saturation phases are distinct.
\paragraph{Converse.}
Suppose state~$y$ is a descendant of an $l$-successor of a state~$x$,
with no intervening states.
Call~$x$ the parent state of~$y$ since all intervening nodes are not states.
We require that $\{ \psi \mid \paa{\pcv{l}}{\psi} \in \Gamma_y \} \subseteq \Gamma_x$,
since~$y$ is then compatible with being a $l$-successor of~$x$
in the putative model under construction.
If some~$\paa{\pcv{l}}{\psi} \in \Gamma_y$ has~$\psi \notin \Gamma_x$ then~$x$ is ``too small'',
and must be ``restarted'' as an alternative node~$x^{+}$ containing all such~$\psi$.
If any such~$\psi$ is a complex formula to which an $\alpha/\beta$-rule is applicable
then~$x^{+}$ is not a state and may have to be ``saturated'' further.
The job of creating these alternatives is done by
\emph{special nodes}~\cite{gore-widmann-alci-tableaux09}.
Each special node monitors a state and creates the alternatives when needed.
\paragraph{Detecting Fulfilled and Unfulfilled Eventualities.}
Suppose the current node~$x$ contains an eventuality~$e_x$.
There are three possibilities.
The first is that~$e_x$ can be fulfilled in the part of the graph which is ``older'' than~$x$.
Else, it may be possible to reach a node~$z$ in the parts of the graph ``newer'' than~$x$
such that~$z$ contains a reduction~$e_z$ of~$e_x$.
Since this ``newer'' part of the graph is not fully explored yet,
future expansions may enable us to fulfil~$e_x$ via~$z$,
so the pair~$(z, e_z)$ is a ``potential rescuer'' of~$e_x$.
The only remaining case is that~$e_x$ cannot be fulfilled in the ``older'' part of the graph,
and has no potential rescuers.
Thus future expansions of the graph cannot possibly help to fulfil~$e_x$
since it cannot reach these ``newer'' parts of the future graph.
In this case~$x$ can be ``closed''.
The technical machinery to maintain this information for PDL is
from~\cite{gore-widmann-pdl-cade09}.
However, the presence of ``converse'' and the resulting need for alternative nodes
requires a more elaborate scheme for CPDL.
\section{The Algorithm}
Our algorithm builds a directed graph~$G$
consisting of nodes and directed edges.
We first explain the structure of~$G$ in more detail.
\begin{definition}
Let~$X$ and~$Y$ be sets.
We define~$\opt{X} := X \uplus \{ \bot \}$
where~$\bot$ indicates the undefined value and~$\uplus$ is the disjoint union.
If~$f: X \to Y$ is a function and~$x \in X$ and~$y \in Y$
then the function~$\updfkt{f}{x}{y}: X \to Y$ is defined as
$\updfkt{f}{x}{y}(x') := y$ if~$x' = x$
and $\updfkt{f}{x}{y}(x') := f(x')$ if~$x' \not= x$.
\end{definition}
\begin{definition}
\label{def:nodes}
Let~$G = (V, E)$ be a graph
where~$V$ is a set of nodes and~$E$ is a set of directed edges.
Each node~$x \in V$ has six attributes:
$\Gamma_x \subseteq \mathrm{Fml}$, $\mathop{\mathrm{ann}}_x: \mathrm{Ev} \to \opt{\mathrm{Fml}}$,
$\mathrm{pst}_x \in \opt{V}$, $\mathrm{ppr}_x \in \opt{\mathrm{LPrg}}$,
$\mathrm{idx}_x \in \opt{Nat}$, and~$\mathrm{sts}_x \in \mathfrak{S}$
where~$\mathfrak{S} := \{ \mathtt{unexp}, \mathtt{undef} \} \cup \{ \mathop{\mathtt{closed}}(\mathrm{alt}) \mid \mathrm{alt} \subseteq \pows{\mathrm{Fml}} \} \cup
\{ \mathop{\mathtt{open}}(\mathop{\mathrm{prs}},\mathrm{alt}) \mid \mathop{\mathrm{prs}}: \mathrm{Ev} \to \opt{(\pows{V \times \mathrm{Ev}})} \;\&\; \mathrm{alt} \subseteq \pows{\mathrm{Fml}} \}$.
Each directed edge~$e \in E$ is labelled with a label~$l_e \in \opt{(\mathrm{Fml} \cup \pows{\mathrm{Fml}} \cup \{ \mathrm{cs} \})}$
where~$\mathrm{cs}$ is just a constant.
\end{definition}
All attributes of a node~$x \in V$ are initially set at the creation of~$x$,
possibly with the value~$\bot$ (if allowed).
Only the attributes~$\mathrm{idx}_x$ and~$\mathrm{sts}_x$ are changed at a later time.
We use the function~$\mathop{\mbox{create-new-node}}(\Gamma, \mathop{\mathrm{ann}}, \mathrm{pst}, \mathrm{ppr}, \mathrm{idx}, \mathrm{sts})$
to create a new node and initialise its attributes in the obvious way.
The finite set~$\Gamma_x$ contains the formulae which are assigned to~$x$.
The attribute~$\mathop{\mathrm{ann}}_x$ is defined for the eventualities in~$\Gamma_x$ at most.
If~$\mathop{\mathrm{ann}}_x(\varphi) = \varphi'$ then~$\varphi' \in \Gamma_x$ and~$\varphi \mathrel{\rightsquigarrow} \varphi'$.
The intuitive meaning is that~$\varphi$ has already been ``reduced'' to~$\varphi'$ in~$x$.
For a state (as defined below) we always have that~$\mathop{\mathrm{ann}}_x$ is undefined everywhere
since we do not need the attribute for states.
The node~$x$ is called a \emph{state} iff both attributes~$\mathrm{pst}_x$ and~$\mathrm{ppr}_x$ are undefined.
For all other nodes, the attribute~$\mathrm{pst}_x$ identifies the,
as we will ensure, unique ancestor~$p \in V$ of~$x$
such that~$p$ is a state and there is no other state between~$p$ and~$x$ in~$G$.
We call~$p$ the \emph{\underline{p}arent \underline{st}ate} of~$x$.
The creation of the child of~$p$ which lies on the path from~$p$ to~$x$ (it could be~$x$)
was caused by a $\pea{\mathrm{lp}}{}$-formula~$\pea{l}{\varphi}$ in~$\Gamma_p$.
The literal program~$l$ which we call the \emph{\underline{p}arent \underline{pr}ogram} of~$x$
is stored in~$\mathrm{ppr}_x$.
Hence, for nodes which are not states, both~$\mathrm{pst}_x$ and~$\mathrm{ppr}_x$ are defined.
The attribute~$\mathrm{sts}_x$ describes the \emph{status} of~$x$.
Unlike the attributes described so far, its value may be modified several times.
The value~$\mathtt{unexp}$, which is the initial value of each node,
indicates that the node has not yet been expanded.
When a node is expanded, its status becomes either~$\mathop{\mathtt{closed}}(\cdot)$
if it contains an immediate contradiction,
or~$\mathtt{undef}$ to indicate that the node has been expanded
but that its ``real'' status is to be determined.
Eventually, the status of each node is set to either~$\mathop{\mathtt{closed}}(\cdot)$ or~$\mathop{\mathtt{open}}(\cdot, \cdot)$.
If the status is~$\mathop{\mathtt{open}}(\cdot, \cdot)$, it might be modified several times later on,
either to~$\mathop{\mathtt{closed}}(\cdot)$ or to~$\mathop{\mathtt{open}}(\cdot, \cdot)$ (with different arguments),
but once it becomes $\mathop{\mathtt{closed}}(\cdot)$, it will never change again.
We call a node \emph{undefined} if its status is~$\mathtt{unexp}$ or~$\mathtt{undef}$
and \emph{defined} otherwise.
Hence a node is undefined initially,
becomes defined eventually, and then never becomes undefined again.
Furthermore, we call~$x$ \emph{closed} iff its status is~$\mathop{\mathtt{closed}}(\mathrm{alt})$
for some~$\mathrm{alt} \subseteq \pows{\mathrm{Fml}}$.
In this case, we define~$\mathrm{alt}_x := \mathrm{alt}$.
We call~$x$ \emph{open} iff its status is~$\mathop{\mathtt{open}}(\mathop{\mathrm{prs}},\mathrm{alt})$
for some~$\mathop{\mathrm{prs}}: \mathrm{Ev} \to \opt{(\pows{V \times \mathrm{Ev}})}$ and some~$\mathrm{alt} \subseteq \pows{\mathrm{Fml}}$.
In this case, we define~$\mathop{\mathrm{prs}}_x := \mathop{\mathrm{prs}}$ and~$\mathrm{alt}_x := \mathrm{alt}$.
To avoid some clumsy case distinctions,
we define~$\mathrm{alt}_x := \emptyset$ if~$x$ is undefined.
The value~$\mathop{\mathtt{closed}}(\mathrm{alt})$ indicates
that the node is ``useless'' for building an interpretation
because it is either unsatisfiable or ``too small''.
In the latter case,
the set~$\mathrm{alt}$ of \emph{alternative sets} contains information about missing formulae.
Finally, the value~$\mathop{\mathtt{open}}(\mathop{\mathrm{prs}}, \mathrm{alt})$ indicates
that there is still hope that~$x$ is ``useful''
and the function~$\mathop{\mathrm{prs}}_x$ contains information about each eventuality~$e_x \in \Gamma_x$
as explained in the overview.
Although~$x$ itself may be useful,
we need its alternative sets in case it becomes closed later on.
Hence it also has a set of alternative sets.
The attribute~$\mathrm{idx}_x$ serves as a time stamp.
It is set to~$\bot$ at creation time of~$x$
and becomes defined when~$x$ becomes defined.
When this happens, the value of~$\mathrm{idx}_x$ is set
such that~$\mathrm{idx}_x > \mathrm{idx}_y$ for all nodes~$y$ which became defined earlier than~$x$.
We define~$y \mathrel{\sqsubset} x$ iff~$\mathrm{idx}_y \not= \bot$ and either~$\mathrm{idx}_x = \bot$ or~$\mathrm{idx}_y < \mathrm{idx}_x$.
Note that~$y \mathrel{\sqsubset} x$ depends on the current state of the graph.
However, once~$y \mathrel{\sqsubset} x$ holds, it will do so for the rest of the time.
To track eventualities, we label an edge between a state and one of its children
by the $\pea{\mathrm{lp}}{}$-formula~$\pea{l}{\varphi}$ which creates this child.
Additionally, we label edges from special nodes (see overview)
to their \underline{c}orresponding \underline{s}tates with the marker~$\mathrm{cs}$.
We also label edges from special nodes to its alternative nodes
with the corresponding alternative set.
\begin{definition}
Let~$\mathop{\mathrm{ann^\bot}}: \mathrm{Ev} \to \opt{\mathrm{Fml}}$ and~$\mathop{\mathrm{prs^\bot}}: \mathrm{Ev} \to \opt{(\pows{V \times \mathrm{Ev}})}$
be the functions which are undefined everywhere.
For a node~$x \in V$ and a label~$l \in \mathrm{Fml} \cup \pows{\mathrm{Fml}} \cup \{ \mathrm{cs} \}$,
let~$\mathop{\mathrm{getChild}}(x, l)$ be the node~$y \in V$
such that there exists an edge~$e \in E$ from~$x$ to~$y$ with~$l_e = l$.
If~$y$ does not exists or is not unique, let the result be~$\bot$.
For a function~$\mathop{\mathrm{prs}}: \mathrm{Ev} \to \opt{(\pows{V \times \mathrm{Ev}})}$,
a node~$x \in V$, and an eventuality~$\varphi \in \mathrm{Ev}$,
we define the set~$\mathop{\mathrm{reach}}(\mathop{\mathrm{prs}}, x, \varphi)$ of eventualities as follows:
\[ \mathop{\mathrm{reach}}(\mathop{\mathrm{prs}}, x, \varphi) \begin{array}[t]{l}
:= \Big\{ \psi \in \mathrm{Ev} \mid
\exists k \in \bbbn_0.\: \exists \varphi_0, \dotsc, \varphi_k \in \mathrm{Ev}.\: \Big( \psi = \varphi_k \;\&\;\\
(x, \varphi_0) \in \mathop{\mathrm{prs}}(\varphi) \;\&\;
\forall i \in \{ 0, \dotsc, k-1 \}.\: (x, \varphi_{i+1}) \in \mathop{\mathrm{prs}}(\varphi_i) \Big) \Big\} \enspace.
\end{array} \]
The function~$\mathop{\mathrm{defer}}: V \times \mathrm{Ev} \to \opt{\mathrm{Fml}}$ is defined as follows:
\begin{displaymath}
\mathop{\mathrm{defer}}(x, \varphi) :=
\left\{
\begin{array}{ll}
\psi & \text{ if }
\begin{array}[t]{l}
\exists k \in \bbbn_0.\: \exists \varphi_0, \dotsc, \varphi_k \in \mathrm{Fml}.\: \Big( \varphi_0 = \varphi \;\&\; \varphi_k = \psi \;\&\;\\
\forall i \in \{ 0, \dotsc, k-1 \}.\: \big( \varphi_i \in \mathrm{Ev} \;\&\; \mathop{\mathrm{ann}}_x(\varphi_i) = \varphi_{i+1} \big) \;\&\;\\
\big( \varphi_k \notin \mathrm{Ev} \text{ or } \mathop{\mathrm{ann}}_x(\varphi_k) = \bot \big) \Big)
\end{array}\\
\bot & \text{ otherwise.}
\end{array}
\right .
\end{displaymath}
\end{definition}
The function~$\mathop{\mathrm{getChild}}(x, l)$ retrieves a particular child of~$x$.
It is easy to see that, during the algorithm,
the child is always unique if it exists.
Intuitively, the function~$\mathop{\mathrm{reach}}(\mathop{\mathrm{prs}}, x, \varphi)$ computes all eventualities
which can be ``reached'' from~$\varphi$ inside~$x$ according to~$\mathop{\mathrm{prs}}$.
If a potential rescuer~$(x, \psi)$ is contained in~$\mathop{\mathrm{prs}}(\varphi)$,
the potential rescuers of~$\psi$ are somehow relevant for~$\varphi$ at~$x$.
Therefore~$\psi$ itself is relevant for~$\varphi$ at~$x$.
The function~$\mathop{\mathrm{reach}}(\mathop{\mathrm{prs}}, x, \varphi)$ computes exactly the transitive closure
of this relevance relation.
Intuitively, the function~$\mathop{\mathrm{defer}}(x, \varphi)$ follows the ``$\mathop{\mathrm{ann}}_x$-chain''.
That is, it computes~$\varphi_1 := \mathop{\mathrm{ann}}_x(\varphi)$, $\varphi_2 := \mathop{\mathrm{ann}}_x(\varphi_1)$, and so on.
There are two possible outcomes.
The first outcome is that we eventually encounter a~$\varphi_k$
which is either not an eventuality or has~$\mathop{\mathrm{ann}}_x(\varphi_k) = \bot$.
Consequently, we cannot follow the ``$\mathop{\mathrm{ann}}_x$-chain'' any more.
In this case we stop and return~$\mathop{\mathrm{defer}}(x, \varphi) := \varphi_k$.
The second outcome is that we can follow the ``$\mathop{\mathrm{ann}}_x$-chain'' indefinitely.
Then, as~$\Gamma_x$ is finite, there must exist a cycle~$\varphi_0, \dotsc, \varphi_n, \varphi_0$ of eventualities
such that~$\mathop{\mathrm{ann}}_x(\varphi_i) = \varphi_{i+1}$ for all~$0 \leq i < n$, and~$\mathop{\mathrm{ann}}_x(\varphi_n) = \varphi_0$.
In this case we say that~$x$ (or~$\Gamma_x$) contains an \emph{``at a world'' cycle}
and return~$\mathop{\mathrm{defer}}(x, \varphi) := \bot$.
Next we comment on all procedures given in pseudocode.
\begin{procedure}[t]
\caption{\texttt{is-sat}{}($\phi$) for testing whether a formula~$\phi$ is satisfiable}
\dontprintsemicolon
\SetVline
\Input{a formula~$\phi \in \mathrm{Fml}$ in negation normal form}
\Output{true iff~$\phi$ is satisfiable}
\BlankLine
$G :=$ a new empty graph; \quad $\mathrm{idx} := 1$ \;
let~$d \in \mathrm{APrg}$ be a dummy atomic program which does not occur in~$\phi$ \;
$\mathrm{rt} := \mathop{\mbox{create-new-node}}(\{ \pea{d}{\phi} \}, \mathop{\mathrm{ann^\bot}}, \bot, \bot, \bot, \mathtt{unexp})$ \;
insert~$\mathrm{rt}$ in~$G$ \;
\While{one of the rules in Table~\ref{tabd:rules} is applicable}{
apply any one of the applicable rules in Table~\ref{tabd:rules} \;
}
\lIf{$\mathrm{sts}_\mathrm{rt} = \mathop{\mathtt{open}}(\cdot,\cdot)$}{\Return true} \lElse{\Return false} \;
\end{procedure}
\begin{table}[t]
\caption{Rules used in the procedure~\texttt{is-sat}{}}
\label{tabd:rules}
\centering
\begin{tabular}{|ll|}
\hline
Rule~1: & Some node~$x$ has not been expanded yet.\\
Condition: & $\exists x \in V.\: \mathrm{sts}_x = \mathtt{unexp}$\\
Action: & \texttt{expand}($x$)\\
\hline
Rule~2: & The status of some node~$x$ is still undefined.\\
Condition: & $\exists x \in V.\: \mathrm{sts}_x = \mathtt{undef}$\\
Action: & $\mathrm{sts}_x :=$ \texttt{det-status}($x$) \;\&\; $\mathrm{idx}_x := \mathrm{idx}$ \;\&\; $\mathrm{idx} := \mathrm{idx} + 1$\\
\hline
Rule~3: & Some open node~$x$ is not up-to-date.\\
Condition: & $\exists x \in V.\: \mathop{\mathtt{open}}(\cdot,\cdot) = \mathrm{sts}_x \not= $ \texttt{det-status}($x$)\\
Action: & $\mathrm{sts}_x :=$ \texttt{det-status}($x$)\\
\hline
Rule~4: & All nodes are up-to-date, and some~$x$ has an unfulfilled eventuality~$\varphi$.\\
Condition: & Rule~3 is not applicable and\\
& $\exists x \in V.\: \mathrm{sts}_x = \mathop{\mathtt{open}}(\mathop{\mathrm{prs}}_x, \mathrm{alt}_x) \;\&\; \exists \varphi \in \mathrm{Ev} \cap \Gamma_x.\: \mathop{\mathrm{prs}}_x(\varphi) = \emptyset$\\
Action: & $\mathrm{sts}_x := \mathop{\mathtt{closed}}(\mathrm{alt}_x)$\\
\hline
\end{tabular}
\end{table}
\noindent\textbf{Procedure}{}~\texttt{is-sat}{}($\phi$) is invoked to determine
whether a formula~$\phi \in \mathrm{Fml}$ in negation normal form is satisfiable.
It creates a root node~$\mathrm{rt}$
and initialises the graph~$G$ to contain only~$\mathrm{rt}$.
The dummy program~$d$ is used to make~$\mathrm{rt}$ a state
so that each node in~$G$ which is not a state has a parent state.
The global variable~$\mathrm{idx}$ is used to set the time stamps of the nodes accordingly.
While at least one of the rules in Table~\ref{tabd:rules} is applicable,
that is its condition is true,
the algorithm applies any applicable rule.
If no rules are applicable,
the algorithm returns satisfiable iff~$\mathrm{rt}$ is open.
Rule~1 picks an unexpanded node and expands it.
Rule~2 picks an expanded but undefined node and computes its (initial) status.
It also sets the correct time stamp.
Rule~3 picks an open node whose status has changed and recomputes its status.
Its meaning is, that if we compute \texttt{det-status}($x$) on the current graph
then its result is different from the value in~$\mathrm{sts}_x$,
and consequently, we update~$\mathrm{sts}_x$ accordingly.
Rule~4 is only applicable if all nodes are up-to-date.
It picks an open node containing an eventuality~$\varphi$
which is currently not fulfilled in the graph
and which does not have any potential rescuers either.
As this indicates that~$\varphi$ can never be fulfilled,
the node is closed.
This description leaves several questions open, most notably:
``How do we check efficiently whether Rule~3 is applicable?'' and
``Which rule should be taken if several rules are applicable?''.
We address these issues in Section~\ref{secd:strategy}.
\begin{procedure}[t]
\caption{\texttt{expand}{}($x$) for expanding a node~$x$}
\dontprintsemicolon
\SetVline
\Input{a node~$x \in V$ with~$\mathrm{sts}_x = \mathtt{unexp}$}
\BlankLine
\If{$\exists\varphi \in \Gamma_x.\: \pneg{\varphi} \in \Gamma_x \text{ or } (\varphi \in \mathrm{Ev} \;\&\; \mathop{\mathrm{defer}}(x, \varphi) = \bot)$}{
$\mathrm{idx}_x := \mathrm{idx}$; \quad $\mathrm{idx} := \mathrm{idx} + 1$; \quad $\mathrm{sts}_x := \mathop{\mathtt{closed}}(\emptyset)$ \;
}
\Else($(\!*${} $x$ does not contain a contradiction $*\!)${}){
$\mathrm{sts}_x := \mathtt{undef}$ \;
\uIf($(\!*${} $x$ is a state $*\!)${}){$\mathrm{pst}_x = \bot$}{
let~$\pea{l_1}{\varphi_1}, \dotsc , \pea{l_k}{\varphi_k}$ be all of the $\pea{\mathrm{lp}}{}$-formulae in~$\Gamma_x$ \;
\For{$i \longleftarrow 1$ \KwTo $k$}{
$\Gamma_i := \{ \varphi_i \} \cup \{ \psi \mid \paa{l_i}{\psi} \in \Gamma_x \}$ \;
$y_i := \mathop{\mbox{create-new-node}}(\Gamma_i, \mathop{\mathrm{ann^\bot}}, x, l_i, \bot, \mathtt{unexp})$ \;
insert~$y_i$, and an edge from~$x$ to~$y_i$ labelled with~$\pea{l_i}{\varphi_i}$, into~$G$ \;
}
}
\uElseIf{$\exists\alpha \in \Gamma_x.\: \{ \alpha_1, \dotsc, \alpha_k \} \not\subseteq \Gamma_x \text{ or } (\alpha \in \mathrm{Ev} \;\&\; \mathop{\mathrm{ann}}_x(\alpha) = \bot)$}{
$\Gamma := \Gamma_x \cup \{ \alpha_1, \dotsc, \alpha_k \}$ \;
$\mathop{\mathrm{ann}} :=$ \lIf{$\alpha \in \mathrm{Ev}$}{$\updfkt{\mathop{\mathrm{ann}}_x}{\alpha}{\alpha_1}$} \lElse{$\mathop{\mathrm{ann}}_x$} \;
$y := \mathop{\mbox{create-new-node}}(\Gamma, \mathop{\mathrm{ann}}, \mathrm{pst}_x, \mathrm{ppr}_x, \bot, \mathtt{unexp})$ \;
insert~$y$, and an edge from~$x$ to~$y$, into~$G$ \;
}
\uElseIf{$\exists\beta \in \Gamma_x.\: \{ \beta_1, \beta_2 \} \cap \Gamma_x = \emptyset \text{ or } (\beta \in \mathrm{Ev} \;\&\; \mathop{\mathrm{ann}}_x(\beta) = \bot)$}{
\For{$i \longleftarrow$ 1 \KwTo 2}{
$\Gamma_i := \Gamma_x \cup \{ \beta_i \}$ \;
$\mathop{\mathrm{ann}}_i :=$ \lIf{$\beta \in \mathrm{Ev}$}{$\updfkt{\mathop{\mathrm{ann}}_x}{\beta}{\beta_i}$} \lElse{$\mathop{\mathrm{ann}}_x$} \;
$y_i := \mathop{\mbox{create-new-node}}(\Gamma_i, \mathop{\mathrm{ann}}_i, \mathrm{pst}_x, \mathrm{ppr}_x, \bot, \mathtt{unexp})$ \;
insert~$y_i$, and an edge from~$x$ to~$y_i$, into~$G$ \;
}
}
\Else($(\!*${} $x$ is a special node $*\!)${}){
\uIf($(\!*${} state already exists in~$G$ $*\!)${}){
$\exists y \in V.\: \Gamma_y = \Gamma_x \;\&\; \mathrm{pst}_y = \bot$
}{
insert an edge from~$x$ to~$y$ labelled with~$\mathrm{cs}$ into~$G$ \;
}
\Else($(\!*${} state does not exist in~$G$ yet $*\!)${}){
$y := \mathop{\mbox{create-new-node}}(\Gamma_x, \mathop{\mathrm{ann^\bot}}, \bot, \bot, \bot, \mathtt{unexp})$ \;
insert~$y$, and an edge from~$x$ to~$y$ labelled with~$\mathrm{cs}$, into~$G$ \;
}
}
}
\end{procedure}
\noindent\textbf{Procedure}{}~\texttt{expand}{}($x$) expands a node~$x$.
If~$\Gamma_x$ contains an immediate contradiction or an ``at a world'' cycle
then we close~$x$ and set the time stamp accordingly.
For the other cases, we assume implicitly that~$\Gamma_x$ does not contain either of these.
If~$x$ is a state, that is~$\mathrm{pst}_x = \bot$,
then we do the following for each $\pea{\mathrm{lp}}{}$-formula~$\pea{l_i}{\varphi_i}$.
We create a new node~$y_i$
whose associated set contains~$\varphi_i$ and all~$\psi$ such that~$\paa{l_i}{\psi} \in \Gamma_x$.
As none of the eventualities in~$\Gamma_{y_i}$ is reduced yet, there are no annotations.
The parent state of~$y_i$ is obviously~$x$ and its parent program is~$l_i$.
In order to relate~$y_i$ to~$\pea{l_i}{\varphi_i}$,
we label the edge from~$x$ to~$y_i$ with~$\pea{l_i}{\varphi_i}$.
We call~$y_i$ the \emph{successor} of~$\pea{l_i}{\varphi_i}$.
If~$x$ is not a state and~$\Gamma_x$ contains an $\alpha$-formula~$\alpha$
whose decompositions are not in~$\Gamma_x$,
or which is an unannotated eventuality,
we call~$x$ an \emph{$\alpha$-node}.
In this case, we create a new node~$y$
whose associated set is the result of adding all decompositions of~$\alpha$ to~$\Gamma_x$.
If~$\alpha$ is an eventuality
then~$\mathop{\mathrm{ann}}_y$ extends~$\mathop{\mathrm{ann}}_x$ by mapping~$\alpha$ to~$\alpha_1$.
The parent state and the parent program of~$y$ are inherited from~$x$.
Note that~$\mathrm{pst}_x$ and~$\mathrm{ppr}_x$ are defined as~$x$ is not a state.
Also note that~$\Gamma_y \supsetneq \Gamma_x$
or~$\alpha$ is an eventuality which is annotated in~$\mathop{\mathrm{ann}}_y$ but not in~$\mathop{\mathrm{ann}}_x$.
If~$x$ is neither a state nor an $\alpha$-node
and~$\Gamma_x$ contains a $\beta$-formula~$\beta$
such that neither of its immediate subformulae is in~$\Gamma_x$,
or such that~$\beta$ is an unannotated eventuality,
we call~$x$ a \emph{$\beta$-node}.
For each decomposition~$\beta_i$ we do the following.
We create a new node~$y_i$
whose associated set is the result of adding~$\beta_i$ to~$\Gamma_x$.
If~$\beta$ is an eventuality
then~$\mathop{\mathrm{ann}}_{y_i}$ extends~$\mathop{\mathrm{ann}}_x$ by mapping~$\alpha$ to~$\beta_i$.
The parent state and the parent program of~$y$ are inherited from~$x$.
Note that~$\mathrm{pst}_x$ and~$\mathrm{ppr}_x$ are defined as~$x$ is not a state.
Also note that~$\Gamma_{y_i} \supsetneq \Gamma_x$
or~$\beta$ is an eventuality which is annotated in~$\mathop{\mathrm{ann}}_{y_i}$ but not in~$\mathop{\mathrm{ann}}_x$.
If~$x$ is neither a state nor an $\alpha$-node nor a $\beta$-node,
it must be fully saturated and we call it a \emph{special node}.
Intuitively, a special node sits between a saturation phase and a state
and is needed to handle the ``special'' issue arising from converse programs,
as explained in the overview.
Like $\alpha$- and $\beta$-nodes,
special nodes have a unique parent state and a unique parent program.
In this case we check whether there already exists a state~$y$ in~$G$
which has the same set of formulae as the special node.
If such a state~$y$ exists, we link~$x$ to~$y$;
else we create such a state and link~$x$ to it.
In both cases we label the edge with the marker~$\mathrm{cs}$
since a special node can have several children (see below)
and we want to uniquely identify the $\mathrm{cs}$-child~$y$ of~$x$.
Note that there is only at most one state for each set of formulae
and that states are always fully saturated since special nodes are.
\begin{procedure}[t]
\caption{\texttt{det-status}{}($x$) for determining the status of a node~$x$}
\dontprintsemicolon
\SetVline
\Input{a node~$x \in V$ with~$\mathtt{unexp} \not= \mathrm{sts}_x \not= \mathop{\mathtt{closed}}(\cdot)$}
\BlankLine
\lIf{$x$ is an $\alpha$-or a $\beta$-node}{$\mathrm{sts}_x :=$ \texttt{det-sts-$\beta$}($x$)} \;
\lElseIf{$x$ is a state}{$\mathrm{sts}_x :=$ \texttt{det-sts-state}($x$)} \;
\Else($(\!*${} $x$ is a special node, in particular~$\mathrm{pst}_x \not= \bot \not= \mathrm{ppr}_x$ $*\!)${}){
$\Gamma_\mathrm{alt} := \{ \varphi \mid \paa{\pnnp{\mathrm{ppr}_x}}{\varphi} \in \Gamma_x \} \setminus \Gamma_{\mathrm{pst}_x}$ \;
\lIf{$\Gamma_\mathrm{alt} = \emptyset$}{$\mathrm{sts}_x :=$ \texttt{det-sts-spl}($x$)}
\lElse{$sts_x := \mathop{\mathtt{closed}}(\Gamma_\mathrm{alt})$} \;
}
\end{procedure}
\noindent\textbf{Procedure}{}~\texttt{det-status}{}($x$) determines the current status of a node~$x$.
Its result will always be~$\mathop{\mathtt{closed}}(\cdot)$ or~$\mathop{\mathtt{open}}(\cdot,\cdot)$.
If~$x$ is an $\alpha$/$\beta$-node or a state,
the procedure just calls the corresponding sub-procedure.
If~$x$ is a special node,
we determine the set~$\Gamma_\mathrm{alt}$ of all formulae~$\varphi$
such that~$\paa{\pnnp{\mathrm{ppr}_x}}{\varphi}$ is in~$\Gamma_x$
but~$\varphi$ is not in the set of the parent state of~$x$.
If there is no such formula, that is~$\Gamma_\mathrm{alt}$ is the empty set,
we say that~$x$ is \emph{compatible} with its parent state~$\mathrm{pst}_x$.
Note that incompatibilities can only arise because of converse programs.
If~$x$ is compatible with~$\mathrm{pst}_x$, all is well,
so we determine its status via the corresponding sub-procedure.
Else we cannot connect~$\mathrm{pst}_x$ to a state
with~$\Gamma_x$ assigned to it in the putative model as explained in the overview,
and, thus, we can close~$x$.
That does not, however, mean that~$\mathrm{pst}_x$ is unsatisfiable;
maybe it is just missing some formulae.
We cannot extend~$\mathrm{pst}_x$ directly
as this may have side-effects elsewhere;
but to tell~$\mathrm{pst}_x$ what went wrong, we remember~$\Gamma_\mathrm{alt}$.
The meaning is that if we create an alternative node for~$\mathrm{pst}_x$
by adding the formulae in~$\Gamma_\mathrm{alt}$,
we might be more successful in building an interpretation.
\begin{procedure}[t]
\caption{\texttt{det-sts-$\beta$}{}($x$) for determining the status of an $\alpha$- or a $\beta$-node}
\dontprintsemicolon
\SetVline
\Input{an $\alpha$- or a $\beta$-node~$x \in V$ with~$\mathtt{unexp} \not= \mathrm{sts}_x \not= \mathop{\mathtt{closed}}(\cdot)$}
\Output{the new status of~$x$}
\BlankLine
let~$y_1, \dotsc, y_k \in V$ be all children of~$x$ \;
$\mathrm{alt} := \bigcup_{i=1}^k \mathrm{alt}_{y_i}$ \;
\lIf{$\forall i \in \{ 1, \dotsc, k \}.\: \mathrm{sts}_{y_i} = \mathop{\mathtt{closed}}(\cdot)$}{\Return $\mathop{\mathtt{closed}}(\mathrm{alt})$} \;
\Else($(\!*${} at least one child is not closed $*\!)${}){
$\mathop{\mathrm{prs}} := \mathop{\mathrm{prs^\bot}}$ \;
\ForEach{$\varphi \in \Gamma_x \cap \mathrm{Ev}$}{
\lFor{$i \longleftarrow 1$ \KwTo $k$}{$\Lambda_i :=$ \texttt{det-prs-child}($x, y_i, \varphi$)} \;
$\Lambda :=$ \lIf{$\exists i \in \{ 1, \dotsc, k \}.\: \Lambda_i = \bot$}{$\bot$} \lElse{$\bigcup_{i=1}^k \Lambda_i$} \;
$\mathop{\mathrm{prs}} := \updfkt{\mathop{\mathrm{prs}}}{\varphi}{\Lambda}$ \;
}
$\mathop{\mathrm{prs}}' :=$ \texttt{filter}($x, \mathop{\mathrm{prs}}$) \;
\Return $\mathop{\mathtt{open}}(\mathop{\mathrm{prs}}', \mathrm{alt})$ \;
}
\end{procedure}
\noindent\textbf{Procedure}{}~\texttt{det-sts-$\beta$}{}($x$)
computes the status of an $\alpha$- or a $\beta$-node~$x \in V$.
For this task, an $\alpha$-node can be seen as a $\beta$-node with exactly one child.
The set of alternative sets of~$x$
is the union of the sets of alternative sets of all children.
If all children of~$x$ are closed then~$x$ must also be closed.
Otherwise we compute the set of potential rescuers
for each eventuality~$\varphi$ in~$\Gamma_x$ as follows.
For each child~$y_i$ of~$x$
we determine the potential rescuers of~$\varphi$ which result from following~$y_i$
by invoking \texttt{det-prs-child}{}.
If the set of potential rescuers corresponding to some~$y_i$ is~$\bot$
then~$\varphi$ can currently be fulfilled via~$y_i$ and~$\mathop{\mathrm{prs}}_x(\varphi)$ is set to~$\bot$.
Else~$\varphi$ cannot currently be fulfilled in~$G$,
but each child returned a set of potential rescuers,
and the set of potential rescuers for~$\varphi$ is their union.
Finally, we deal with potential rescuers in~$\mathop{\mathrm{prs}}$
of the form~$(x, \chi)$ for some~$\chi \in \mathrm{Ev}$ by calling \texttt{filter}{}.
\begin{procedure}[t]
\caption{\texttt{det-sts-state}{}($x$) for determining the status of a state}
\dontprintsemicolon
\SetVline
\Input{a state~$x \in V$ with~$\mathtt{unexp} \not= \mathrm{sts}_x \not= \mathop{\mathtt{closed}}(\cdot)$}
\Output{the new status of~$x$}
\BlankLine
let~$\pea{l_1}{\varphi_1}, \dotsc , \pea{l_k}{\varphi_k}$ be all of the $\pea{\mathrm{lp}}{}$-formulae in~$\Gamma_x$ \;
\lFor{$i \longleftarrow 1$ \KwTo $k$}{$y_i := \mathop{\mathrm{getChild}}(x, \pea{l_i}{\varphi_i})$} \;
\lIf{$\exists i \in \{ 1, \dotsc, k \}.\: \mathrm{sts}_{y_i} = \mathop{\mathtt{closed}}(\mathrm{alt})$}{\Return $\mathop{\mathtt{closed}}(\mathrm{alt})$} \;
\Else($(\!*${} no child is closed $*\!)${}){
$\mathrm{alt} := \bigcup_{i=1}^k \mathrm{alt}_{y_i}$ \;
$\mathop{\mathrm{prs}} := \mathop{\mathrm{prs^\bot}}$ \;
\For{$i \longleftarrow 1$ \KwTo $k$}{
\If{$\varphi_i \in \mathrm{Ev}$}{
$\Lambda :=$ \texttt{det-prs-child}($x, y_i, \varphi_i$) \;
$\mathop{\mathrm{prs}} := \updfkt{\mathop{\mathrm{prs}}}{\pea{l_i}{\varphi_i}}{\Lambda}$ \;
}
}
$\mathop{\mathrm{prs}}' :=$ \texttt{filter}($x, \mathop{\mathrm{prs}}$) \;
\Return $\mathop{\mathtt{open}}(\mathop{\mathrm{prs}}', \mathrm{alt})$ \;
}
\end{procedure}
\noindent\textbf{Procedure}{}~\texttt{det-sts-state}{}($x$) computes the status of a state~$x \in V$.
We obtain the successors for all $\pea{\mathrm{lp}}{}$-formulae in~$\Gamma_x$.
If any successor is closed then~$x$ is closed with the same set of alternative sets.
Else the set of alternative sets of~$x$
is the union of the sets of alternative sets of all children
and we compute the potential rescuers
for each eventuality~$\pea{l_i}{\varphi_i}$ in~$\Gamma_x$
by invoking \texttt{det-prs-child}{}.
Finally, we deal with potential rescuers in~$\mathop{\mathrm{prs}}$
of the form~$(x, \chi)$ for some~$\chi \in \mathrm{Ev}$ by calling \texttt{filter}{}.
Note that we do not consider eventualities which are not $\pea{\mathrm{lp}}{}$-formulae.
The intuitive reason is that the potential rescuers of such eventualities
are determined by following the annotation chain (see below).
However, different special nodes which have the same set, and hence all link to~$x$,
might have different annotations.
Hence we cannot (and do not need to) fix the potential rescuer sets
for eventualities in~$x$ which are not $\pea{\mathrm{lp}}{}$-formulae.
\begin{procedure}[t]
\caption{\texttt{det-sts-spl}($x$) for determining the status of a special node}
\dontprintsemicolon
\SetVline
\Input{a special node~$x \in V$ with~$\mathtt{unexp} \not= \mathrm{sts}_x \not= \mathop{\mathtt{closed}}(\cdot)$}
\Output{the new status of~$x$}
\BlankLine
$y_0 := \mathop{\mathrm{getChild}}(x, \mathrm{cs})$ \;
let~$\Gamma_1, \dotsc , \Gamma_{j}$ be all the sets in the set~$\mathrm{alt}_{y_0}$ \;
\For{$i \longleftarrow 1$ \KwTo $j$}{
$y_i := \mathop{\mathrm{getChild}}(x, \Gamma_i)$ \;
\If($(\!*${} child does not exist $*\!)${}){$y_i = \bot$}{
$y_i := \mathop{\mbox{create-new-node}}(\Gamma_x \cup \Gamma_i, \mathop{\mathrm{ann}}_x, \mathrm{pst}_x, \mathrm{ppr}_x, \bot, \mathtt{unexp})$ \;
insert~$y_i$, and an edge from~$x$ to~$y_i$ labelled with~$\Gamma_i$, into~$G$ \;
}
}
let~$y_{j+1}, \dotsc, y_k$ be all the remaining children of~$x$ \;
$\mathrm{alt} := \bigcup_{i=1}^k \mathrm{alt}_{y_i}$ \;
\lIf{$\forall i \in \{ 0, \dotsc, k \}.\: \mathrm{sts}_{y_i} = \mathop{\mathtt{closed}}(\cdot)$}{\Return $\mathop{\mathtt{closed}}(\mathrm{alt})$} \;
\Else($(\!*${} at least one child is not closed $*\!)${}){
$\mathop{\mathrm{prs}} := \mathop{\mathrm{prs^\bot}}$ \;
\ForEach{$\varphi \in \Gamma_x \cap \mathrm{Ev}$}{
$\varphi' := \mathop{\mathrm{defer}}(x, \varphi)$ \;
\If{$\varphi' \in \mathrm{Ev}$}{
\lFor{$i \longleftarrow 0$ \KwTo $k$}{$\Lambda_i :=$ \texttt{det-prs-child}($x, y_i, \varphi'$)} \;
$\Lambda :=$ \lIf{$\exists i \in \{ 0, \dotsc, k \}.\: \Lambda_i = \bot$}{$\bot$} \lElse{$\bigcup_{i=0}^k \Lambda_i$} \;
$\mathop{\mathrm{prs}} := \updfkt{\mathop{\mathrm{prs}}}{\varphi}{\Lambda}$ \;
}
}
$\mathop{\mathrm{prs}}' :=$ \texttt{filter}($x, \mathop{\mathrm{prs}}$) \;
\Return $\mathop{\mathtt{open}}(\mathop{\mathrm{prs}}', \mathrm{alt})$ \;
}
\end{procedure}
\noindent\textbf{Procedure}{}~\texttt{det-sts-spl}($x$)
computes the status of a special node~$x \in V$.
First, we retrieve the state~$y_0$ corresponding to~$x$,
namely the unique $\mathrm{cs}$-child of~$x$.
For all alternative sets~$\Gamma_i$ of~$y_0$ we do the following.
If there does not exist a child of~$x$
such that the corresponding edge is labelled with~$\Gamma_i$,
we create a new node~$y_i$
whose associated set is the result of adding the formulae in~$\Gamma_i$ to~$\Gamma_x$.
The annotations, the parent state, and the parent program of~$y_i$ are inherited from~$x$.
We label the new edge from~$x$ to~$y_i$ with~$\Gamma_i$.
In other words we unpack the information
stored in the alternative sets in~$\mathrm{alt}_{y_0}$ into actual nodes
which are all children of~$x$.
Note that each~$\Gamma_i \neq \emptyset$ by construction in \texttt{det-status}{}.
Some children of~$x$ may not be referenced from~$\mathrm{alt}_{y_0}$,
but we consider them anyway.
\begin{procedure}[p]
\caption{\texttt{det-prs-child}{}($x, y, \varphi$) for passing a $\mathop{\mathrm{prs}}$-entry of a child to a parent}
\dontprintsemicolon
\SetVline
\Input{two nodes~$x, y \in V$ and a formula~$\varphi \in \Gamma_y \cap \mathrm{Ev}$}
\Output{$\bot$ or a set of node-formula pairs}
\Remark{if \texttt{det-prs-child}{}($x, y, \varphi$) has been invoked before
with exactly the same arguments and
\emph{under the same invocation of \mbox{\rm \texttt{det-sts-$\beta$}{}}, \mbox{\rm \texttt{det-sts-state}{}}
or \mbox{\rm \texttt{det-sts-spl}{}}},
the procedure is not executed a second time
but returns the cached result of the first invocation.
We do not model this behaviour explicitly in the pseudocode.}
\BlankLine
\lIf{$\mathrm{sts}_y = \mathop{\mathtt{closed}}(\cdot)$}{\Return $\emptyset$} \;
\lElseIf{$\mathrm{sts}_y = \mathtt{unexp} \text{ or } \mathrm{sts}_y = \mathtt{undef} \text{ or } \text{not } y \mathrel{\sqsubset} x$}{\Return $\{ (y, \varphi) \}$} \;
\Else($(\!*${} \mbox{$\mathrm{sts}_y = \mathop{\mathtt{open}}(\cdot,\cdot) \;\&\; y \mathrel{\sqsubset} x$} $*\!)${}){
\lIf{$\mathop{\mathrm{prs}}_y(\varphi) = \bot$}{\Return $\bot$} \;
\Else($(\!*${} \mbox{$\mathop{\mathrm{prs}}_y(\varphi)$} is defined $*\!)${}){
let~$(z_1, \varphi_1), \dotsc, (z_k, \varphi_k)$ be all of the pairs in~$\mathop{\mathrm{prs}}_y(\varphi)$\;
\lFor{$i \longleftarrow 1$ \KwTo $k$}{$\Lambda_i :=$ \texttt{det-prs-child}($x, z_i, \varphi_i$)} \;
\lIf{$\exists j \in \{ 1, \dotsc, k \}.\: \Lambda_j = \bot$}{\Return $\bot$} \lElse{\Return $\bigcup_{i=1}^k \Lambda_i$} \;
}
}
\end{procedure}
The set of alternative sets of~$x$
is the union of the sets of alternative sets of all children;
with the exception of~$y_0$
since the alternative sets of~$y_0$ are not related to~$\mathrm{pst}_x$
but affect~$x$ directly as we have seen.
If all children of~$x$ are closed then~$x$ must also be closed.
Otherwise we compute the set of potential rescuers
for each eventuality~$\varphi$ in~$\Gamma_x$ as follows.
First, we determine~$\varphi' := \mathop{\mathrm{defer}}(x, \varphi)$.
Note that~$\varphi'$ is defined
because the special node~$x$ cannot contain an ``at a world'' cycle by definition.
If~$\varphi'$ is not an eventuality
then~$\varphi'$ is fulfilled in~$x$ and~$\mathop{\mathrm{prs}}(\varphi)$ remains~$\bot$.
If~$\varphi'$ is an eventuality, it must be a $\pea{\mathrm{lp}}{}$-formula as~$x$ is a special node.
We use~$\varphi'$ instead of~$\varphi$
since only $\pea{\mathrm{lp}}{}$-formula have a meaningful interpretation in~$\mathop{\mathrm{prs}}_{y_0}$ (see above).
For each child~$y_i$ of~$x$
we determine the potential rescuers of~$\varphi'$ by invoking \texttt{det-prs-child}{}.
If the set of potential rescuers corresponding to some~$y_i$ is~$\bot$
then~$\varphi'$ can currently be fulfilled via~$y_i$ and so~$\mathop{\mathrm{prs}}_x(\varphi)$ is set to~$\bot$.
Otherwise~$\varphi'$ cannot currently be fulfilled in~$G$,
but each child returned a set of potential rescuers,
and the set of potential rescuers for~$\varphi$ is their union.
Finally, we deal with potential rescuers in~$\mathop{\mathrm{prs}}$
of the form~$(x, \chi)$ for some~$\chi \in \mathrm{Ev}$ by calling \texttt{filter}{}.
\noindent\textbf{Procedure}{}~\texttt{det-prs-child}{}($x, y, \varphi$)
determines whether an eventuality~$\psi \in \Gamma_x$,
which is not passed as an argument,
can be fulfilled via~$y$
such that~$\varphi$ is part of the corresponding fulfilling path;
or else which potential rescuers~$\psi$ can reach via~$y$ and~$\varphi$.
If~$y$ is closed, it cannot help to fulfil~$\psi$
as indicated by the empty set.
If~$x$ is undefined or did not become defined before~$x$
then~$(y, \varphi)$ itself is a potential rescuer of~$x$.
Else, if~$\varphi$ can be fulfilled, i.e.~$\mathop{\mathrm{prs}}_y(\varphi) = \bot$,
then~$\psi$ can be fulfilled too, so we return~$\bot$.
Otherwise we invoke the procedure recursively
on all potential rescuers in~$\mathop{\mathrm{prs}}_y(\varphi)$.
If at least one of these invocations returns~$\bot$
then~$\psi$ can be fulfilled via~$y$ and~$\varphi$
and the corresponding rescuer in~$\mathop{\mathrm{prs}}_y(\varphi)$.
If all invocations return a set of potential rescuers,
the set of potential rescuers for~$\psi$ is their union.
The recursion is well-defined
because if~$(z_i, \varphi_i) \in \mathop{\mathrm{prs}}_y(\varphi)$
then either~$z_i$ is still undefined or~$z_i$ became defined later than~$y$.
Each invocation of \texttt{det-prs-child}{} can be uniquely assigned
to the invocation of \texttt{det-sts-$\beta$}{}, \texttt{det-sts-state}{}, or \texttt{det-sts-spl}{}
which (possibly indirectly) invoked it.
To meet our complexity bound,
we require that under the same invocation of \texttt{det-sts-$\beta$}{}, \texttt{det-sts-state}{}, or \texttt{det-sts-spl}{},
the procedure \texttt{det-prs-child}{} is only executed at most once for each argument triple.
Instead of executing it a second time with the same arguments,
it uses the cached result of the first invocation.
Since \texttt{det-prs-child}{} does not modify the graph,
the second invocation would return the same result as the first one.
An easy implementation of the cache is
to store the result of \texttt{det-prs-child}{}($x, y, \varphi$) in the node~$y$
together with~$\varphi$ and a unique id number
for each invocation of \texttt{det-sts-$\beta$}{}, \texttt{det-sts-state}{}, or \texttt{det-sts-spl}{}.
\begin{procedure}[t]
\caption{\texttt{filter}{}($x, \mathop{\mathrm{prs}}$) for handling self-loops in~$\mathop{\mathrm{prs}}$ chains in~$G$}
\dontprintsemicolon
\SetVline
\Input{a node~$x \in V$ and a function~$\mathop{\mathrm{prs}}: \mathrm{Ev} \to \opt{(\pows{V \times \mathrm{Ev}})}$}
\Output{$\mathop{\mathrm{prs}}$ where self-loops have been handled}
\BlankLine
$\mathop{\mathrm{prs}}' := \mathop{\mathrm{prs^\bot}}$ \;
\ForEach{$\varphi \in \Gamma_x \cap \mathrm{Ev}$ such that $\mathop{\mathrm{prs}}(\varphi) \not= \bot$}{
$\Delta := \{ \varphi \} \cup \mathop{\mathrm{reach}}(\mathop{\mathrm{prs}}, x, \varphi)$ \;
\If{$\text{not } \exists \chi \in \Delta.\: \mathop{\mathrm{prs}}(\chi) = \bot$}{
$\Lambda := \bigcup_{\chi \in \Delta} \big\{ (z, \psi) \in \mathop{\mathrm{prs}}(\chi) \mid z \not= x \big\}$\;
$\mathop{\mathrm{prs}}' := \updfkt{\mathop{\mathrm{prs}}'}{\varphi}{\Lambda}$ \;
}
}
\Return $\mathop{\mathrm{prs}}'$ \;
\end{procedure}
\noindent\textbf{Procedure}{}~\texttt{filter}{}($x, \mathop{\mathrm{prs}}$) deals with the potential rescuers
for each eventuality of a node~$x$
which are of the form~$(x, \psi)$ for some~$\psi \in \mathrm{Ev}$.
The second argument of \texttt{filter}{} is a provisional~$\mathop{\mathrm{prs}}$ for~$x$.
If an eventuality~$\varphi \in \Gamma_x$ is currently fulfillable in~$G$
there is nothing to be done,
so let~$(x, \psi) \in \mathop{\mathrm{prs}}(\varphi)$.
If~$\psi = \varphi$ then~$(x, \varphi)$ cannot be a potential rescuer for~$\varphi$ in~$x$
and should not appear in~$\mathop{\mathrm{prs}}(\varphi)$.
But what about potential rescuers of the form~$(x, \psi)$ with~$\psi \not= \varphi$?
Since we want the nodes in the potential rescuers
to become defined later than~$x$,
we cannot keep~$(x, \psi)$ in~$\mathop{\mathrm{prs}}(\varphi)$;
but we cannot just ignore the pair either.
Intuitively~$(x, \psi) \in \mathop{\mathrm{prs}}(\varphi)$ means
that~$\varphi \in \Gamma_x$ can ``reach''~$\psi \in \Gamma_x$ by following a loop in~$G$
which starts at~$x$ and returns to~$x$ itself.
Thus if~$\psi$ can be fulfilled in~$G$, so can~$\varphi$;
and all potential rescuers of~$\psi$ are also potential rescuers of~$\varphi$.
The function~$\mathop{\mathrm{reach}}(\mathop{\mathrm{prs}}, x, \varphi)$ computes all eventualities in~$x$
which are ``reachable'' from~$\varphi$ in the sense above,
where transitivity is taken into account.
That is, it detects all self-loops from~$x$ to itself
which are relevant for fulfilling~$\varphi$.
We add~$\varphi$ as it is not in~$\mathop{\mathrm{reach}}(\mathop{\mathrm{prs}}, x, \varphi)$.
If any of these eventualities is fulfilled in~$G$
then~$\varphi$ can be fulfilled and is consequently undefined in the resulting~$\mathop{\mathrm{prs}}'$.
Otherwise we take all their potential rescuers
whose nodes are not~$x$.
\begin{theorem}[Soundness, Completeness and Complexity]
Let~$\phi \in \mathrm{Fml}$ be a formula in negation normal form of size~$n$.
The procedure $\texttt{is-sat}(\phi)$ terminates, runs in \textsc{exptime}{} in~$n$,
and~$\phi$ is satisfiable iff $\texttt{is-sat}(\phi)$ returns true.
\end{theorem}
\section{Implementation, Optimisations, and Strategy}
\label{secd:strategy}
It should be fairly straightforward to implement our algorithm.
It remains to show
an efficient way to find nodes which are not up-to-date.
It is not too hard to see
that the status of a node~$x$ can become outdated
only if its children change their status or
\texttt{det-prs-child}($x, y, \cdot$) was invoked when~$x$'s previous status was determined
and~$y$ now changes its status.
If we keep track of nodes of the second kind by inserting additional ``update''-edges
as described in \cite{gore-widmann-pdl-cade09},
we can use a queue for all nodes that might need updating.
When the status of a node is modified,
we queue all parents and all nodes linked by ``update''-edges.
We have omitted several refinements from our description for clarity.
The most important is that if a state~$s$ is closed,
all non-states which have~$s$ as a parent state are ignorable
since their status cannot influence any other node~$t$
unless~$t$ also has~$s$ as a parent state.
Moreover, if every special node parent~$x$ of a state~$s'$ is
incompatible or itself has a closed parent state,
then~$s'$ and the nodes having~$s'$ as parent state are ignorable.
This applies transitively,
but if~$s'$ gets a new parent whose parent state is not closed
then~$s'$ becomes ``active'' again.
Another issue is which rule to choose if several are applicable.
As we have seen, it is advantageous to close nodes as early as possible.
Apart from immediate contradictions,
we have Rule~4 which closes a node because it contains an unfulfillable eventuality.
If we can apply Rule~4 early while the graph is still small,
we might prevent big parts of the graph being built needlessly later.
Trying to apply Rule~4 has several consequences on the strategy of how to apply rules.
First, it is important to keep all nodes up-to-date
since Rule~4 is not applicable otherwise.
Second, it is preferable that a node~$x$ cannot reach open nodes
which became defined (or will be defined) after~$x$ did.
Hence, we should try to use Rule~2 on a node only if all children are already defined.
\section{An Example}
To demonstrate how the algorithm works,
we invoke it on the satisfiable toy formula~$\pea{a}{\phi}$ where~$\phi := \pea{\prp{a}}{\paa{\pcv{a}}{p}}$.
To save space, Fig.~\ref{figd:ex1} only shows the core subgraph of the tableau.
Remember that the order of rule applications is not fixed
but the example will follow some of the guidelines given in Section~\ref{secd:strategy}.
\begin{figure}[p]
\centering
\begin{math}
\xymatrix{
*++[F]{
\begin{tabular}{c}
(1) \hfill\phantom{M} state \\
$\{\: \phi, \ \pea{a}{\phi} \:\}$ \\
$\bot, \bot$ \hfill $5$
\end{tabular}
}
\ar[r]^-{\pea{a}{\phi}}
&
*++[F]{
\begin{tabular}{c}
(2) \hfill\phantom{M} $\beta$-node \\
$\{\: \phi \:\}$ \\
$1, a$ \hfill $4$
\end{tabular}
}
\ar[r]
\ar[d]
&
*++[F]{
\begin{tabular}{c}
(3) \hfill\phantom{I} special node \\
$\{\: \phi \mathrel{\rightsquigarrow} \paa{\pcv{a}}{p} \:\}$ \\
$1, a$ \hfill $2$
\end{tabular}
}
\ar[d]_-{\mathrm{cs}}
\\
*++[F]{
\begin{tabular}{c}
(6) \hfill\phantom{I} special node \\
$\{\: \phi \mathrel{\rightsquigarrow} \pea{a}{\phi}, \ p \:\}$ \\
$1, a$ \hfill $9$
\end{tabular}
}
\ar[d]_-{\mathrm{cs}}
&
*++[F]{
\begin{tabular}{c}
(4) \hfill\phantom{I} special node \\
$\{\: \phi \mathrel{\rightsquigarrow} \pea{a}{\phi} \:\}$ \\
$1, a$ \hfill $3$
\end{tabular}
}
\ar[lu]_-{\mathrm{cs}}
\ar[l]_-{\{p\}}
&
*++[F]{
\begin{tabular}{c}
(5) \hfill\phantom{M} state \\
$\{\: \phi, \ \paa{\pcv{a}}{p} \:\}$ \\
$\bot, \bot$ \hfill $1$
\end{tabular}
}
\\
*++[F]{
\begin{tabular}{c}
(7) \hfill\phantom{M} state \\
$\{\: \phi, \ \pea{a}{\phi}, \ p \:\}$ \\
$\bot, \bot$ \hfill $8$
\end{tabular}
}
\ar[r]^-{\pea{a}{\phi}}
&
*++[F]{
\begin{tabular}{c}
(8) \hfill\phantom{M} $\beta$-node \\
$\{\: \phi \:\}$ \\
$7, a$ \hfill $7$
\end{tabular}
}
\ar[r]
\ar[ld]
&
*++[F]{
\begin{tabular}{c}
(9) \hfill\phantom{I} special node \\
$\{\: \phi \mathrel{\rightsquigarrow} \paa{\pcv{a}}{p} \:\}$ \\
$7, a$ \hfill $6$
\end{tabular}
}
\ar[u]^-{\mathrm{cs}}
\\
*++[F]{
\begin{tabular}{c}
(10) \hfill\phantom{I} special node \\
$\{\: \phi \mathrel{\rightsquigarrow} \pea{a}{\phi} \:\}$ \\
$7, a$ \hfill $10$
\end{tabular}
}
\ar`l/20pt[uuu] `[uuu]_-{\mathrm{cs}} [uuu]
\ar[r]^-{\{p\}}
&
*++[F]{
\begin{tabular}{c}
(11) \hfill\phantom{I} special node \\
$\{\: \phi \mathrel{\rightsquigarrow} \pea{a}{\phi}, \ p \:\}$ \\
$7, a$ \hfill $11$
\end{tabular}
}
\ar`u/10pt[l] `[ul]_<<{\mathrm{cs}} [lu]
}
\end{math}
\caption[]{An example: The graph~$G$ just before setting the status of node~(2)}
\label{figd:ex1}
\end{figure}
The nodes in Fig.~\ref{figd:ex1} are numbered in order of creation.
The annotation~$\mathop{\mathrm{ann}}$ is given using~``$\mathrel{\rightsquigarrow}$'' in~$\Gamma$.
For example, in node~(3), we have~$\Gamma_3 = \{\: \phi,\ \paa{\pcv{a}}{p} \:\}$,
and~$\mathop{\mathrm{ann}}_3$ maps the eventuality~$\phi$ to~$\paa{\pcv{a}}{p}$
and is undefined elsewhere.
The bottom line of a node contains the parent state and the parent program on the left,
and the time stamp on the right.
We do not show the status of a node
since it changes during the algorithm, but explain it in the text.
If we write~$\mathrm{sts}_x = \mathop{\mathtt{open}}(\Lambda,\cdot)$ where~$\Lambda \subseteq V \times \mathrm{Ev}$,
we mean that~$\mathop{\mathrm{prs}}_x$ maps all eventualities in~$\Gamma_x$,
with the exception of non-$\pea{\mathrm{lp}}{}$-formulae if~$x$ is a state, to~$\Lambda$
and is undefined elsewhere.
We only consider the core subgraph of~$\phi$ and start by expanding node~(1)
which creates~(2).
Then we expand~(2) and create~(3) and~(4) which are both special nodes.
Next we expand~(3) and create the state~(5).
Expanding~(5) creates no new nodes since~$\Gamma_5$ contains no $\pea{\mathrm{lp}}{}$-formula.
Now we define~(5) and then~(3).
This results in setting~$\mathrm{sts}_5 := \mathop{\mathtt{open}}(\mathop{\mathrm{prs^\bot}}, \emptyset)$ according to \texttt{det-sts-state}{},
and~$\mathrm{sts}_3 := \mathop{\mathtt{closed}}(\{p\})$ since~(3) is not compatible with its parent state~(1).
Expanding~(4) inserts the edge from~(4) to~(1)
and defining~(4) sets~$\mathrm{sts}_4 := \mathop{\mathtt{open}}(\{(1, \pea{a}{\phi})\}, \emptyset)$ according to \texttt{det-sts-spl}{}.
Note that~(6) does not exist yet.
Next we define~(2) and then~(1)
which results in setting~$\mathrm{sts}_2 := \mathop{\mathtt{open}}(\{(1, \pea{a}{\phi})\}, \{p\})$ according to \texttt{det-sts-$\beta$}{}
and~$\mathrm{sts}_1 := \mathop{\mathtt{open}}(\emptyset, \{p\})$ thanks to \texttt{filter}{}.
Note that~$\pea{a}{\phi} \in \Gamma_1$ has an empty set of potential rescuers.
In PDL, we could thus close~(1),
but converse programs complicate matters for CPDL
as reflected by the fact that Rule~4 is not applicable for~(1)
because~(4) is not up-to-date.
Updating~(4) creates~(6) and sets~$\mathrm{sts}_4 := \mathop{\mathtt{open}}(\{(1, \pea{a}{\phi}), (6, \pea{a}{\phi})\}, \emptyset)$.
Updating~(2) and then~(1) sets~$\mathrm{sts}_2 := \mathop{\mathtt{open}}(\{(1, \pea{a}{\phi}), (6, \pea{a}{\phi})\}, \{p\})$
and~$\mathrm{sts}_1 := \mathop{\mathtt{open}}(\{(6, \pea{a}{\phi})\}, \{p\})$.
Now all nodes are up-to-date,
but Rule~4 is not applicable for~(1)
because the set of potential rescuers for~$\phi$ is no longer empty.
Next we expand~(6), which creates~(7),
then~(7), which creates~(8),
then~(8), which creates~(9) and~(10),
and finally~(9), which creates no new nodes.
Node~(9) is similar to~(3), but unlike~(3),
it is compatible with its parent state~(7)
which results in~$\mathrm{sts}_9 := \mathop{\mathtt{open}}(\bot, \emptyset)$.
Using our strategy from the last section,
we would now expand~(10)
so that~(8) can become defined after both its children became defined.
Since~(9) fulfils all its eventualities,
we choose to define~(8) instead and set~$\mathrm{sts}_8 := \mathop{\mathtt{open}}(\bot, \emptyset)$.
Next we define~(7) and then~(6)
which sets~$\mathrm{sts}_7 := \mathop{\mathtt{open}}(\bot, \emptyset)$ and~$\mathrm{sts}_6 := \mathop{\mathtt{open}}(\bot, \emptyset)$.
The status of~(4) is not affected since~(6) was defined after~(4),
giving ``(6) $\not\mathrel{\sqsubset}$ (4)'' in \texttt{det-prs-child}(4, 6, $\pea{a}{\phi}$).
We expand~(10) which inserts the edge from~(10) to~(1).
Then we define~(10) which creates~(11)
and sets~$\mathrm{sts}_{10} := \mathop{\mathtt{open}}(\bot, \emptyset)$.
Note that the invocation of \texttt{det-prs-child}(10, 1, $\pea{a}{\phi}$)
in the invocation \texttt{det-sts-spl}(10)
leads to the recursive invocation \texttt{det-prs-child}(10, 6, $\pea{a}{\phi}$).
Expanding and defining~(11) yields~$\mathrm{sts}_{11} := \mathop{\mathtt{open}}(\bot, \emptyset)$.
Finally, no rule is applicable in the shown subgraph.
\bibliographystyle{splncs}
|
\section{Introduction}
The possible existence of meson-nucleus bound systems is now one of the hot topics in hadron physics. Especially, an antikaon is expected to be bound in a finite nucleus, thanks to a strong attraction between an $I=0$ $\overline{K}N$ pair~\cite{Dote_HNP09}, and recent experimental findings of the signature of such a kaon-nucleus bound state~\cite{FINUDA05, DISTO09} have brought a heated debate on whether it really exists with a decay width narrow enough to be observed experimentally.
In contrast, the interaction between an $\eta$ meson and a nucleon are also known to be also attractive, while its strength is not well-determined. The $\eta N$ scattering length has a large ambiguity, between $0.270\,\mathrm{fm}$ and $1.050\,\mathrm{fm}$ for its real part and between $0.190\,\mathrm{fm}$ and $0.399\,\mathrm{fm}$ for its imaginary part~\cite{Haider02}. It is not possible to carry out neither an X-ray measurement of mesic atom nor a scattering experiment, in order to derive information on $\eta N$ interaction, different from the case for an antikaon. Instead, experimental data of hadronic production and photoproduction of $\eta$ mesons are mainly used in theoretical analyses. It should be noted that the $S_{11}(1535)$ resonance plays a dominant role in near-threshold $\eta$ production, since it lies just about $50\,\mathrm{MeV}$ above the $\eta N$ threshold, and strongly couples with the $\eta N$ channel, as seen in its large decay branching ratio into $\eta+N$ (45--60\%~\cite{PDG08}).
If the existence of $\eta$-mesic nuclei is confirmed experimentally, experimental observables such as the binding energy and decay width will impose a constraint on theoretical models about (in-medium) $\eta N$ interaction and the $S_{11}(1535)$ resonance. Haider and Liu predicted their existence for the first time~\cite{Haider86}, by using the scattering length of $(0.28+0.29i)\,\mathrm{fm}$ or $(0.27+0.22i)\,\mathrm{fm}$, and they found a bound state with the mass number $A \geq 12$ may exist. If the scattering length is larger than they applied, there may be a possibility for lighter $\eta$-mesic nuclei to exist.
In this paper, we discuss experimental ideas to explore $\eta N$ interaction and $\eta$-mesic nuclei by using the ($\pi$, $N$) reactions on different nuclear targets which will be feasible at the J-PARC facility.
\section{($\pi$, $N$) reaction at J-PARC}
Secondary particles, such as pions and kaons, are produced by the bombardment of intense proton beam accelerated by the J-PARC $50\,\mathrm{GeV}$ Proton Synchrotron onto the production target. Then, they are extracted and transported into experimental areas.
At present, there are two beamlines, K1.8 (available momentum up to $\sim 2\,\mathrm{GeV}/c$) and K1.8BR (up to $1.1\,\mathrm{GeV}/c$), and further beamlines are planned. For simplicity, we discuss experimental plans at the K1.8BR area, where most of the detectors for the E15 experiment~\cite{E15Prop} may be reused after a small modification.
The E15 experiment aims to search for a deeply-bound kaonic nuclear state $K^-pp$ by the $^3\mathrm{He}$($K^-$, $n$) reaction. The layout of the experimental area together with the detectors to be installed is shown in Fig.~\ref{fig1}. A $K^-$ is transported through the K1.8BR beamline, and momentum-analyzed. Scattered neutrons by the ($K^-$, $n$) reaction on the helium-3 target are detected by the Neutron Counter, located $15\,\mathrm{m}$ downstream from the target. The decay particles of $K^-pp\to \Lambda +p\to p+\pi^-+p$ are detected by the Cylindrical Detector System which surrounds the target. Their trajectories can be reconstructed by the Cylindrical Drift Chamber installed in a solenoid magnet which operates at $0.5\,\mathrm{T}$ parallel to the beam axis. The Cylindrical Detector Hodoscopes, which locates outside of the CDC, are used for the trigger purpose and the particle identification.
Recently, a new measurement of the $^3\mathrm{He}$($K^-$, $p$) reaction in the E15 experiment, by adding a detector system for scattered protons, was proposed (J-PARC P28~\cite{P28Prop}), and the proposal has been approved as a part of the E15 experiment. Because a beam sweeping magnet is installed downstream of the target, scattered protons are also bent away from the beam axis (see Fig.~\ref{fig1}). Hence, an array of TOF counters for scattered protons will be located next to the charge veto counters for neutrons. In order to enable an inclusive measurement, a set of drift chambers will be installed downstream of the target, so that the trajectories of the protons can be reconstructed.
The present experimental setup is, of course, optimized for the E15 experiment. However, the ($\pi^\pm$, $n$) and/or ($\pi^\pm$, $p$) reactions on a different nuclear target can be investigated with almost the same setup. As discussed below, it is vitally important to detect rescattered particles in the $\pi^++d$ reaction and decay particles from $\eta$-mesic nuclei. Therefore, the usage of the Cylindrical Detector System will be desirable, too.
\begin{figure}[t]
\begin{center}
\psfig{file=k18br.eps,width=4.5in}
\caption{Setup for the J-PARC E15 experiment.}
\label{fig1}
\end{center}
\end{figure}
\section{Baryon resonance production in the $\pi+d$ reaction}
In this section, we consider a \textit{double-scattering reaction}:
\begin{equation}
\pi^++d \to p+p+\eta,\label{eq1}
\end{equation}
where both the protons in the final state are involved (Fig.~\ref{fig2a}). This reaction must be distinguished from a \textit{quasi-free reaction}:
\begin{equation}
\pi^+ +n \to p+\eta,\label{eq2}
\end{equation}
where the proton in the deuteron is a spectator. (Fig.~\ref{fig2b}).
The double-scattering reaction is a two-step process, namely the quasi-free reaction (\ref{eq2}) followed by the rescattering:
\begin{equation}
\eta + p \to \eta + p.\label{eq3}
\end{equation}
The experimental goal is to obtain the missing-mass spectrum for the $d(\pi^+,\ p)X$ reaction with tagging $X$ as the $p\eta$ final state. It is expected that a bump corresponding to the $S_{11}(1535)$ resonance due to the rescattering (\ref{eq3}) will be observed in the spectrum.
Since we are interested in $\eta N$ interaction at low energy, the beam momentum should not be far from the recoilless condition for the $n(\pi^+,\ p)\eta$ reaction. As shown in Fig.~\ref{fig3}, the optimal beam momentum will be around $0.8\, \mbox{--}\,1.1\,\mathrm{GeV}/c$.
The momentum of the $S_{11}(1535)$ ($p^*$) resonance will be almost at rest (Fig.~\ref{fig3}), as expected.
\begin{figure}[t]
\begin{center}
\subfigure[double-scattering reaction]{\psfig{file=reaction1.eps,scale=0.35}\label{fig2a}}
\hspace{0.5cm}
\subfigure[quasi-free reaction]{\raisebox{9.9mm}{\psfig{file=reaction2.eps,scale=0.35}}\label{fig2b}}
\caption{Diagrams for the $\pi^+d\to pp\eta$ reaction.}
\end{center}
\end{figure}
Then, one possibility to separate the double-scattering reaction from the quasi-free one is to detect a proton from the $S_{11}$ decay ($p^*\to p+\eta$). While a proton from the decay of a $S_{11}$ with its mass $1.535\,\mathrm{GeV}/c^2$ has its momentum of $\sim 186\,\mathrm{MeV}/c$, the Fermi momentum of a nucleon in the deuteron hardly exceed $200\,\mathrm{MeV}/c$. Moreover by detecting a decay proton, a missing $\eta$ can be identified by use of the missing-mass of the $d(\pi^+,\ pp)X$ reaction.
From an experimental point of view, the present CDC for the E15 experiment does not allow for the detection of such a slow proton because of the existence of its inner support. For instance, the range of a $200\,\mathrm{MeV}/c$ proton is only $5\,\mathrm{mm}$ in a plastic scintillator. Accordingly, a target has to be thin enough, less than a few mm\footnote{Therefore, a solid $\mathrm{CD}_2$ will be used, instead of liquid $\mathrm{D}_2$.}, and we have to equip a new detector system between the CDC and the target. For example, it is considered to install two barrels of plastic scintillators with different distances from the center; the counter in the inner barrel measures the total energy of a proton, and that in the outer barrel serves as a veto counter to confirm the proton is stopped in the inner counter.
\begin{figure}[t]
\begin{center}
\psfig{file=momentum_transfer_mod.eps,width=2.6in}
\caption{Momentum transfer for the $n$($\pi^+$, $p$)$\eta$ reaction (solid line) and the $d$($\pi^+$, $p$)$p^*$ reaction (dashed lines) for three different $p^*$ masses (1485, 1535, 1585$\,\mathrm{MeV}/c^2$).}
\label{fig3}
\end{center}
\end{figure}
By the way, the $\pi^-+p\to \eta + n$ reaction has been investigated by several groups~\cite{pipetan, Prakhov}. The total cross section increases very rapidly from the threshold (center-of-mass energy $W\sim 1.487\,\mathrm{GeV}$), and reaches more than $2\,\mathrm{mb}$ in the $S_{11}(1535)$ resonance region. While the bump below $W\sim 1.6\,\mathrm\,\mathrm{MeV}$ can be explained by the predominant contribution of the $S_{11}(1535)$ resonance, the cross section above $1.6\,\mathrm{GeV}$ is strongly model-dependent~\cite{pipetan_th}. It is because more resonances, such as $S_{11}(1650)$, $P_{11}(1710)$, $D_{13}(1520)$, $D_{13}(1700)$, $P_{13}(1720)$, and so on, contribute to the reaction.
A similar behavior should hold for the $\pi^++n\to \eta + p$ reaction, due to the isospin symmetry. Therefore, an incident pion momentum around $0.8\,\mathrm{GeV}/c$, which corresponds to $W\sim 1.56\,\mathrm{GeV}$, may be suitable in order to avoid difficulties in model calculations. A measurement with a higher beam momentum may provide information on different resonances, including the second $S_{11}$ resonance.
It should be noted that not only $\eta$ but also $\pi$ and vector mesons ($\rho$, $\omega$) can propagate as an intermediate meson, even if the center-of-mass energy of the pion and the neutron is adjusted to the $S_{11}(1535)$ region. This is also the case for the $p+p\to p+p+\eta$ reaction near the threshold, which has been experimentally studied by COSY-11 and COSY-TOF (see, for example, Ref.~\refcite{Moskal09}). Theoretical calculations for the $p+p$ reaction~\cite{pp_th} take into account these intermediate mesons between two nucleons.
Therefore, a new calculation for the $\pi^++d$ reaction, based on a model for the $\pi^++p$ and $p+p$ reactions, is awaited. This novel reaction might reveal unique features, i.e., a possibility to make a constraint to parameters such as the a $MNN^*$ coupling constant.
It is worth noting that an analogous reaction has been already discussed with regard to the $\Lambda(1405)$ resonance. In the chiral unitary model, it is shown that a $\Lambda(1405)$, which is dynamically generated in the meson-baryon scattering, is a superposition of two poles~\cite{Jido03}. The lower pole ($\sim 1390\,\mathrm{MeV}$) and higher pole ($\sim 1420\,\mathrm{MeV}$) couple strongly with the $\pi\Sigma$ and $\overline{K}N$ channel, respectively. In order to confirm that the resonance in the $\overline{K}N$ channel locates around $1420\,\mathrm{MeV}$ instead of $1405\,\mathrm{MeV}$, the scattering of $\overline{K}N\to\pi\Sigma$ has to be examined. However, it is impossible to produce the $\Lambda(1405)$ resonance in a direct $K^-+p$ scattering, because it lies below the $\overline{K}N$ threshold. Jido~\textit{et al.}~\cite{Jido09} discussed a two-step process of the $K^-+d \to \Lambda(1405) +n$ reaction, where the resonance can be produced by the scattering of an nucleon in the deuteron and an intermediate antikaon, and compared their calculation with an old experiment by Braun \textit{et al.}~\cite{Braun77}\footnote{Since they used a deuteron bubble chamber, the scattered neutron was not detected. Instead, they obtained the $\pi^+\Sigma^-$ invariant-mass spectrum.}\footnote{Not only $\Lambda(1405)+n$, but also $\Sigma^-+p$, $\Sigma^-(1385)+p$, and $\Lambda(1520)+n$ were identified as final states.} In J-PARC, an experimental proposal to investigate the $d(K^-,\ n)$ reaction, together with the identification of the decay mode of $\Lambda(1405)$ into $\pi^0\Sigma^0$ as well as $\pi^\pm \Sigma^\mp$ has been submitted~\cite{P31Prop}.
Turning back to the double-scattering reaction (\ref{eq1}), the situation is different in that the $S_{11}(1535)$ resonance is above the $\eta N$ threshold. Nevertheless, there is a possibility that one can access information on the $\eta N\to \eta N$ scattering, caused by an intermediate $\eta$ meson, which cannot be investigated by a direct reaction with ``$\eta$ meson beam''.
\section{Search for $\eta$-mesic nuclei}
After Haider and Liu predicted the existence of $\eta$-mesic nuclei in 1986~\cite{Haider86}, $\eta$-nucleon and $\eta$-nucleus interaction have been of particular interest both theoretically and experimentally. Observation of $\eta$-mesic nuclei, as well as investigation of near-threshold $\eta$ production, will provide much better understanding of the interaction.
The TAPS/MAMI experiment investigated the photoproduction of $\eta$-mesic $^3\mathrm{He}$ by the $\gamma +{}^3\mathrm{He}\to \eta+X$ and $\pi^0+p+X$ reaction~\cite{Pfeiffer04}. The total cross section of the coherent $\eta$ production exhibits a peak-like structure at the threshold. An enhancement of correlated $\pi^0 p$ pairs with relative angles near $180^\circ$ is also observed near the $\eta$ production threshold, which may be attributed to the decay of $\eta$-mesic nuclei. By a simultaneous fit of the cross sections for the two channels, the binding energy $-4.4\pm 4.2\,\mathrm{MeV}$ and the width $25.6\pm 6.1\,\mathrm{MeV}$ are obtained.
However, Hanhart claimed that it is not possible to conclude whether the observed enhancement corresponds to a bound state or a virtual state, due to the lack of statistics~\cite{Hanhart05}.
Recently, the COSY-GEM experiment investigated the ${}^{27}\mathrm{Al}(p,\ {}^3\mathrm{He})$ reaction at the recoilless kinematics~\cite{COSYGEM09}. They observed back-to-back $\pi^-$-$p$ pairs, whose energy corresponds to $S_{11}(1535)$ at rest. The missing-mass spectrum, with requiring back-to-back $\pi^-$-$p$ pairs in coincidence, shows an enhancement below the $^{25}\mathrm{Mg}+\eta$ threshold. They obtained the binding energy ($-13.13\pm 1.64\,\mathrm{MeV})$ and the width ($10.22\pm 2.98\,\mathrm{MeV}$) of the ${}^{25}\mathrm{Mg}\otimes \eta$ system.
Both the experiments tagged a pair of a pion and a nucleon as the decay product of $\eta$-mesic nuclei. Since an $\eta$ meson and a nucleon strongly couples with a $S_{11}(1535)$ resonance ($N^*$), a main decay channel is considered to be $N^*\to N\pi$. The $N^*\to N\eta$ decay, which has a large branching ratio in free space, is suppressed because of the Pauli blocking, even if the state is above the $\eta$ emission threshold. Other decay modes, namely $N^*N\to NN$ and $NN\pi$, also contributes to the total decay width. The width of $NN$ decay is estimated to be negligibly small~\cite{Chiang91, Nagahiro03}. Experimentally, the $N^*\to N\pi$ is easiest to identify, as its back-to-back topology is a clean signal.
Historically, the first experiment to search for $\eta$-mesic nuclei was done at BNL by using the $(\pi^+,\ p)$ reaction on lithium, carbon, oxygen, and aluminum targets~\cite{Chrien88}, motivated by a theoretical calculation by Liu and Haider~\cite{Liu86}. A peak structure corresponding to $\eta$-mesic nuclei was not observed for any target.
We consider that it is worthwhile to reconsider this reaction for two reasons, and plan to carry out a new experiment at J-PARC~\cite{Itahashi_LoI}.
First, the proton spectrometer was set to $15^\circ$ against the beam in the BNL experiment, and the momentum transfer is as large as $200\,\mathrm{MeV}/c$ for the incident beam momentum $800\,\mathrm{MeV}/c$ (see Fig.~\ref{fig4}). A theoretical calculation by Nagahiro \textit{et al.} revealed that many subcomponents of different configurations of a nucleon-hole and an $\eta$ meson will contribute with substantial fractions, and that they will mask a possible peak structure of $\eta$-mesic nuclei~\cite{Nagahiro09}.
\begin{figure}[t]
\begin{center}
\psfig{file=nagahiro_fig4.eps,width=2.75in}
\caption{Momentum transfer for the $A(\pi,\ N)$ reaction (heavy mass limit) as a function of incident pion momentum ($p_\pi$) or kinetic energy ($T_\pi$), for different scattering angle ($\theta_p$) of $0^\circ$ and $15^\circ$. The solid (dashed) line corresponds to the $\eta$ energy at $\omega_\eta=m_\eta$ ($m_\eta-50\,\mathrm{MeV}$). This figure is taken from Ref.~\refcite{Nagahiro09}.}
\label{fig4}
\end{center}
\end{figure}
On the other hand, the recoilless condition can be satisfied by measuring a proton scattered at almost zero degree. As shown in Fig.~\ref{fig4}, the magic momentum, for the $\eta$ energy $\omega_\eta=m_\eta-50\,\mathrm{MeV}$ and $m_\eta$, where $m_\eta$ is the mass of an $\eta$ meson, is $777\,\mathrm{MeV}/c$ and $950\,\mathrm{MeV}/c$, respectively. In this condition only substitutional states are enhanced, since the angular momentum of the target nucleus is conserved after the reaction.
Secondly, the BNL experiment investigated only the inclusive reaction. A signal of $\eta$-mesic nuclei, if exists, may be overwhelmed by a large background irrelevant to $\eta$ production. Such a background will be reduced by detecting $N\pi$ pairs in back-to-back simultaneously.
Taking these points into consideration, we proposed to study the $(\pi^-,\ n)$ reactions on $^7\mathrm{Li}$ target at J-PARC~\cite{Itahashi_LoI} (see also Ref.~\refcite{Nagahiro09}). We have chosen $^7\mathrm{Li}$, in addition to $^{12}\mathrm{C}$ as a candidate of the target, because of less contribution of a proton in the $p_{3/2}$-shell. It is replaced by an $\eta$ meson with its orbital angular momentum 1 after the reaction ($p$-wave component), and it is foreseen that the structure from the $s$-wave component from a proton in the $s_{1/2}$-shell is much more interesting than that from the $p$-wave component.
Furthermore, the missing-mass spectrum will be affected by the in-medium property of the $N^*(1535)$ resonance. Nagahiro \textit{et al.} derived the $\eta$-nucleus optical potentials, based on two kinds of chiral models --- \textit{chiral doublet model} and \textit{chiral unitary model} --- and calculated the spectrum for each potential~\cite{Nagahiro09}.
In the chiral doublet model, the $N^*$ is regarded as the chiral partner of the nucleon, and the mass gap between $N^*$ and $N$ decreases in a finite density because of partial restoration of chiral symmetry. Thereby, a level crossing between the $N^*$-$N^{-1}$ and $\eta$ modes may take place in a nucleus, which results in a deeply-bound $\eta$ state and a bump structure in the unbound region.
On the other hand, the mass gap will not change largely in nuclear medium in case of the chiral unitary model. A shallow bound state, which is similar to the result in Ref.~\refcite{Garcia02}, will appear in the missing-mass spectrum.
Therefore, we expect to obtain information on the in-medium $N^*$ property in relation with the level crossing, via a missing-mass spectroscopy covering the unbound region as well as the bound region.
\section{Conclusion}
We discussed experimental ideas to utilize the $(\pi,\ N)$ reactions for two purposes: investigation of the properties of the $S_{11}(1535)$ resonance, and search for $\eta$-mesic nuclei. It seems that both studies will shed light on the $\eta$-nucleon and $\eta$-nucleus interaction.
\section*{Acknowledgments}
We acknowledge stimulating comments from Dr. M. D\"{o}ring. We would like to thank Dr. H.~Nagahiro, Dr. D.~Jido, and Prof. S.~Hirenzaki for valuable discussions on $\eta$-mesic nuclei.
|
\section{Introduction}
In this paper we study Darboux integrable
semi-discrete chains of the form
\begin{equation}\label{dhyp}
\frac{d}{dx}t(n+1, x)=f(x,t(n, x),t(n+1, x),\frac{d}{dx}t(n, x))\, .
\end{equation}
Here unknown function $t=t(n,x)$ depends on discrete and continuous variables $n$ and $x$ respectively;
function $f=f(x,t,t_1,t_x)$ is assumed to
be locally analytic, and $\frac{\partial f}{\partial t_x}$ is not
identically zero. The last two decades the discrete phenomena have become very popular due
to various important applications (for more details see \cite{Zabrodin}-\cite{GKP} and references therein).
Below we use a subindex to indicate the shift of the discrete
argument: $t_k=t(n+k,x)$, $k\in \ZZ$, and derivatives with respect
to $x$: $t_{[1]}=t_x=\displaystyle{\frac{d}{d x}}t(n,x),$
$t_{[2]}=t_{xx}=\displaystyle{\frac{d^2}{d x^2}}t(n,x)$,
$t_{[m]}=\frac{d^m}{dx^m}t(n,x)$, $m\in \NN$. Introduce {\it the set of
dynamical variables} containing $\{t_k\}_{k=-\infty}^{\infty};$
$\{t_{[m]}\}_{m=1}^{\infty}$.
We denote through $D$ and $D_x$ the shift operator and the
operator of the total derivative with respect to $x$ correspondingly. For
instance, $Dh(n,x)=h(n+1,x)$ and
$D_xh(n,x)=\frac{d}{d x}h(n,x)$.
Functions $I$ and $F$, both depending on $x$, $n$, and a finite number of
dynamical variables, are called respectively {\it $n$- and $x$-integrals}
of (\ref{dhyp}), if $DI=I$ and $D_xF=0$ (see also
\cite{AdlerStartsev}).
Clearly, any function depending on $n$ only, is an $x$-integral, and any
function, depending on $x$ only, is an $n$-integral. Such integrals are called {\it trivial}
integrals. One can see that any $n$-integral $I$ does
not depend on variables $t_{m}$, $m\in \ZZ\backslash\{0\}$,
and any $x$-integral $F$
does not depend on variables $t_{[m]}$, $m\in \NN$.
Chain (\ref{dhyp}) is called {\it Darboux integrable}
if it admits a nontrivial $n$-integral and a nontrivial
$x$-integral.
The basic ideas on integration of partial differential equations of the hyperbolic type
go back to classical works by Laplace, Darboux, Goursat, Vessiot, Monge, Ampere, Legendre,
Egorov, etc. Notice that understanding of integration as finding an explicit formula for a
general solution was later replaced by other, in a sense less obligatory, definitions.
For instance, the Darboux method for integration of hyperbolic type equations consists of
searching for integrals in both directions followed by the reduction of the equation to
two ordinary differential equations. In order to find integrals, provided that they exist,
Darboux used the Laplace cascade method.
An alternative, more algebraic approach based on the characteristic vector fields was used by
Goursat and Vessiot. Namely this method allowed Goursat to get a list of integrable equations
\cite{GOURSAT}. An important contribution to the development of the algebraic method
investigating Darboux integrable equations was made by A.B.~Shabat who introduced the
notion of the characteristic algebra of the hyperbolic equation
\begin{equation}\label{hyp}
u_{x,y}=f(x,y,u,u_x,u_y)\, .
\end{equation}
It turned out that the operator $D_y$ of total differentiation, with
respect to the variable $y$, defines a derivative in the
characteristic algebra in the direction of $x$. Moreover, the
operator $ad_{D_y}$ defined according to the rule
$ad_{D_y}X=[D_y,X]$ acts on the generators of the algebra in a very
simple way. This makes it possible to obtain effective integrability
conditions for the equation (\ref{hyp}) (see \cite{ShabatYamilov}).
A.V. Zhiber and F.Kh. Mukminov investigated the structure of the characteristic
algebras for the so-called quadratic systems containing the Liouville equation
and the sine-Gordon equation (see \cite{ZhiberMukminov}). In \cite{ZhiberMukminov} and
\cite{ZhiberMurtazina} the very nontrivial connection between characteristic algebras
and Lax pairs of the hyperbolic S-integrable equations and systems of equations is studied,
and perspectives on the application of the characteristic algebras to classify such kinds of
equations are discussed.
Recently the concept of the characteristic algebras has been
defined for discrete models. In our articles \cite{hp}-\cite{HZhP2}
an effective algorithm was worked out to classify Darboux integrable
models. By using this algorithm some new classification results were
obtained. In
\cite{h2010} a method of classification of S-integrable discrete
models is suggested based on the concept of characteristic algebra.
The article is organized as follows. In Section 2 characteristic algebras are defined for the chain (\ref{dhyp}). In Section 3 we describe the structure of $n$-integrals and $x$-integrals of the Darboux integrable chains of general
form (\ref{dhyp}) (see Theorems \ref{Thm1} and \ref{Thm2}). Then we show that one can
choose the minimal order $n$-integral and the minimal order $x$-integral of a special
canonical form, important for the purpose of classification (see Theorems
\ref{Thm3} and \ref{Thm4}).
The complete classification of a particular case
$t_{1x}=t_x+d(t_1,t)$ in \cite{HZhP2} was done due to the finiteness
of the characteristic algebras in both directions. However,
algebras themselves were not described. In Subsections 4.1 and 4.2
we fill up this gap and represent the tables of multiplications
for all of these algebras.
The problem of finding explicit solutions for Darboux integrable
models is rather complicated. Even in the mostly studied case of PDE
$u_{xy}=f(x,y,u,u_x,u_y)$ this problem is not completely solved. In
Subsection 4.3 we give the explicit formulas for general solutions
of the integrable chains in the particular case
$t_{1x}=t_x+d(t,t_1)$ (see Theorem \ref{TheoremExplicit}).
It is remarkable that the classification of Darboux integrable chains is closely connected with the classical
problem of description of ODE admitting a fundamental system of solutions (following Vessiot-Guldberg-Lie),
for the details one can see \cite{ibr} and the references therein.
Indeed any $n$-integral defines an ordinary differential equation $F(n,x,t,t_x,...t_{[k]})=p(x)$ for which the corresponding $x$-integral gives a formula $I(n,x,t,t_1,...t_m)=c_n$ allowing one to find a new solution $t_m$ for given set of solutions $t,t_1,...t_{m-1}.$ Iterating this way one finds a solution $t_N=H(t,t_1,...t_{m-1},x,c_1,c_2,...c_k)$ depending on $k$ arbitrary constants. In the case when $I$ does not depend on $x$ explicitly this formula gives general solution in a desired form. Examples are given in the Remark in Section 4.
\section{Characteristic algebras of discrete models}
Let us introduce the characteristic algebras for chain (\ref{dhyp}).
Due to the
requirement of $\frac{\partial f }{\partial t_x} \neq 0$, we can
rewrite (at least locally) chain (\ref{dhyp}) in the inverse form
$$t_x(n-1, x)=g(x,t(n, x),t(n-1, x),t_x(n, x)).$$
Since $x$-integral $F$ does not depend on the variables $t_{[k]}$, $k\in \NN$, then
the equation $D_xF=0$ becomes $KF=0$, where
\begin{equation}\label{gc1} K = \frac{\partial }{\partial x}+t_x\frac{\partial
}{\partial t} +f\frac{\partial }{\partial t_1 }+g\frac{\partial
}{\partial t_{-1}} +f_1\frac{\partial }{\partial
t_2}+g_{-1}\frac{\partial }{\partial t_{-2} }+\ldots\, .
\end{equation}
Also, $XF=0$,
with $X=\frac{\partial}{\partial t_x}$. Consider the linear space over the field of locally analytic functions depending on a finite number of dynamical variables spanned by all multiple commutators of $K$ and $X$. This set is closed with respect to three operations: addition, multiplication by a function, and taking the commutator of two elements. It is called characteristic algebra $L_x$ of chain (\ref{dhyp}) in the $x$-direction. Therefore, any vector field from algebra $L_x$ annulates $F$.
The relation between Darboux integrability of chain (\ref{dhyp}) and its
characteristic algebra $L_x$ is given by the following important criterion.
\begin{thm}\label{thm1} (see \cite{HZhP2})
Chain (\ref{dhyp}) admits a
nontrivial $x$-integral if and only if its characteristic algebra $L_x$ is of
finite dimension.
\end{thm}
The
equation $DI=I$, defining an $n$-integral $I$, in an enlarged form becomes
\begin{equation}\label{n-integral}
I(x,n+1, t_1,f,f_{x},...)=I(x,n,t,t_x,t_{xx},...).
\end{equation}
The left hand side contains the variable $t_1$ while the right
hand side does not. Hence we have $D^{-1}\frac{d}{dt_1}DI=0,$ i.e.
the $n$-integral is in the kernel of the operator
$$
Y_1=D^{-1}Y_0D, $$ where
\begin{equation}\label{Y1}
Y_1=\frac{\partial}{\partial t}+D^{-1}(Y_0f)\frac{\partial}{\partial t_x}+
D^{-1}Y_0(f_x)\frac{\partial}{\partial t_{xx}}+ D^{-1}Y_0(f_{xx})
\frac{\partial}{\partial t_{xxx}}+...,
\end{equation}
and \begin{equation} Y_0=\frac{d}{dt_1}\,.
\end{equation}
One can show that $D^{-j}Y_0D^jI=0$
for any natural $j$.
Direct calculations show that
$$
D^{-j}Y_0D^j=X_{j-1}+Y_j, \qquad j\geq 2,
$$
where
\begin{eqnarray}
&&Y_{j+1}=D^{-1}(Y_jf)\frac{\partial}{\partial t_x}
+D^{-1}Y_j(f_x)\frac{\partial}{\partial t_{xx}}+
D^{-1}Y_j(f_{xx})\frac{\partial}{\partial t_{xxx}}+...,\quad j\geq
1\, , \label{Yj}
\end{eqnarray}
\begin{equation}\label{definitionXj}
X_j=\frac{\partial}{\partial_{t_{-j}}}, \qquad j\geq
1.\end{equation}
Define by $N^*$ the dimension of the linear space spanned by the operators
$\{Y_j \}_1^{\infty}$. Introduce the linear space over the field of locally analytic functions depending on a finite number of dynamical variables spanned by all multiple commutators of the vector fields from $\{Y_j \}_1^{N^*}\cup\{X_j \}_1^{N^*}$. This linear space is closed with respect to three operations: addition, multiplication by a function, and taking the commutator of two elements.
It is called characteristic algebra $L_n$ of chain (\ref{dhyp}) in the $n$-direction.
\begin{thm}\label{thm2}(see \cite{hp})
Equation (\ref{dhyp}) admits a nontrivial $n$-integral if and only
if its characteristic algebra $L_n$ is of
finite dimension.
\end{thm}
\section{On the structure of nontrivial $x$- and $n$- integrals}
We define the {\it order of a nontrivial $n$-integral} $I=I(x,n,t,t_x,
\ldots, t_{[k]})$ with $\frac{\partial I}{\partial t_{[k]}}\ne 0$, as the number $k$.
\begin{thm} \label{Thm1}
Assume equation (\ref{dhyp}) admits a nontrivial $n$-integral. Then for any nontrivial $n$-integral
$I^*(x,n,t,t_x, \ldots, t_{[k]})$ of the smallest order and any $n$-integral $I$ we have
\begin{equation}\label{representationnint}I=\phi(x, I^*, D_xI^*, D^2_xI^*, \ldots),
\end{equation}
where $\phi$ is some function.
\end{thm}
{\it Proof}: Denote by $I^*=I^*(x,n,t, \ldots, t_{[k]})$ an $n$-integral of the smallest order.
Let $I$ be any other $n$-integral, $I=I(x,n, t, \ldots, t_{[r]})$.
Clearly $r\geq k$. Let us introduce new variables $x$, $n$, $t$, $t_x$, $\ldots$, $t_{[k-1]}$, $I^*$,
$D_xI^*$, $\ldots$, $D_x^{r-k}I^*$ instead of the variables $x$, $n$, $t$, $t_x$, $\ldots$, $t_{[k-1]}$,
$t_{[k]}$, $t_{[k+1]}$, $\ldots$, $t_{[r]}$. Now, $I=I(x, n, t, t_x, \ldots, t_{[k-1]}, I^*,
D_xI^*, \ldots, D_x^{r-k}I^*)$. We write the power series for function $I$ in the neighborhood of the point
$( (I^{*})_0,
(D_xI^*)_0, \ldots, (D_x^{r-k}I^*)_0)$:
\begin{equation}\label{nintseries}
I=\sum\limits_{i_0, i_1, \ldots, i_{r-k}} E_{i_0, i_1, \ldots,
i_{r-k}}(I^*-(I^{*})_0)^{i_0}(D_xI^*-(D_xI^*)_0)^{i_1}
\ldots (D_x^{r-k}I^*-(D_x^{r-k}I^*)_0)^{i_{r-k}}\, .
\end{equation}
Then
$$
DI=\sum\limits_{i_0, i_1, \ldots, i_{r-k}} DE_{i_0, i_1, \ldots,
i_{r-k}}(DI^*-(I^{*})_0)^{i_0}(DD_xI^*-(D_xI^*)_0)^{i_1}
\ldots (DD_x^{r-k}I^*-(D_x^{r-k}I^*)_0)^{i_{r-k}}\, .
$$
Since $DI=I$, $DD^j_xI^*=D^j_xDI^*=D^j_xI^*$ and the power series representation
for function $I$ is unique, then
$DE_{i_0, i_1, \ldots, i_{r-k}}=E_{i_0, i_1, \ldots, i_{r-k}}$, i.e. $E_{i_0, i_1,
\ldots, i_{r-k}}(x,n, t, \ldots, t_{[k-1]})$ are all $n$-integrals. Due to the fact that
minimal $n$-integral depends on $x$, $n$, $t$, $\ldots$, $t_{[k]}$,
we conclude that all $E_{i_0, i_1, \ldots, i_{r-k}}(x, n, t, \ldots, t_{[k-1]})$ are trivial
$n$-integrals, i.e. functions depending only on $x$. Now equation (\ref{representationnint})
follows immediately from (\ref{nintseries}).
We define the {\it order of a nontrivial $x$-integral} $F=F(x,n,t_{k},t_{k+1},...t_{m}) $
with $\frac{\partial F}{\partial t_{[m]}}\ne 0$, as the number $m-k$.
\begin{thm} \label{Thm2}
Assume equation (\ref{dhyp}) admits a nontrivial $x$-integral. Then for any nontrivial
$x$-integral $F^*(x,n, t, t_1, \ldots, t_m)$ of the smallest order and any $x$-integral $F$ we have
\begin{equation}\label{representationxint}
F=\xi(n,F^*, DF^*, D^2F^*, \ldots),
\end{equation}
where $\xi$ is some function.
\end{thm}
{\it Proof}: Denote by $F^*=F^*(x,n, t, t_1, \ldots, t_m)$ an
$x$-integral of the smallest order. Let $F$ be any other
$x$-integral, $F=F(x, n,t, t_1, \ldots, t_l)$. Clearly, $l\geq m$. Let
us introduce new variables $x$, $n$, $t$, $t_1$, $\ldots$, $t_{m-1}$,
$F^*$, $DF^*$, $\ldots$, $D^{l-m}F^*$ instead of variables $x$, $n$, $t$,
$t_1$, $\ldots$,$t_{m-1}$, $t_m$, $\ldots$, $t_l$. Now, $F=F(x, n, t, t_1,
\ldots,t_{m-1}, F^*, DF^*, \ldots, D^{l-m}F^*)$. We write the power
series representation of function $F$ in the neighborhood of point
$((F^*)_0, (DF^*)_0, \ldots, (D^{l-m}F^*)_0)$:
\begin{equation}\label{xintseries}
F=\sum\limits_{i_0,i_1, \ldots, i_{l-m}}K_{i_0,i_1, \ldots, i_{l-m}}
(F^*-(F^*)_0)^{i_0}(DF^*-(DF^*)_0)^{i_1}\ldots
(D^{l-m}F^*-(D^{l-m}F^*)_0)^{i_{l-m}}\, .
\end{equation}
Then
$$
D_xF=\sum\limits_{i_0,i_1, \ldots, i_{l-m}}D_x\{K_{i_0,i_1, \ldots, i_{l-m}}\}
(F^*-(F^*)_0)^{i_0}(DF^*-(DF^*)_0)^{i_1}\ldots
(D^{l-m}F^*-(D^{l-m}F^*)_0)^{i_{l-m}}
$$
$$
+\sum\limits_{i_0,i_1, \ldots, i_{l-m}}K_{i_0,i_1, \ldots, i_{l-m}}
D_x\{(F^*-(F^*)_0)^{i_0}(DF^*-(DF^*)_0)^{i_1}\ldots
(D^{l-m}F^*-(D^{l-m}F^*)_0)^{i_{l-m}}\}
$$
Since $D_xD^jF^*=D^jD_xF^*=0$, then
$$D_x\{(F^*-(F^*)_0)^{i_0}(DF^*-(DF^*)_0)^{i_1}\ldots
(D^{l-m}F^*-(D^{l-m}F^*)_0)^{i_{l-m}}\}=0.$$ Therefore,
$$
0=D_xF=\sum\limits_{i_0,i_1, \ldots, i_{l-m}}D_x\{K_{i_0,i_1, \ldots, i_{l-m}}\}
(F^*-(F^*)_0)^{i_0}(DF^*-(DF^*)_0)^{i_1}\ldots
(D^{l-m}F^*-(D^{l-m}F^*)_0)^{i_{l-m}}\, .
$$
Due to the unique representation of the zero power series we have that
$D_x\{K_{i_0,i_1, \ldots, i_{l-m}}\}=0$, i.e. all $K_{i_0,i_1, \ldots, i_{l-m}}
(x, n, t, \ldots, t_{m-1})$ are
$x$-integrals. Since the minimal nontrivial $x$-integral is of order $m$,
then all $K_{i_0,i_1, \ldots, i_{l-m}}$ are
trivial $x$-integrals, i.e. functions depending on $n$ only.
Now the equation (\ref{representationxint}) follows from (\ref{xintseries}).
The next two theorems are discrete versions of Lemma 1.2 from \cite{Zhiber}.
\begin{thm} \label{Thm3}
Among all nontrivial $n$-integrals $I^*(x, n, t, t_x, \ldots, t_{[k]})$ of the
smallest order, with $k\geq 2$, there is an $n$-integral $I^0(x, n, t,t_x,
\ldots , t_{[k]})$ such that
\begin{equation}
\label{linearnint}
I^0(x, n, t, t_x, \ldots, t_{[k]})=a(x, n,t, t_x, \ldots, t_{[k-1]})t_{[k]}+
b(x, n, t, t_x, \ldots, t_{[k-1]})\, .
\end{equation}
\end{thm}
{\it Proof}: Consider nontrivial minimal $n$-integral $I^*(x,n, t, t_x,
\ldots, t_{[k]})$ with $k\geq 2$. Equality $DI^*=I^*$
can be rewritten as
$$
I^*(x, n+1, t_1, f, f_x, \ldots, f_{[k-1]})=I^*(x,n, t,t_x, \ldots, t_{[k]}).
$$
We differentiate both sides of the last equality with respect to $t_{[k]}$:
\begin{equation}\label{**}
\frac{\partial I^*(x,n+1, t_1, f, \ldots, f_{[k-1]})}{\partial f_{[k-1]}}
\cdot\frac{\partial f_{[k-1]}}{\partial t_{[k]}}=
\frac{\partial I^*(x, n, t, \ldots, t_{[k]})}{\partial t_{[k]}}\, .
\end{equation}
In virtue of $\frac{\partial f_{[j]}}{\partial t_{[j+1]}}=f_{t_x}$, the equation
(\ref{**}) can be rewritten as
\begin{equation}\label{***}
\frac{\partial I^*(x, n+1, t_1, f, \ldots, f_{[k-1]})}{\partial f_{[k-1]}}f_{t_x}=
\frac{\partial I^*(x, n, t, \ldots, t_{[k]})}{\partial t_{[k]}}\, .
\end{equation}
Let us differentiate once more with respect to $t_{[k]}$ both sides of the last equation, we have:
$$
\frac{\partial^2 I^*(x,n+1, t_1, f, \ldots, f_{[k-1]})}{\partial^2 f_{[k-1]}}f^2_{t_x}=
\frac{\partial^2 I^*(x, n, t, \ldots, t_{[k]})}{\partial t^2_{[k]}}\, ,
$$
or the same,
$$D\left\{ \frac{\partial^2I^*}{\partial t^2_{[k]}}\right\}f^2_{t_x}=
\frac{\partial^2I^*}{\partial t^2_{[k]}}\, ,
$$
where $I^*=I^*(x,n, t,\ldots, t_{[k]})$. It follows from (\ref{***}) that
$$
D\left\{\frac{\partial^2I^*}{\partial t^2_{[k]}}\right\}\left\{
\frac{\partial I^*}{\partial t_{[k]}}\right\}^2=
\frac{\partial^2I^*}{\partial t^2_{[k]}}D\left\{\left(
\frac{\partial I^*}{\partial t_{[k]}}\right)^2\right\}\, ,
$$
or the same, function
$$
J:=\frac{\frac{\partial^2I^*}{\partial t^2_{[k]}}}{\left(
\frac{\partial I^*}{\partial t_{[k]}}\right)^2}
$$ is an $n$-integral, and by Theorem \ref{Thm1}, we have that $J=\phi(x, I^*)$. Therefore,
$$
\frac{\partial^2I^*}{\partial t^2_{[k]}}=\frac{\partial H(x, I^*)}{\partial I^*}\left(
\frac{\partial I^*}{\partial t_{[k]}}\right)^2, \quad \mbox{where} \quad \frac{\partial H}{\partial I^*}=J,
$$
or
$$
\frac{\partial }{\partial t_{[k]}} \left\{ \ln \frac{\partial I^*}{\partial t_{[k]}}-H(x, I^*)\right\}=0.
$$
Hence, $e^{-H(x, I^*)}
\frac{\partial I^*}{\partial t_{[k]}}=e^g$ for some function $g(x,n, t,t_x, \ldots, t_{[k-1]})$.
Introduce $W$ in such a way that $\frac{\partial W}{\partial I^*}
=e^{-H(x, I^*)}$. Then $\frac{\partial W}{\partial t_{[k]}}=e^g$ and
$W=e^{g(x,n, t, \ldots, t_{[k-1]})}t_{[k]}+l(x,n, t, \ldots, t_{[k-1]})$ is an $n$-integral,
where $l(x,n, t, \ldots, t_{[k-1]})$ is some function.
\begin{thm} \label{Thm4}
Among all nontrivial $x$-integrals $F^*(x, n, t_{-1}, t, t_1, \ldots, t_m)$ of the smallest order,
with $m\geq 1$, there is $x$-integral $F^0(x, n, t_{-1}, t, t_{1}, \ldots, t_m)$ such that
\begin{equation}\label{separatexint}
F^0(x,n, t_{-1}, t, t_{1}, \ldots, t_m)=A(x,n, t_{-1}, t, \ldots, t_{m-1})+B(x,n, t, t_1, \ldots, t_m).
\end{equation}
\end{thm}
{\it Proof}: Consider nontrivial $x$-integral $F^*(x, n, t_{-1}, t, t_1, \ldots, t_m)$ of minimal order.
Since $D_xF^*=0$, then
\begin{equation}\label{O}
\frac{\partial F^*}{\partial x}+g\frac{\partial F^*}{\partial t_{-1}}+
t_x\frac{\partial F^*}{\partial t}+f\frac{\partial F^*}{\partial t_1}+Df
\frac{\partial F^*}{\partial t_2}+\ldots+D^{m-1}f\frac{\partial F^*}{\partial t_m}=0.
\end{equation}
We differentiate both sides of (\ref{O}) with respect to $t_m$ and with respect to
$t_{-1}$ separately and have the following
two equations:
\begin{equation}\label{A}
\{D_x+\frac{\partial }{\partial t_m}(D^{m-1}f)\}\frac{\partial F^*}{\partial t_m}=0,
\end{equation}
\begin{equation}\label{B}
\{D_x+\frac{\partial g }{\partial t_{-1}}\}\frac{\partial F^*}{\partial t_{-1}}=0.
\end{equation}
Let us differentiate (\ref{A}) with respect to $t_{-1}$, we have,
\begin{equation}\label{C}
D_x\frac{\partial^2F^*}{\partial t_m\partial t_{-1}}+\frac{\partial g}{\partial t_{-1}}
\frac{\partial^2F^*}{\partial t_m\partial t_{-1}}+\frac{\partial }{\partial t_m}(D^{m-1}f)
\frac{\partial^2F^*}{\partial t_m\partial t_{-1}}=0\, .
\end{equation}
It follows from (\ref{A}) and (\ref{B}) that $\frac{\partial }{\partial t_m}(D^{m-1}f)=-
\frac{D_xF^*_{t_m}}{F^*_{t_m}}$,
$\frac{\partial g}{\partial t_{-1}}=-\frac{D_x F^*_{t_{-1}}}{F^*_{t_{-1}}}$. Equation (\ref{C}) becomes
$$
D_x\left\{ \ln \frac{F^*_{t_mt_{-1}}}{F^*_{t_m}F^*_{t_{-1}}}\right\}=0.
$$
By Theorem \ref{Thm2} we have, $\frac{F^*_{t_mt_{-1}}}{F^*_{t_m}F^*_{t_{-1}}}=\xi(n, F^*)$, or
$$
\frac{F^*_{t_mt_{-1}}}{F^*_{t_m}}=F^*_{t_{-1}}\xi(n, F^*)=H'(F^*)F^*_{t_{-1}}=
\frac{\partial}{\partial t_{-1}}H(F^*), \quad {\mbox{where }} \quad
\xi(n,F^* )=H'(n, F^*)\, .
$$
Thus,
$\frac{\partial}{\partial t_{-1}}\{ \ln F^*_{t_m} -H(n,F^*)\}=0$, or $e^{-H(n, F^*)}F^*_{t_m}=
C(x,n, t, t_1, \ldots, t_m)$ for some function \\ $C(x,n, t, t_1, \ldots, t_m)$.
Denote by $\tilde{H^*}(n, F)$ such a function that $\tilde{H}'(n, F^*) =e^{-H(n, F^*)}$. Then
$\frac{\partial \tilde{H}(n, F^*)}{\partial t_m}=C(x,n, t,t_1, \ldots, t_m)$. Hence,
$\tilde{H}(n, F^*)=B(x,n, t, t_1, \ldots, t_m)+A(x,n, t_{-1}, t, \ldots, t_{m-1})$.
Since $D_x\tilde{H}(F^*)=\tilde{H}'(n, F^*)D_x(F^*)=0$, then
$\tilde{H}(n, F^*)$ is an $x$-integral in the desired form (\ref{separatexint}).
\section{Particular case: $t_{1x}=t_x+d(t, t_1)$}
Finiteness of the characteristic algebras $L_x$ and $L_n$ was
used in \cite{HZhP1} and \cite{HZhP2} to classify Darboux integrable
semi-discrete chains of special form
\begin{equation}
\label{main}
t_{1x}=t_x+d(t, t_1)\, .
\end{equation}
The statement of this classification result is given by the next theorem from \cite{HZhP2}.
\begin{thm}\label{maintheorem}
Chain (\ref{main}) admits nontrivial $x$- and $n$-integrals if and
only if it is one of the kind:\\
(a) \begin{equation}\label{a}t_{1x}=t_x+A(t_1-t),
\end{equation} where $A(t_1-t)$ is given in
implicit form $A(t_1-t)=\frac{d}{d\theta}P(\theta)$,
$t_1-t=P(\theta)$, with $P(\theta)$ being an arbitrary quasi polynomial, i.e. a
function satisfying an ordinary differential equation
P^{(N+1)}=\mu_NP^{(N)}+\ldots+\mu_1P'+\mu_0P
$with constant coefficients $\mu_k$, $0\leq k\leq N$,\\
(b)\begin{equation}\label{b}t_{1x}=t_x+C_1(t_1^2-t^2)+C_2(t_1-t),\end{equation}
(c) \begin{equation}\label{c}t_{1x}=t_x+\sqrt{C_3e^{2\alpha
t_1}+C_4e^{\alpha(t_1+t)}+C_3e^{2\alpha t}},
\end{equation}
(d) \begin{equation}\label{d}t_{1x}=t_x +C_5(e^{\alpha
t_1}-e^{\alpha t})+C_6(e^{-\alpha t_1}-e^{-\alpha t}),\end{equation}
\noindent where $\alpha\ne 0$, $C_i$, $1\leq i\leq 6$, are
arbitrary constants. Moreover, some nontrivial $x$-integrals $F$ and
$n$-integrals $I$ in each of the cases are
\begin{enumerate}
\item[i)] $F=x-\int^{t_1-t}\frac{ds}{A(s)}$, $I=L(D_x)t_x$, where $L(D_x)$ is a differential operator
which annihilates $\frac{d}{d\theta}P(\theta)$ where $D_x\theta=1$.
\item[ii)]
$F=\frac{(t_3-t_1)(t_2-t)}{(t_3-t_2)(t_1-t)}$, $I=t_x-C_1t^2-C_2t,$
\item[iii)]
$F=arcsinh(ae^{\alpha(t_1-t_2)}+b)+arcsinh(ae^{\alpha(t_1-t)}+b),
\quad a=2C_3/\sqrt{4c_3^2-c_4^2}, \quad b=C_4/\sqrt{4c_3^2-c_4^2}, $
$I=2t_{xx}-\alpha t_x^2-\alpha C_3
\mathrm{e}^{2\alpha t}$,
\item[iv)]
$F=\frac{(\mathrm{e}^{\alpha t}-\mathrm{e}^{\alpha
t_2})(\mathrm{e}^{\alpha t_1}-\mathrm{e}^{\alpha
t_3})}{(\mathrm{e}^{\alpha t}-\mathrm{e}^{ \alpha
t_3})(\mathrm{e}^{\alpha t_1}-\mathrm{e}^{\alpha t_2})}$, $I=t_x-C_5
e^{\alpha t}-C_6 e^{-\alpha t}$.
\end{enumerate}
\end{thm}
Note that all the integrals in Theorem \ref{maintheorem} are given
in their canonical forms (see Theorems \ref{Thm3} and \ref{Thm4}).
\noindent{\bf{Remark}}. In case (c) equation (\ref{c}) is closely connected with
the well-known Steen-Ermakov equation (see \cite{rogers} and the
references therein)
\begin{equation}\label{ermakov}
y''+q(x)y=c_3y^{-3}.
\end{equation}
Indeed for $\alpha=2$ its $n$-integral $2t_{xx}-\alpha t_x^2-\alpha
C_3\mathrm{e}^{2\alpha t}=p(x)$ is reduced to the form
(\ref{ermakov}) by substitution $y=\mathrm{e}^{-t}$. Now it follows
from the $x$-integral $F$ that for arbitrary three solutions
$y(x)$, $z(x)$, $w(x)$ of the Steen-Ermakov equation the following
function
$$ R(y,z,w)=arcsinh({aw^2}{y^{-2}}+b)+arcsinh({az^2}{y^{-2}}+b)$$
does not depend on $x$. Remind that the Riccati equation connected with the cases (b) and (d) has a
similar property: the cross-ratio of its four solutions is a constant.
\subsection{Characteristic algebras $L_x$ for Darboux integrable equations
$t_{1x}=t_x+d(t,t_1)$}
It was proved (see \cite{HZhP1}) that if equation $t_{1x}=t_x+d(t,t_1)$ admits a
nontrivial $x$-integral, then it admits a nontrivial $x$-integral not depending on $x$.
Introduce new vector fields
$$
\tilde{X}=[X, K]=\sum\limits_{k=-\infty, }^\infty \frac{\partial }{\partial t_k}\, \qquad
J:=[\tilde{X}, K]\, .
$$
\subsubsection {Case 1: $t_{1x}=t_x+A(t_1-t)$}
Direct calculations show that the multiplication table for characteristic algebra $L_x$ is the following
\begin{center}
\begin{tabular}{|c||c|c|c|}
\hline
$L_x$ & $X$ & $K$ & $\tilde{X}$ \\
\hline
\hline
$X$ & 0 & $\tilde{X}$ & 0 \\
\hline
$K$ & $-\tilde{X}$ & 0 & 0 \\
\hline
$\tilde{X}$ & 0 & 0 & 0 \\
\hline
\end{tabular}
\end{center}
\subsubsection{Case 2: $t_{1x}=t_x+C_1(t_1^2-t^2)+C_2(t_1-t)$}
Direct calculations show that
$$J=2C_1\sum\limits_{k=-\infty, k\ne 0}^\infty (t_k-t) \frac{\partial }{\partial t_k}$$
and
$$
[J, K]=2C_1^2\sum\limits_{k=-\infty, k\ne 0}^\infty (t_k-t)^2 \frac{\partial }{\partial t_k}
=2C_1(K-t_x \tilde{X})-(2C_1t+C_2)J\, \,$$
and the multiplication table for characteristic algebra $L_x$ is the following
\begin{center}
\begin{tabular}{|c||c|c|c|c|}
\hline
$L_x$ & $X$ & $K$ & $\tilde{X}$ & $J$\\
\hline
\hline
$X$ & 0 & $\tilde{X}$ & 0 & 0\\
\hline
$K$ & $-\tilde{X}$ & 0 & $-J$ & $-2C_1(K-t_x \tilde{X})+(2C_1t+C_2)J$\\
\hline
$\tilde{X}$ & 0 & $J$ & 0 & 0\\
\hline
$J$ & 0 & $2C_1(K-t_x \tilde{X})-(2C_1t+C_2)J$ & 0& 0\\
\hline
\end{tabular}
\end{center}
\subsubsection{Case 3: $t_{1x}=t_x +\sqrt{C_3e^{2\alpha t_1}+C_4e^{\alpha (t_1+t)}+C_3e^{2\alpha t}}$}
Direct calculations show that $[\tilde{X}, K]=\alpha K-\alpha t_x \tilde{X}$, and the multiplication
table for characteristic algebra $L_x$ is the following
\begin{center}
\begin{tabular}{|c||c|c|c|}
\hline
$L_x$ & $X$ & $K$ & $\tilde{X}$ \\
\hline
\hline
$X$ & 0 & $\tilde{X}$ & 0 \\
\hline
$K$ & $-\tilde{X}$ & 0 & $-\alpha K+\alpha t_x \tilde{X}$ \\
\hline
$\tilde{X}$ & 0 & $\alpha K-\alpha t_x \tilde{X}$ & 0 \\
\hline
\end{tabular}
\end{center}
\subsubsection{Case 4: $t_{1x}=t_x+C_5(e^{\alpha t_1}-e^{\alpha t})
+C_6(e^{-\alpha t_1}-e^{-\alpha t})$}
Direct calculations show that
$$
J=\alpha \sum\limits_{k=-\infty, k\ne 0}^\infty \{ C_5(e^{\alpha t_k}-e^{\alpha t})-
C_6(e^{-\alpha t_k}-e^{-\alpha t})\}
\frac{\partial }{\partial t_k}
$$
and
$$
[J, K]=2C_5C_6\alpha^2 \sum\limits_{k=-\infty, k\ne 0}^\infty \{ e^{\alpha (t-t_k)}+
e^{\alpha (t_k-t)}-2\}
\frac{\partial }{\partial t_k}
$$
$$
= \alpha^2(C_5e^{\alpha t}+C_6e^{-\alpha t})(K-t_x \tilde{X})+\alpha (C_6e^{-\alpha t}
-C_5e^{\alpha t})J.
$$
Denote by
$$
\beta_1=\alpha^2(C_5e^{\alpha t}+C_6e^{-\alpha t}), \qquad \beta_2=\alpha (C_6e^{-\alpha t}
-C_5e^{\alpha t})\, .
$$
The multiplication table for characteristic algebra $L_x$ is
\begin{center}
\begin{tabular}{|c||c|c|c|c|}
\hline
$L_x$ & $X$ & $K$ & $\tilde{X}$ & $J$\\
\hline
\hline
$X$ & 0 & $\tilde{X}$ & 0 & 0\\
\hline
$K$ & $-\tilde{X}$ & 0 & $-J$ &$-\beta_1(K-t_x \tilde{X})-\beta_2 J $\\
\hline
$\tilde{X}$ & 0 & $J$ & 0 &$\alpha^2K-\alpha^2\tilde{X}$\\
\hline
$J$ & 0 & $\beta_1(K-t_x \tilde{X})+\beta_2 J$ & $\alpha^2\tilde{X}-\alpha^2K$& 0\\
\hline
\end{tabular}
\end{center}
\subsection{Characteristic algebras $L_n$ for Darboux integrable equation
$t_{1x}=t_x+d(t,t_1)$}
\subsubsection {Case 1: $t_{1x}=t_x+A(t_1-t)$}
Characteristic algebra $L_n$ is generated only by two vector fields $X_1$ and $Y_1$, and can be of
any finite dimension.
If $A(t_1-t)=t_1-t+c$, where $c$ is some constant, then characteristic algebra $L_n$ is trivial,
consisting of $X_1$ and $Y_1$ only, with commutativity relation $[X_1, Y_1]=0$. If $A(t_1-t)\ne t_1-t+c$,
one can choose a basis in $L_n$
consisting of $W=\frac{\partial }{\partial \theta}$, $Z=\sum_{k=0}^{k=\infty}
D_x^kp(\theta)\partial/\partial t_{[k]}$, with $\theta=x+\alpha_n$,
$C_1=[W, Z]$, $C_{k+1}=[W, C_k]$, $1\leq k\leq N-1$. Its multiplication table for $L_n$ is the following
\begin{center}
\begin{tabular}{|c||c|c|c|c|c|c|c|c|c|}
\hline
$L_n$ & $W$ & $Z$ & $C_1$ & $C_2$ & $\ldots$& $C_k$ & $\ldots$ & $C_{N-1}$ & $C_N$ \\
\hline
\hline
$W$ & 0 & $C_1$ & $C_2$ & $C_3$ &$\ldots$ & $C_{k+1}$ & $\ldots$ & $C_N$ &$K$ \\
\hline
$Z$ & $-C_1$ & 0 & 0 & 0 & $\ldots$ & 0 & $\ldots$ &0 &0 \\
\hline
$C_1$ & $-C_2$ & 0 & 0 & 0 &$\ldots$ & 0 & $\ldots$ &0&0 \\
\hline
$\vdots$ & $\vdots$ & $\vdots$ & $\vdots$ & $\vdots$ & $\vdots$ & $\vdots$ & $\vdots$ &
$\vdots$ & $\vdots$ \\
\hline
$C_N$ & $-K$ & 0 & 0 & 0 &$\ldots$ & 0 & $\ldots$& 0&0 \\
\hline
\end{tabular}
\end{center}
where $K=\mu_0 Z+\mu_1C_1+\ldots + \mu_N C_N$.
\subsubsection{Cases 2 and 4: $t_{1x}=t_x+C_1(t_1^2-t^2)+C_2(t_1-t)$ and $t_{1x}=t_x+
C_5(e^{\alpha t_1}-e^{\alpha t})
+C_6(e^{-\alpha t_1}-e^{-\alpha t})$}
In both cases characteristic algebra $L_n$ is trivial, consisting of $X_1$ and $Y_1$ only, with
commutativity relation $[X_1, Y_1]=0$.
\subsubsection{Case 3: $t_{1x}=t_x +\sqrt{C_3e^{2\alpha t_1}+C_4e^{\alpha (t_1+t)}+C_3e^{2\alpha t}}$}
Denote by $\tilde{X}_1=A(\tau_{-1})e^{-\alpha\tau_{-1}}\frac{\partial}{\partial \tau_{-1}}$
and $\tilde{Y}_1=A(\tau_{-1})Y_1$, $C_2=[\tilde{X}_1, \tilde{Y}_1]$. Direct calculations show
that the multiplication table for algebra $L_n$ is the following
\begin{center}
\begin{tabular}{|c||c|c|c|}
\hline
$L_n$ & $\tilde{X}_1$ & $\tilde{Y}_1$ & $C_2 $ \\
\hline
\hline
$\tilde{X}_1$ & 0 & $C_2$ & $\alpha^2C_3\tilde{Y}_1+C_4/(2C_3)\tilde{X}_1 $ \\
\hline
$\tilde{Y}_1$& $-C_2$ & 0 &$ K$ \\
\hline
$C_2$ & $-\alpha^2C_3\tilde{Y}_1-C_4/(2C_3)\tilde{X}_1 $ & $-K$ & 0 \\
\hline
\end{tabular}
\end{center}
where $K=-(\alpha^2C_4/2)\tilde{Y}_1+(2\alpha^2C_4e^{\alpha\tau_{-1}}-
\alpha^2C_3)\tilde{X}_1$.
\subsection{Explicit solutions for Darbour integrable chains from Theorem \ref{maintheorem}}
Below we find explicit solutions for Darboux integrable chains of
special form (\ref{main}).
\begin{thm}\label{TheoremExplicit}
(a) The explicit solution of equation (\ref{a}) is \begin{equation}\label{aexp}
t(n,x)=t(0,x)+\sum^{n-1}_{j=0} R(x+P_j),
\end{equation}
where $t(0,x)$ and $P_j$ are arbitrary functions of $x$ and $j$
respectively, and $A(\tau)=R'(\theta)$, $t_1-t=R(\theta)$.\\
(b) The explicit solution of equation (\ref{b}) is \begin{equation}\label{bexp}
t(n,x)=\frac{1}{C_1}\left(\frac{\psi_{xx}}{2\psi_x}-\frac{\psi_x}{P_n+\psi} \right)-\frac{C_2}{2C_1},
\end{equation}
where $\psi=\psi(x)$ is an arbitrary function depending on $x$ and $P_n$
is an arbitrary function depending on $n$ only. \\
(c) The explicit solution $t(n,x)$ of equation (\ref{c}) satisfies \begin{equation}\label{cexp}
e^{\alpha t(n,x)}=\frac{\mu'(x)(R_1(P_n-P_{n+1}))}{0.25\alpha^2(\mu(x)+(P_n+P_{n+1})+
R_3(P_n-P_{n+1}))^2-C_3(R_1(P_n-P_{n+1}))^2}\, ,
\end{equation}
where $R_1=2\alpha/\sqrt{2C_3+C_4}$, $R_3=\sqrt{2C_3-C_4}/\sqrt{2C_3+C_4}$, $\mu$ and $P_n$ are
arbitrary functions depending respectively on variables $x$ and $n$. \\
(d) Equation (\ref{d}) does not admit any explicit formula for general solution of the form
\begin{equation}\label{2case4}
t=H(x, \psi(x), \psi'(x), \ldots, \psi^{(k)}(x), P_n, P_{n+1}, \ldots, P_{n+m}).
\end{equation}
However, equation (\ref{d}) admits general solution in more complicated form
\begin{equation}\label{case4}
e^{\alpha t(n,x)}=\frac{1}{\alpha C_5}\left(\frac{\psi_{xx}}{2\psi_x}-\frac{\psi_x}{P_n+\psi}\right)+\frac{1}{\alpha C_5} w\, ,
\end{equation}
where the nonlocal variable $w=w(x)$ is a solution of the first order ODE
$$
w_x'+w^2-\alpha^2C_5C_6=-\left(\frac{\psi_{xx}}{2\psi_x}\right)_x+
\left(\frac{\psi_{xx}}{2\psi_x}\right)^2.
$$
\end{thm}
\noindent{\it{Proof}}: In a trivial case (a) the explicit solution
was described in \cite{HZhP2}.\\
In case $(b)$, the equation (\ref{b}) has an $n$-integral
$I=t_x-C_1t^2-C_2t$. Since $DI=I$ then we have the following Riccati
equation $t_x-C_1t^2-C_2t=C(x)$ to solve and obtain the explicit
solution (\ref{bexp}).
In case $(c)$, to find the explicit solution of equation (\ref{c})
we look for the Cole-Hopf type substitution $t=H(v,v_1,v_x)$ that reduces the equation
to the semi-discrete D'Alembert equation $v_{1x}=v_x$ for which the solution is $v=\mu(x)+P_n$.
Let us find function $H(v,v_1,v_x)$. Since $v_{1x}=v_x$, then $t_1=H(v_1,v_2,v_x)=:\bar{H}$ and
$t_x=v_{xx}H_{v_x}+v_x(H_v+H_{v_1})$, $t_{1x}=v_{xx}\bar{H}_{v_x}+v_x(\bar{H}_{v_1}+\bar{H}_{v_2})$.
In new variables equation
(\ref{c}) becomes
\begin{equation}\label{3case3}
v_{xx}\{\bar{H}_{v_x}-H_{v_x}\}=v_x(H_v+H_{v_1}-\bar{H}_{v_1}-\bar{H}_{v_2})+
\sqrt{C_3e^{2\alpha \bar{H}}+C_4e^{\alpha(H+\bar{H})}+C_3e^{2\alpha H}}\,.
\end{equation}
The right side of (\ref{3case3}) does not depend on $v_{xx}$, but the left side does unless
$\bar{H}_{v_x}=H_{v_x}$. It implies that $H(v,v_1,v_x)=\psi(v_x)+A(v,v_1)$, where $\psi$ is a
function of one variable $v_x$ and $A$ is a function depending on $v$ and $v_1$. Now,
$t=H(v,v_1,v_x)=\psi(v_x)+A(v,v_1)$,
$t_1=\psi(v_x)+A(v_1,v_2)=:\psi(v_x)+\bar{A}$, and equality (\ref{3case3}) becomes
\begin{equation}\label{4case3}
(\bar{A}_{v_1}+\bar{A}_{v_2}-A_v-A_{v_1})v_x=e^{\alpha \psi(v_x)}
\sqrt{C_3e^{2\alpha \bar{A}}+C_4e^{\alpha(A+\bar{A})}+C_3e^{2\alpha A}}
\end{equation}
that shows $e^{\alpha \psi(v_x)}=Rv_x$, where $R$ is some constant. We have,
\begin{equation}\label{2case3}
e^{\alpha H(v,v_1,v_x)}=Rv_xe^{\alpha A(v,v_1)}\, .
\end{equation}
Let us find function $A(v,v_1)$. In variables $v_k$, $v_{[k]}$ an $n$-integral
$I=2t_{xx}-\alpha t_x^2-\alpha C_3e^{2\alpha t}
=q(x)$ of equation (\ref{c}) becomes
$$
I=\left( \frac{2v_{xxx}}{\alpha v_x}-\frac{3v^2_{xx}}{\alpha v^2_x}\right)+
v_x^2(2A_{vv}+4A_{vv_1}+2A_{v_1v_1}-\alpha(A_v+A_{v_1})^2
-\alpha C_3R^2e^{2\alpha A})
$$
$$
=:s(v_x,v_{xx}, v_{xxx})+v_x^2p(v,v_1)=q(x)\, .
$$
One can see
$$p(v,v_1)=2A_{vv}+4A_{vv_1}+2A_{v_1v_1}-\alpha(A_v+A_{v_1})^2
-\alpha C_3R^2e^{2\alpha A}=0,$$
that can be rewritten in variables $\xi=(v+v_1)/2$ and $\eta=(v-v_1)/2$ in the following form
$$
2A_{\xi\xi}-\alpha (A_\xi)^2-\alpha C_3R^2e^{2\alpha A}=0
$$
that implies
\begin{equation}\label{5case3}
e^{-\alpha A(\xi,\eta)}=\frac{\alpha^2}{4}C_1(\eta)(\xi+C_2(\eta))^2-\frac{C_3R^2}{C_1(\eta)},
\end{equation}
where $C_1(\eta)$ and $C_2(\eta)$ are some functions depending on $\eta$ only.
Equation (\ref{4case3}) implies not only $e^{\alpha \psi(v_x)}=Rv_x$, but also
$$
\bar{A}_{v_1}+\bar{A}_{v_2}-A_v-A_{v_1}=R
\sqrt{C_3e^{2\alpha \bar{A}}+C_4e^{\alpha(A+\bar{A})}+C_3e^{2\alpha A}}
$$
that in variables $\xi$ and $\eta$ becomes
\begin{equation}\label{7case3}
\bar{A}_{\xi_1}-A_\xi=R\sqrt{C_3e^{2\alpha \bar{A}}+C_4e^{\alpha(A+\bar{A})}+C_3e^{2\alpha A}}\, .
\end{equation}
We find function $A$ from (\ref{5case3}) and substitute it into (\ref{7case3}),
remembering that $\xi_1=\xi-\eta-\eta_1$. We have,
$$
\frac{4}{\alpha^2C_3R^2}\left\{-\frac{\xi-\eta-\eta_1+C_2(\eta_1)}{
(\xi-\eta-\eta_1+C_2(\eta_1))^2-4\alpha^{-2}C_3R^2}+
\frac{\xi+C_2(\eta)}{(\xi+C_2(\eta))^2-4\alpha^{-2}C_3R^2}\right\}^2
$$
$$
=\left(\frac{4\alpha^{-2}C_1^{-1}(\eta_1)}{(\xi-\eta-\eta_1+C_2(\eta_1))^2-
4\alpha^{-2}C_3R^2C_1^{-2}(\eta_1)}\right)^2 +
\left( \frac{4\alpha^{-2}C_1^{-1}(\eta)}{(\xi+C_2(\eta))^2-4\alpha^{-2}C_3R^2C_1^{-2}(\eta)}\right)^2
$$
$$
+\frac{16\alpha^{-4}C_4C_3^{-1}C_1^{-1}(\eta)C_1^{-1}(\eta_1)}{((\xi+C_2(\eta))^2-
4\alpha^{-2}C_3R^2C_1^{-2}(\eta))
((\xi-\eta-\eta_1+C_2(\eta_1))^2-4\alpha^{-2}C_3R^2C_1^{-2}(\eta_1))}\, ,
$$
or, equivalently,
\begin{equation}\label{6case3}
(C_2(\eta)+\eta-(C_2(\eta_1)-\eta_1))^2-4\alpha^{-2}C_3R^2(C_1^{-2}(\eta)+C_1^{-2}(\eta_1))\,
-4\alpha^{-2}C_4R^2C_1^{-1}(\eta)C_1^{-1}(\eta_1)=0.\end{equation}
To find $C_1(\eta)$ and $C_2(\eta)$, let us differentiate both sides of (\ref{6case3})
with respect to $\eta$
and then with respect to $\eta_1$. We have,
$$
\frac{-2C_4R^2C_1'(\eta)}{C_1^2(\eta)(C_2'(\eta)+1)}=\frac{(C_2'(\eta_1)-1)C_1^2(\eta_1)}{C_1'(\eta_1)}\, .
$$
We use the fact that the left side of the last equation depends on $\eta$ only, but the right side depends only on
$\eta_1$ and obtain
\begin{equation}\label{8case3}
C_1(\eta)=\frac{1}{R_1\eta+R_2}, \qquad C_2(\eta)=R_3\eta+R_4,
\end{equation}
where $R_1=2(D+2\alpha^{-2}C_4R^2D^{-1})^{-1}$, $R_3=(-D+2\alpha^{-2}C_4R^2D^{-1})(D+2\alpha^{-2}C_4R^2D^{-1})^{-1}$, and $D$, $R_2$, $R_4$ are some constants.
Combining formulas (\ref{2case3}), (\ref{5case3}) and (\ref{8case3}) we see that
\begin{equation}\label{9case3}
e^{\alpha t}=\frac{\mu'(x)R(R_1(P_n-P_{n+1})+R_2)}{0.25\alpha^2(\mu(x)+(P_n+P_{n+1})+
R_3(P_n-P_{n+1})+R_4)^2-C_3(R_1(P_n-P_{n+1})+R_2)^2}\, .
\end{equation}
Without loss of generality, we may assume $R=1$ and $R_4=0$. Substitution
of (\ref{9case3}) into (\ref{c}) shows that $R_2=0$,
$R_1= 2\alpha/\sqrt{2C_3+C_4}$ and $R_3=\sqrt{2C_3-C_4}/\sqrt{2C_3+C_4}$.
Let us study case $(d)$. For the sake of convenience we set $C_5=C_6=\alpha=1$.
Suppose that there exists function $H$ such that for any choice of functions $\psi(x)$ and $P_n$
function (\ref{2case4}) solves (\ref{d}).
Consider the variables $\psi$, $\psi'$, $\ldots$, $\psi^{(j)}$, $\ldots$,
$P_n$, $P_{n\pm 1}$, $\ldots$, $P_{n\pm k}$, $\ldots$ as new dynamical variables. Substitution of $t=H$,
$t_1=H(x,\psi(x), \psi'(x), \ldots, \psi^{(k)}, P_{n+1}, P_{n+2}, \ldots, P_{n+m+1})=:\bar{H}, $
$t_x=H_x+\psi'H_{\psi}+\psi''H_{\psi'}+\ldots+\psi^{(k+1)}H_{\psi^{(k)}},$
$t_{1x}=\bar{H}_x+\psi'\bar{H}_{\psi}+\psi''\bar{H}_{\psi'}+\ldots+\psi^{(k+1)}\bar{H}_{\psi^{(k)}}$
into (\ref{d}) yields
\begin{equation}\label{3case4}
\bar{H}_x+\psi'\bar{H}_{\psi}+\psi''\bar{H}_{\psi'}+\ldots+\psi^{(k+1)}\bar{H}_{\psi^{(k)}}=H_x+
\psi'H_{\psi}+\psi''H_{\psi'}+\ldots+\psi^{(k+1)}H_{\psi^{(k)}}
\end{equation}
$$
+e^{\bar{H}}+e^{-\bar{H}}-e^H-e^{-H}.
$$
Evidently, $H_{\psi^{(k)}}=\bar{H}_{\psi^{(k)}}$ and consequently $H$ can be represented as
\begin{equation}\label{H}
H=h(x, \psi, \psi', \ldots, \psi^{(k)})+r(x,\psi, \psi', \ldots, \psi^{(k-1)}, P_n, P_{n+1}, \ldots, P_{n+m}).
\end{equation}
Substitute $H$ found in (\ref{H}) in the relation $t_x-2\cosh t=C(x)$ obtained from the $n$-integral
for (\ref{d}). We have,
\begin{equation}\label{4case4}h_x+r_x+
(h_\psi+r_\psi)\psi'+(h_{\psi'}+r_{\psi'})\psi''+\ldots
\end{equation}
$$
(h_{\psi^{(k-1)}}+r_{\psi^{(k-1)}})\psi^{(k)}
+h_{\psi^{(k)}}\psi^{(k+1)}-2\cosh(h+r)=C(x).
$$
Differentiate it with respect to $P_{n+m}:=z$
\begin{equation}
r_{xz}+
r_{\psi z}\psi' + r_{\psi' z}\psi''+\dots + r_{\psi^{k-1} z}\psi^{(k)}-2r_z\cosh(h+r) =0.
\end{equation}
We find $h+r$ from the last equation
\begin{equation}\label{5case4}
h+r=arccosh \left( \frac{r_{xz}+r_{\psi z} \psi' + \dots +r_{\psi^{k-1} z} \psi^{(k)}}{2r_z}\right)\, .
\end{equation}
Denote $r_{xz}+r_{\psi z} \psi' + \dots +r_{\psi^{k-2} z}
\psi^{(k-1)}=A(x,\psi, \psi', \psi^{(k-1)}, P_n,\dots,P_{n+m})$.
Differentiate (\ref{5case4}) with respect to $z=P_{n+m}$ and set
\begin{equation}
r_z= \frac{\left(\frac{A+r_{\psi^{(k-1)} z} \psi^{(k)}} {2r_z}\right)_z}{
\sqrt{\left(\frac{A+r_{\psi^{(k-1)} z} \psi^{(k)}} {2r_z}\right)^2-1}},
\end{equation}
where $A$ and $r$ do not depend on $\psi^{(k)}$. Let us denote $p:=r_z$, $y:=\psi^{(k-1)}$ and $\xi:=\psi^{(k)}$
then
\begin{equation}
p^2\left( \frac{A+p_y\xi}{2p}\right)^2-p^2=\left( \frac{A+p_y\xi}{2p}\right)_z=\left(
\frac{(A_z+p_{yz}\xi)p-p_z(A+p_y\xi)}{2p^2}\right)^2\,
\end{equation}
or
\begin{equation}\label{6case4}
p^4(A+p_y\xi)^2-4p^6=((A_zp-p_zA)+(p_{yz}p-p_yp_z)\xi)^2 .
\end{equation}
Comparison of the coefficients in (\ref{6case4}) gives rise to three
equalities: $p^4p_y^2=(p_{yz}p-p_yp_z)^2$, $p^4Ap_y
=(A_zp-p_zA)(p_{yz}p-p_yp_z)$
and $p^4A^2-4p^6=(A_zp-p_zA)^2$,
that are consistent only if $p=r_z=r_{P_{n+m}}=0$. This condition
$H_{P_{n+m}}=0$ contradicts to our assumption that $H$ essentially
depends on $P_{n+m}$. (Note that if $H_{P_k}=0$ for all $k\in \ZZ$,
then we have a trivial solution $t=H(x)$ for equation (\ref{d}).)
Therefore, in case (d) the solution can not be represented in form
(\ref{2case4}).
One can use $n$-integral $I=t_x-C_5e^{\alpha t}-C_6e^{-\alpha t}$, solve the equation $I=C(x)$
which with the help of the substitution $u=e^{\alpha t}$ can be brought to Riccati equation, and
see that equation (\ref{d}) admits a general solution in more complicated form (\ref{case4}).
\section{Conclusions}
Darboux integrable semi-discrete chains are studied. Structures of their integrals are described. It is proved that if the chain admits an $n$-integral of order $k$ then it admits also an $n$-integral linearly depending on the highest order variable $t_{[k]}$. Similarly,
if the chain admits an $x$-integral $F(x,t_{-k},t_{-k+1},...t_{m})$, then there is an $x$-integral $F^0(x,t_{-k},t_{-k+1},...t_{m})$
solving the equation
$$\frac{\partial^2F^0}{\partial t_{-k}\partial t_{m}}=0.$$
The previously found list of Darboux integrable chains of a particular form $t_{1x}=t_x+d(t,t_1)$ is studied in details. The tables of multiplication for the corresponding characteristic algebras are given, explicit formulas for general solutions are constructed.
The problem of complete classification of Darboux integrable chains (\ref{dhyp}) is still open. Another important open problem is connected with systems of discrete equations: find Darboux integrable discrete versions of the exponential type hyperbolic systems corresponding to the Cartan matrices of semi-simple Lie algebras.
\section*{Acknowledgments}
This work is partially supported by the Scientific and
Technological Research Council of Turkey (T\"{U}B{\.{I}}TAK) grant $\# $209 T 062,
Russian Foundation for Basic
Research (RFBR) (grants $\#$ 09-01-92431KE-a, $\#$
08-01-00440-a, $\#$ 10-01-91222-CT-a and $\#$ 10-01-00088-a), and MK-8247.2010.1.
|
\section{Introduction}
Quantum critical behavior in metals is a topic of prime interest for the
theory and experiment in the field of condensed matter physics.
\cite{sachdev_book, vojta_review03, loehneysen_review06, belitz_review05,
abanov03, gegenwart08,stewart01} Despite significant
effort made over the last couple of years, several intriguing puzzles remain
unresolved by the quickly developing theory of quantum phase transitions.
These include both, fundamental problems regarding the correct low energy action
to describe quantum critical systems, and the physical mechanisms responsible for
unusual behavior of specific compounds.
In the conventional scenario of quantum criticality the critical temperature
$T_c$ for a second order transition is driven to zero by a tuning parameter,
so that a quantum critical point emerges. However, it is also possible that
below a tricritical temperature the transition becomes first
order. Then a particular brand of quantum criticality
arises when the finite temperature ($T$) phase boundary terminates at $T=0$
with the tricritical point. Strictly speaking, in real systems this scenario is
hardly feasible, since it requires fine tuning of two non-thermal parameters.
However, proximity to quantum tricriticality has been invoked to explain a number
of experimentally observed properties of metallic compounds.
\cite{pfleiderer01, uhlarz04, belitz05, green05, misawa08, misawa09}
Although the relevance of quantum tricriticality was emphasized in a number of recent
works,\cite{belitz05, green05, misawa08, misawa09} a renormalization group (RG) study
addressing crossovers between critical and tricritical behavior, in particular in the
quantum-critical, non-Fermi liquid regime, is missing.
In this paper we present a renormalization group analysis for a model system
displaying such crossovers.
We build upon an earlier work,\cite{jakubczyk09} where renormalization group
flows of a quantum $\phi^6$-model were already studied. The main focus of that
work was to show that order parameter fluctuations may turn first order transitions
occurring within the bare model into continuous transitions.
The paper is structured as follows. In Sec.~II we present a phenomenological
picture of the problem to be analyzed, and in Sec.~III we define the corresponding
model action.
In Sec.~IV we describe the functional RG framework and derive the RG flow equations.
In Sec.~V we analyze the linearized forms of the flow equations obtaining
analytical solutions in spatial dimensions $d>2$.
In Sec.~VI we present example numerical results for the
correlation length in the quantum critical regime, and the shapes of the phase
boundaries. In the final Sec.~VII we lay out our conclusions.
\section{Phenomenological picture}
Before plunging into the details of the formalism let us consider the
different scenarios which are envisaged in the following.
In Fig.~\ref{Phase_diag_pheno} (a) we schematically depict a generic phenomenological
phase diagram of a system exhibiting a Gaussian quantum critical point (QCP).
At $T=0$ the system can be tuned between an ordered and a disordered state
by varying a non-thermal control parameter $r$.
At finite $T$, in the disordered phase, a crossover region, schematically represented here
with a line, separates the Fermi liquid and quantum critical regimes.
For $r<0$, the $T_c$-line separates the phases of the system at finite $T$. The
classical critical region where non-Gaussian fluctuations occur is bounded by the
Ginzburg lines and vanishes as $T \to 0$.
\begin{figure}[ht!]
\begin{center}
\includegraphics[width=2.1in]{Phase_diag_pheno.eps}
\includegraphics[width=2.1in]{Phase_diag_pheno3.eps}
\includegraphics[width=2.1in]{Phase_diag_pheno2.eps}
\caption{(Color online) Different scenarios for generic phase diagrams.
Bold solid (dotted) lines mark second (first) order phase transitions,
thin solid lines indicate the crossover between quantum critical and
Fermi liquid behavior, while the thin dotted lines indicate the
boundaries of the Ginzburg region.
(a) Schematic phase diagram of a system exhibiting a
quantum critical point, at which a line of second order finite $T$ phase
transitions terminates. A crossover region, represented here with a line, separates
the Fermi liquid and quantum critical regimes in the disordered phase.
(b) Schematic phase diagram of a system exhibiting a quantum tricritical point.
(c) Schematic phase diagram of a system exhibiting a tricritical point at a finite
temperature. The transition is first order below the tricritical
temperature. Critical quantum fluctuations are absent. Varying another
control coupling one can tune between the depicted situations (see the main
text).}
\label{Phase_diag_pheno}
\end{center}
\end{figure}
By varying another system parameter the phase diagram can be continuously deformed so
that the transition becomes first order at sufficiently low $T$, below a
tricritical temperature $T^{\rm tri}$ (see Fig.~\ref{Phase_diag_pheno} (c)). For a
particular value of this parameter the tricritical point is located exactly at
$T=0$ so that the transition is second order for all $T>0$ but the scaling
properties of the system for $T\to 0$ are governed by the quantum tricritical point (QTCP)
rather than the quantum critical point. This special scenario displaying a
different scaling behavior is depicted in Fig.~\ref{Phase_diag_pheno} (b).
The purpose of the paper is to compute how the crossovers between the
scenarios of Fig.~\ref{Phase_diag_pheno} (a) through (b) to
Fig.~\ref{Phase_diag_pheno} (c) occur.
Even in scenario (a) a ``hidden'' tricritical point, which would be
present at $T^{\rm tri}<0$, i.e. the vicinity to a first order transition,
can affect the scaling behavior at higher temperatures.
\section{Bare action}
The conventional description of quantum criticality, which we shall rely upon here,
invokes the Hertz action.\cite{hertz76} This describes a bosonic mode overdamped by
particle-hole excitations across the Fermi level and is applicable under the assumption
that the electronic degrees of freedom may be integrated out.
The original framework of Hertz \cite{hertz76} and Millis \cite{millis93} was recently
extended to account for a number of systems and phenomena not covered by the original
studies.
These include field-tuned quantum critical points, \cite{fischer_rosch05} metamagnetic
transitions,\cite{millis_schofield02, SMGL02} phase transitions induced by a non-equilibrium
drive \cite{mitra_takei06} as well as dimensional crossovers \cite{garst08} and quantum
criticality involving multiple time scales. \cite{zacharias09}
We focus on the case of discrete symmetry breaking, which can be described by a
scalar order parameter field $\phi$, and an action of the form
\begin{eqnarray}
S[\phi] =
\frac{T}{2} \sum_{\omega_n} \int\frac{d^dp}{(2\pi)^d} \,
\phi_p \left(Z_{\omega} \frac{|\omega_{n}|}{|\mathbf{p}|^{z-2}}
+ Z\mathbf{p}^{2} \right) \phi_{-p} + {\cal U}[\phi] \; .
\label{eq:lagrangian}
\end{eqnarray}
Here $\phi_p$ with $p = (\mathbf{p},\omega_n)$ is the momentum representation of the
order parameter field, where $\omega_n = 2\pi n T$ with integer $n$ denotes the (bosonic)
Matsubara frequencies.
Momentum and energy units can be chosen such that the prefactors in front of
$\frac{|\omega_{n}|}{|\mathbf{p}|^{z-2}}$ and $\mathbf{p}^{2}$ are equal to unity.
The action is regularized in the ultraviolet by restricting momenta to
$|\mathbf{p}| \leq \Lambda_0$. The term $\frac{|\omega_{n}|}{|\mathbf{p}|^{z-2}}$,
where $z\ge 2$ is the dynamical exponent, effectively accounts for overdamping of
the order-parameter fluctuations by fermionic excitations across the Fermi surface.
The expression Eq.~(\ref{eq:lagrangian}) is valid for $|\mathbf{p}|$, and
$\frac{|\omega_{n}|}{|\mathbf{p}|^{z-2}}$ sufficiently small, which is the
limit relevant for the physical situation considered.\cite{millis93}
The quantity ${\cal U}[\phi]$ is a local effective potential which we expand to
sixth order in $\phi$:
\begin{equation}
{\cal U}[\phi] =
\int_0^{1/T} d\tau \int \! d^d x \, U \left(\phi\left(x,\tau\right)\right) \; ,
\end{equation}
where
\begin{equation}
U(\phi) = a_2 \phi^2 + a_4 \phi^4 + a_6 \phi^6 \; .
\label{eq:ef_potential}
\end{equation}
This ansatz assumes the absence of a field explicitly breaking the
inversion symmetry.
We require $a_6 > 0$ to stabilize the system at large $|\phi|$.
The bare effective potential yields the well known phase diagram, \cite{lawrie84}
exhibiting a second-order transition for $a_4 > 0$, a first order transition for
$a_4 < 0$ and a tricritical point at $a_2 = a_4 = 0$. Within the present model the
mass parameter $a_2$ plays the role of the non-thermal control parameter ($r$) to
tune the transition.
By varying the quartic coupling $a_4$ one can deform the phase diagram so that the
transition is second or first order at low $T$ (compare Fig.~\ref{Phase_diag_pheno}).
In conventional quantum criticality, the temperature dependence of the coefficients
$a_2$, $a_4$, $a_6$ can be neglected, as it leads only to subleading corrections.
However, it turns out that in a quantum tricritical regime the generic quadratic
temperature dependence of these coefficients yields a dominant contribution to the
temperature dependence of the correlation length.
\section{Flow equations}
In the present study we apply the one-particle irreducible variant of the
functional RG. \cite{berges_review02, delamotte07, pawlowski_review07, gies_notes06}
The derivation of the flow equations follows Ref.~\onlinecite{jakubczyk09}, where the present
quantum $\phi^6$ model was applied to analyze the effect of fluctuations on the
order of quantum phase transitions and the shapes of the finite $T$ phase
boundaries. The starting point is the exact functional evolution equation \cite{wetterich93}
\begin{eqnarray}
\frac{\partial}{\partial \Lambda}\Gamma^{\Lambda}\left[\phi\right]=
\frac{1}{2}\text{Tr}\frac{\partial_\Lambda R^{\Lambda}}{\Gamma^{(2)}\left[\phi\right]
+ R^{\Lambda}}
\label{eq:flow_eqn}
\end{eqnarray}
describing the flow of the effective action $\Gamma^{\Lambda}[\phi]$, which is the
generating functional for one-particle irreducible vertex functions as the
infrared cutoff scale $\Lambda$ is reduced.
Here $\Gamma^{(2)}\left[\phi\right] = \delta^{2}\Gamma^{\Lambda}[\phi]/\delta \phi^{2}$,
$\text{Tr} = T \sum_{\omega_{n}} \int \frac{d^{d} p}{\left(2\pi\right)^{d}}$,
finally $R^{\Lambda}$ denotes the cutoff function added to the inverse propagator
to cut off modes with momentum below the scale $\Lambda$.
We implement the Litim cutoff \cite{litim01}
\begin{equation}
R^{\Lambda}(\mathbf{p}) =
Z \left( \Lambda^{2}-\mathbf{p}^{2}\right)
\theta\left(\Lambda^{2}-\mathbf{p}^{2} \right) \; .
\label{Litim_fun}
\end{equation}
For $\Lambda=\Lambda_0$ the quantity
$\Gamma^{\Lambda}$ is just the bare action, while in the limit $\Lambda\to 0$
it converges to the full effective action, i.e., the Gibbs free energy.
In most cases it is impossible to solve Eq.~(\ref{eq:flow_eqn}) in the full
functional space.
One may however cast it onto a suitably chosen set of flowing couplings.
The approximation applied here amounts to assuming that the effective potential
preserves the form given by Eq. (\ref{eq:ef_potential}) with flowing
couplings, while the inverse propagator retains its initial form
\begin{eqnarray}
G^{-1}(\textbf{p},\omega_n) &=&
\Gamma^{(2)}\left[\phi=\phi_0\right] + R^{\Lambda}(\textbf{p}) \nonumber \\
&=& Z_{\omega} \frac{|\omega_{n}|}{|\mathbf{p}|^{z-2}} +
Z\mathbf{p}^2 + 2a_2 + R^{\Lambda}(\textbf{p}) \; ,
\label{Green}
\end{eqnarray}
where $\phi_0$ is the minimum of $U(\phi)$.
The present study deals with situations where the quantum critical point is Gaussian.
In the following we shall disregard the flow of $Z$, which is equivalent to neglecting
the anomalous dimension of the order parameter field.
Such a parameterization of the effective action does not capture the anomalous scaling
behavior of the propagator in the narrow vicinity of the classical second-order phase
transition at finite temperature.
Analogous truncations retaining the flow of $Z$ were applied in
Refs.~\onlinecite{jakubczyk09, jakubczyk08}.
The results of Ref.~\onlinecite{jakubczyk08} indicate that universal aspects of the
shape of the phase boundary are not affected by non-Gaussian thermal fluctuations.
In the following we shall also neglect the renormalization of the factor $Z_{\omega}$.
The flow of $Z_{\omega}$ was computed in Ref.~\onlinecite{jakubczyk08}, and shown to
be negligible.
The present truncation reproduces the essential features of the system except for the narrow
vicinity of the second order transition at $T>0$, where nonetheless the shape of
the phase boundary is described correctly.
Note, however, that the approach can be adapted to capture the anomalous
dimension of the order parameter field \cite{jakubczyk08} and to deal with cases
where the fixed point associated with the quantum critical point is not Gaussian.
\cite{strack09}
The present truncation is analogous to the zeroth order derivative expansion,
\cite{berges_review02, delamotte07} which was applied extensively in the context
of classical critical phenomena. The essential quantum ingredient present here is
the Landau damping term $\frac{|\omega_n|}{|\textbf{p}|^{z-2}}$.
On top of the derivative expansion we performed a polynomial expansion of the
effective potential.
Relying on the truncation described above, the flow equation Eq. (\ref{eq:flow_eqn})
can be reduced to a set of three ordinary differential equations which determine the
flow of the couplings $a_2$, $a_4$, $a_6$ as functions of the cutoff scale $\Lambda$.
In practice, whenever the effective potential features non-zero minima at $\pm
\phi_0$, we find it more convenient to write the flow equations in terms of
the variables $\rho_0=\frac{1}{2} \phi_0^2$, $a_4$, $a_6$. The coupling $a_2$
is then obtained from
\begin{equation}
a_2=-4\rho_0(a_4+3a_6\rho_0)\;.
\label{a_2}
\end{equation}
As long as there exists a non-zero $\rho_0$, the evolution of the effective
potential is given by the flow equations derived in
Ref.~\onlinecite{jakubczyk09},
\begin{equation}
\partial_t \rho_0 =
2v_d Z^{-1}\Lambda^{d-2}\left(3+2\frac{6a_6\rho_0}{6a_6\rho_0+a_4}\right)l_1^d\;,
\label{flow_rho_0}
\end{equation}
\vspace{0.1cm}
\begin{equation}
\partial_t a_4 =
12v_d\Lambda^d\left[\frac{4}{3}l_2^d\frac{(30a_6\rho_0+3a_4)^2}{Z^2\Lambda^4}
-5l_1^d\frac{a_6}{Z\Lambda^2}\right]-6\rho_0\partial_t a_6\;,
\label{flow_a4}
\end{equation}
\begin{equation}
\partial_t a_6 =
16v_d\Lambda^d\left[-\frac{8}{3}l_3^d\frac{(30a_6\rho_0+3a_4)^3}{Z^3\Lambda^6}+15l_2^d
a_6\frac{30a_6\rho_0+3a_4}{Z^2\Lambda^4}\right],
\label{flow_a6}
\end{equation}
where $v_d^{-1}=2^{d+1}\pi^{d/2}\Gamma(d/2)$ and $t = \log(\Lambda/\Lambda_0) \leq 0$.
The threshold \cite{berges_review02} functions $l_i^d(\delta)$ are determined from
\begin{equation}
l_0^d(\delta) =
\frac{1}{4}v_d^{-1}\Lambda^{-d}\text{Tr} \, \frac{\partial_t R^{\Lambda}(\mathbf{p})}
{Z_{\omega}\frac{|\omega_n|}{|\textbf{p}|} + Z\textbf{p}^2 +
R^\Lambda(\mathbf{p}) + Z\Lambda^2\delta}\; ,
\label{l0}
\end{equation}
\begin{equation}
l_1^d(\delta) = -\frac{\partial}{\partial \delta}l_0^d(\delta)\; ,
\label{l1}
\end{equation}
\begin{equation}
l_n^d(\delta) =
-\frac{1}{n-1}\frac{\partial}{\partial \delta}l_{n-1}^d(\delta)\;,\;\;\; n\geq 2\;.
\end{equation}
In Eqs.~(\ref{flow_rho_0}, \ref{flow_a4}, \ref{flow_a6})
$\delta=\delta(\rho)=\frac{1}{Z\Lambda^2}(U'(\rho)+2\rho U''(\rho))$ and the
threshold functions are evaluated at the value of $\delta$ corresponding to
$\rho_0$, that is, $l_n^d=l_n^d(\delta)|_{\rho=\rho_0}$.
The flow equations are illustrated in terms of Feynman diagrams in Fig.~\ref{Diagrams}.
For $\rho_0=0$ the flow equations are simplified as the interaction vertices
involving an odd number of legs (see Fig.~\ref{Diagrams}) vanish. In terms of
the couplings $a_2$, $a_4$, $a_6$ they yield
\begin{equation}
\partial_t a_2 = -24v_d\frac{a_4}{Z\Lambda^{2-d}}l_1^d\;,
\end{equation}
\begin{equation}
\partial_t a_4 =
v_d\Lambda^d \left[\left(12\frac{a_4}{Z\Lambda^2}\right)^2l_2^d -
60\frac{a_6}{Z\Lambda^2}l_1^d\right] \; ,
\end{equation}
\begin{equation}
\partial_t a_6 = v_d \Lambda^d
\left[-\frac{2}{3}\left(12\frac{a_4}{Z\Lambda^2}\right)^3l_3^d+
\left(12\frac{a_4}{Z\Lambda^2}\right)\left(60\frac{a_6}{Z\Lambda^2}\right)l_2^d\right]
\;.
\end{equation}
\begin{figure}[ht!]
\begin{center}
\includegraphics[width=3.5in]{Diagrams.eps}
\caption{Feynman diagrams representing the contributions to the
flow equations for the couplings parameterizing the effective
potential. Interactions involving an odd number of legs vanish
if the system is in the disordered phase.}
\label{Diagrams}
\end{center}
\end{figure}
\section{Linearized flow equations}
Essential features of the solutions to the equations provided in the previous
section (the values of the exponents describing the system in the quantum-critical
regime in particular) can be computed analytically by taking only the dominant
terms in the flow equations into account.
Here we analyze the linearized version of the flow equations in spatial dimension $d>2$.
We shall use the following rescaled variables:
\begin{equation}
\delta = \frac{2a_2}{Z\Lambda^2},\;\; u =
\frac{a_4}{Z^2\Lambda^{4-d}},\;\; v =
\frac{a_6}{Z^3\Lambda^{6-2d}}\;.
\end{equation}
We linearize the flow equations around the zero temperature Gaussian fixed point
\begin{equation}
\partial_t\delta=-2\delta-48v_d l_1^d(\delta)u\;,
\label{eq:delta_flow_i}
\end{equation}
\begin{equation}
\partial_t u =(d-4)u-60v_dl_1^d(\delta)v\;,
\label{eq:u_flow_lin_i}
\end{equation}
\begin{equation}
\partial_t v = (2d-6)v\;.
\label{eq:v_flow}
\end{equation}
All the resulting flow equations involve contributions from rescaling. In
addition, renormalization of the couplings $\delta$, $u$ by the diagrams
involving just one interaction vertex is retained. After this step, the terms
involved are analogous to those analyzed by Millis, \cite{millis93} where
$v=0$, so that only mass renormalization from the tadpole diagram was
considered. Note that in the present case of discrete symmetry-breaking,
Eq.~(\ref{eq:delta_flow_i}, \ref{eq:u_flow_lin_i}, \ref{eq:v_flow}) are applicable
both in the symmetric and symmetry-broken phases.
In the above form we still find the flow equations rather hard to solve by
hand as the threshold function $l_1^d(\delta)$ involves integrals (see
Eqs.~(\ref{l0},\ref{l1})). Inspired by the seminal work by Millis,\cite{millis93}
we additionally expand the function $l_1^d$ for
$\tilde{T}=\frac{2\pi T Z_{\omega}}{Z\Lambda ^z}\ll 1$ and $\tilde{T}\gg 1$.
In the regime $\tilde{T}\ll 1$ the flow is dominated by quantum physics,
while for $\tilde{T}\gg 1$ the quantum contributions to the flow equations
are negligible compared to the classical ones. Similarly to
Ref.~\onlinecite{millis93} we shall integrate the flow equations from
$\Lambda=\Lambda_0$ to $\Lambda=\Lambda^*$ determined by $\tilde{T}(\Lambda^*)=1$
using the equations valid strictly speaking for $\tilde{T}\ll 1$. The result
will then serve as the initial condition for the solution of the flow from
$\Lambda^*$ to $\Lambda =0$, where we shall in turn use the asymptotic flow
equations valid in the classical sector $\tilde{T}\gg 1$.
Performing the abovementioned expansion of the threshold function $l_1^d$ we find
for $\tilde{T}\ll 1$
\begin{equation}
\partial_t \delta \approx -2\delta -48v_d\frac{4T}{\tilde{T}}\frac{u}{d+z-2}\;,
\label{eq:delta_flow_lin_ini}
\end{equation}
\begin{equation}
\partial_t u \approx (d-4)u-60v_d\frac{4T}{\tilde{T}}\frac{v}{d+z-2}\; ;
\end{equation}
while for $\tilde{T}\gg 1$
\begin{equation}
\partial_t \delta \approx -2\delta -48v_d\frac{2T}{d}u\;,
\label{eq:delta_flow_lin_fin}
\end{equation}
\begin{equation}
\partial_t u =(d-4) u -60 v_d \frac{2T}{d}v\;.
\end{equation}
The above equations, together with Eq.~(\ref{eq:v_flow}), form a set of linear
ordinary differential equations of first order. Eq.~(\ref{eq:v_flow}) is
solved immediately. Plugging the result into the flow of $u$ and integrating,
we calculate $u(\Lambda)$ for $\Lambda>\Lambda^*$ and
$\Lambda<\Lambda^*$. Demanding that $a_4^r\equiv\lim_{\Lambda\to 0}
a_4(\Lambda)=\lim_{\Lambda\to 0} Z^2\Lambda^{4-d}u(\Lambda)>0$ we find a condition for
the occurrence of a second order transition. If $a_4^r(T)=0$ at some $T\geq 0$,
there is a tricritical point
at this $T$. We find that the transition is always second order at
sufficiently high $T$. It may turn first order as $T$ approaches zero. The
condition $a_4^r>0$ yields
\begin{equation}
u_0 > -v_0 \left( C + C' T^{(d+z-2)/z}\right)\;,
\label{eq:condition_for_u}
\end{equation}
where $u_0 = u(\Lambda=\Lambda_0)$, $v_0 = v(\Lambda_0)$, and
$C$, $C'$ are positive constants (depending on $\Lambda_0$, $Z_\omega$, $Z$,
$d$ and $z$). The explicit expressions for these constants are given in the
Appendix.
If the condition for $u_0$ given by Eq.~(\ref{eq:condition_for_u}) is fulfilled
for $T=0$, it clearly holds for any $T\geq 0$. In such case there is a line of
critical points terminating at $T=0$ with a QCP. This is the scenario (a) in
Fig.~\ref{Phase_diag_pheno}. For $v_0=0$ one
obviously recovers $u_0>0$ and the standard quantum critical behavior. A
quantum tricritical point exists if
$u_0= u_0^{\rm tri}=-Cv_0$, which corresponds to the quantum
tricritical scenario (b) in Fig.~\ref{Phase_diag_pheno}.
In case the condition Eq.~(\ref{eq:condition_for_u})
is violated at $T=0$, it is still always fulfilled at sufficiently high
temperatures, namely for $T > T^{\rm tri}$, where
\begin{equation}
T^{\rm tri} = [-(u_0/v_0+C)/C']^{z/(d+z-2)} \; .
\end{equation}
In this case the transition becomes first order for $T<T^{\rm tri}$,
which is scenario (c) in Fig.~\ref{Phase_diag_pheno}.
We now discuss the flow of $\delta$. Plugging the solution obtained for
$u(\Lambda)$ into Eq.~(\ref{eq:delta_flow_lin_ini},
\ref{eq:delta_flow_lin_fin}), we integrate Eq.~(\ref{eq:delta_flow_lin_ini})
from $\Lambda_0$ to $\Lambda^*$. The solution at $\Lambda^*$ is the initial
condition for Eq.~(\ref{eq:delta_flow_lin_fin}). In cases where critical
behavior occurs, the correlation length can be extracted as
$\xi^{-2}\propto\lim_{\Lambda\to 0}\Lambda^2\delta$.
The result has the following structure:
\begin{equation}
\xi^{-2} = D_0(T) + D_1 T^{(d+z-2)/z} + v_0 D_2 T^{(2(d+z)-4)/z} \; ,
\label{eq:corr_length_lin}
\end{equation}
where $D_0$ depends on $\delta_0$, $u_0$ and $v_0$, while $D_1$ and $D_2$ do not
depend on $\delta_0$.
Notice that $D_1 = \bar D_1(u_0 - u_0^{\rm tri})$, where $\bar D_1 > 0$.
Explicit expressions are quoted in the Appendix.
The generically quadratic temperature dependence of the bare model parameters
$\delta_0$, $u_0$ and $v_0$ for low $T$ induces a corresponding quadratic
temperature dependence of $D_0$, $D_1$, $D_2$. While the temperature dependence
of $D_1$ and $D_2$ is always negligible, the quadratic $T$-dependence of
$D_0(T) = D_0 + D'_0 T^2$ dominates in the tricritical regime (see below) and must
therefore be kept.\cite{misawa08,misawa09}
Criticality occurs at $T=0$ when $\delta_0 = \delta_0^{\rm cr}$ such that $D_0 = 0$.
At finite $T$ the behavior of $\xi^{-2}$ is governed by several terms.
The quantum critical term $D_1 T^{(d+z-2)/z}$ dominates for sufficiently
low $T$, provided $D_1 > 0$, and reproduces the result for $\xi^{-2}(T)$ derived
already by Millis.\cite{millis93} Compared to that term, the quadratic temperature
dependence from $D_0(T)$ and the term $v_0 D_2 T^{(2(d+z)-4)/z}$ coming from the
$\phi^6$ interaction are subleading corrections.
On the other hand, if $u_0 = u_0^{\rm tri}$, the coefficient $D_1$ becomes zero and
the correlation length exhibits a different behavior,
\begin{equation}
\xi^{-2} = D'_0 T^2 + v_0 D_2 T^{(2(d+z)-4)/z} \; .
\end{equation}
The exponent of the second term is $3$ for $z=2$, $d=3$, and $8/3$ for $z=3$, $d=3$.
The latter exponent was obtained already from a self-consistent fluctuation
resummation.\cite{green05}
At low temperatures the ''trivial'' quadratic term thus dominates the temperature
dependence of $\xi$ in the tricritical regime.
The relevance of the quadratic temperature dependences of the bare parameters in the
tricritical regime was noticed already by Misawa et al.\cite{misawa08,misawa09}
However, the exponent for the temperature dependence of the order parameter
susceptibility obtained by these authors for the quantum tricritical regime
turned out to be identical to that for conventional quantum criticality, that is,
$3/2$ for $z=2$.
That result, obtained from a self-consistent summation of staggered and homogeneous
fluctuations, is clearly at variance with our result.\cite{fn1}
For quantum critical systems close to quantum tricriticality (small $D_1$), the
correlation length follows a tricritical temperature dependence above a crossover
temperature
\begin{equation}
T_{\rm cross} = \left( \frac{D_1}{D'_0} \right)^{z/(2+z-d)}\;.
\label{Tcross1}
\end{equation}
For the special case of a very small $D'_0$, tricritical behavior with an
exponent $\frac{2(d+z)-4}{z}$ is observed for temperatures above
\begin{equation}
T_{\rm cross} = \left( \frac{D_1}{v_0 D_2} \right)^{z/(d+z-2)} \; .
\label{Tcross2}
\end{equation}
Eq.~(\ref{eq:corr_length_lin}) also yields the shape of the
phase boundary. From the condition $\xi^{-2}=0$ one finds
$T_c \sim |\delta_0-\delta_0^{\rm cr}|^{\psi}$, where
$\psi = \psi^{\rm qc} = \frac{z}{d+z-2}$
for quantum criticality, reproducing the result by Millis.\cite{millis93}
This value crosses over to $\psi = \psi^{\rm tri} = \frac{1}{2}$ when the
quartic coupling approaches the tricritical value $u_0^{\rm tri}$.
Only for the special case $D'_0 = 0$ one obtains
$\psi^{\rm tri} = \frac{z}{2(d+z)-4}$. For small $D'_0 \neq 0$ a crossover
between different exponents occurs.
As already stated, for $u_0 < u_0^{\rm tri}$ a classical tricritical point exists at
$T = T^{\rm tri}>0$. In Eq.~(\ref{eq:corr_length_lin}) this manifests itself by a
negative value of $D_1$. From the condition $\xi^{-2} = 0$ we find the shape of
the phase boundary above $T = T^{\rm tri}$ to follow
$|\delta_0-\delta_0^{\rm tri}| \approx a|T-T^{\rm tri}|+b|T-T^{\rm tri}|^2$.
This reproduces a MFT result for classical tricritical points.\cite{lawrie84}
From the solution for $\delta(\Lambda)$ we also read off the shape of the
crossover line separating the Fermi liquid and the quantum critical regimes.
The condition for the occurrence of the quantum disordered (Fermi liquid)
regime \cite{millis93} is that the correlation length becomes of the order
of the inverse upper cutoff before classical scaling is reached,
that is, $\delta (\Lambda^*) > \Lambda_0^2$. From this condition we
find the shape of the crossover line as $\delta_0-\delta_0^{\rm cr}\sim
T^{2/z}$, as can also be deduced by a phenomenological
reasoning.\cite{sachdev_book, vojta_review03}
The linearized flow equations discussed above are not applicable in two
dimensions. The reason is easily understood from the diagrammatic interpretation
provided in Fig.~\ref{Diagrams}.
The tadpole diagram renormalizing the $a_2$ coefficient exhibits a logarithmic
infrared singularity for $\xi \to \infty$ at $T>0$.
The divergence is cured when one accounts for the renormalization of the quartic
interaction coupling via the terms of the order $a_4^2$.
\section{Numerical solutions to the flow equations}
Here we present results from direct numerical solutions to the flow equations
delivered in Sec. IV. Information concerning the physical state of the system
is read off from the renormalized values of the quantities $a_2$, $a_4$, $a_6$
in the limit $\Lambda \to 0$, $a_n^r=\lim_{\Lambda\to 0}a_n(\Lambda)$. In all
numerical computations we put $Z=1$, $Z_\omega =1$, $\Lambda_0=1$ and
$a^0_6=a_6(\Lambda_0)=1$. We omit the
quadratic $T$-dependence of $a_2$ in this section. We analyze solutions to the
flow equations upon varying the parameters
$a_2^0=a_2(\Lambda_0)$, $a^0_4=a_4(\Lambda_0)$ and $T$.
We refer to Ref.~\onlinecite{jakubczyk09} for example plots of the
numerically computed phase boundaries including cases involving a first order
transition. Here we focus on the situation where $a^0_4$ is close to a
$a_4^{0,\rm tri}$, a value where a tricritical point exists at $T=0$ for some
$a^0_2$. In the analytical calculation based on the linearized equations one finds
the quantum tricritical point for the choice
$a_4^0 = a_4^{0,\rm tri} = -\frac{120\Lambda_0^{d+z-2}v_d}{\pi(d+z-2)^2}a_6^0$.
For $d=z=3$, $\Lambda_0 = a_6^0 = 1$ this gives
$a_4^{0,\rm tri} \approx -0.0302358$.
In the numerical solution for the full flow equations
a value for $a_4^{0,\rm tri}$ emerges on fine tuning $a^0_2$ and $a^0_4$,
such that $a^r_2\approx a^r_4\approx 0$, where the value for $a^0_4$ tends to
come out a bit (by less than 1\%) smaller than the analytical estimate.
Once the transition line is established, we can fix $a^0_4$ and
investigate the quantum critical regime spanning over the quantum critical or
tricritical point.
In particular, we extract the correlation length $\xi$ as function
of $T$. In the region of interest the system is typically ordered within the
bare model ($a^0_4<0$, $a^0_2>0$, but $(a^0_4)^2>4a^0_2a_6^0$), but becomes
disordered in the course of the flow, as the order parameter becomes zero for
some finite $\Lambda$. At this scale we have to switch to the set of
equations valid in the symmetric phase (see Sec.~IV).
We first show results for pure tricritical scaling of the inverse correlation
length $\xi^{-2}$ as a function of temperature (see Fig.~\ref{xi_tricrit_d3z3}).
We set $d=z=3$ and we choose the parameter $a^0_4\approx
a_4^{0,\rm tri}$ and $a^0_2$ tuned to the zero temperature quantum tricritical
point. This corresponds to the situation depicted in Fig.~1 (b), where the QTCP
is approached from above.
\begin{figure}[ht!]
\begin{center}
\includegraphics[width=3.0in]{figures_qucrit_joh/xim2vsTparaset_4.eps}
\caption{(Color online) The logarithm of the inverse correlation length
$\xi^{-2}$ versus $\log(T)$ in the quantum tricritical regime in
$d=z=3$. The special choice of parameters is $a_4^0 = -0.030270317439$ and
$a_2^0=0.00018421858$.}
\label{xi_tricrit_d3z3}
\end{center}
\end{figure}
Over a wide range of temperature we find that the slope $\epsilon=2.66$ gives
a good fit to the numerical data, corresponding to the analytically
derived value $1/\psi^{\rm tri}= 8/3$. The leveling off at low temperatures
could be avoided by further fine-tuning $a^0_4$, while for higher temperature the
slope decreases as non-universal features start to play a role towards the cut-off
scale.
If the value of $a^0_4$ is increased slightly above $a_4^{0,\rm tri}$ and
$a^0_2$ again tuned to the zero temperature quantum critical point the
crossover behavior in the correlation length as derived in equation
(\ref{eq:corr_length_lin}) appears.
Results of the numerical computations demonstrating this crossover behavior of
the correlation length are displayed in Fig.~\ref{xi_crossover_d3z3}.
\begin{figure}[ht!]
\begin{center}
\includegraphics[width=3.0in]{figures_qucrit_joh/xim2vsTparaset_2.eps}
\caption{(Color online) Correlation length versus temperature in the quantum
critical regime in $d=z=3$. The plot parameters are $a_4^0 = -0.03$ slightly
above the tricritical value and $a_2^0=0.0001809109296$ tuned to the zero
temperature transition point. The vertical dashed line gives the analytical
crossover temperature as obtained from Eq. (\ref{Tcross2}).
The fits (straight lines) are discussed in the text.}
\label{xi_crossover_d3z3}
\end{center}
\end{figure}
At low temperature the slope $\epsilon = 1.33$ fits well to the quantum critical
scaling $1/\psi_{\rm qc} = 4/3$. For larger temperatures the fit gives
$\epsilon = 2.39$, which lies about 10$\%$ below the value expected
$1/\psi^{\rm tri}$.
This can be explained with deviations from the power law at higher energy
as noted in the discussion of Fig.~\ref{xi_tricrit_d3z3}.
On choosing a value of $a_4^0$ closer to the tricritical one the crossover
temperature decreases. The (low $T$) region characterized by the standard quantum
critical scaling shrinks and tricritical scaling down to $T=0$ sets in, which is
accompanied by $\epsilon$ approaching the value $8/3$ (see Fig.~3).
We now turn to the phase boundary. The result computed for the quartic
coupling $a^0_4$ fine-tuned to its tricritical value is exhibited in
Fig.~\ref{phase_boundary_d3z3}.
\begin{figure}[ht!]
\begin{center}
\includegraphics[width=3.0in]{Tr_line_a4-0.03027_d3z3.eps}
\caption{(Color online) Transition line in $d=z=3$. The quartic coupling is
chosen close to the tricritical value $a_4^0=-0.03027 \approx a_4^{0,\rm tri}$.}
\label{phase_boundary_d3z3}
\end{center}
\end{figure}
The shape of the phase boundary follows the power law $T\sim |a_2^0-a_2^{0,\rm cr}|^\psi$.
In the double logarithmic plot we find the slope $\epsilon\approx 0.378$, which compares
well to the value $\psi^{\rm tri}=3/8$ obtained analytically in Sec.~V. When the
quartic coupling is increased, the system shows the Hertz-Millis
scaling with $\psi^{\rm qc}=3/4$ for low temperature, and at larger
temperature tricritical scaling with $\psi^{\rm tri}=3/8$ is still visible for
the right choice of parameters (compare Ref.~\onlinecite{jakubczyk09}).
\section{Summary}
We have presented a study of crossover behavior occurring in the vicinity of
metallic quantum critical and tricritical points. The analysis is based on
renormalization group flow equations derived within the functional RG framework
applied to the Hertz action, retaining a $\phi^6$ term. It complements and
extends an earlier work (Ref.~\onlinecite{jakubczyk09}), providing
results in the quantum critical regime and delivering analytical insights.
We focused on crossovers occurring in the temperature dependence of the
correlation length and the shape of the phase boundary in the quantum critical
regime above the quantum critical and tricritical points in the phase diagram.
The linearized form of the flow equations could be solved analytically in $d>2$,
yielding exponents for the power-laws obeyed by the correlation length and
phase boundary.
The analytical solutions were confirmed and complemented by a numerical evaluation of the
full integro-differential flow equations.
In the quantum tricritical regime the power-law contribution to the inverse
correlation length generated by the $\phi^6$ interaction provides only a
subleading correction to the generic quadratic temperature dependence induced
by the temperature dependence of the coefficients in the bare action.
For quantum critical systems close to quantum tricriticality, we computed the
crossover temperature above which tricritical scaling is observed.
In the situation where a tricritical point occurs at $T>0$, we obtained
an expression for the tricritical temperature, and recovered the shift exponent
corresponding to the classical theory of tricriticality.
\begin{acknowledgments}
The authors would like to thank M. Nieszporski, S. Takei, H. Yamase for
valuable discussions, and S. Takei for a careful reading of the manuscript.
PJ was partially supported by the German Science Foundation through the
research group FOR 723.
\end{acknowledgments}
\begin{appendix}
\section{Explicit expressions for the constants in Section V}
In this Appendix we give explicit expressions for the constants appearing in
the solutions to the flow equations in Sec. V.
For the equations for $u(\Lambda)$ we have introduced
\begin{equation}
C=A\Lambda_0^z\;,
\end{equation}
and
\begin{equation}
C'=\Lambda_0^{2-d}\left(\frac{2\pi
Z_{\omega}}{Z}\right)^{\frac{d-2}{z}}\left(B-\frac{2\pi Z_{\omega} }{Z}A\right)\;.
\end{equation}
For $\delta(\Lambda)$ one has
\begin{equation}
D_0=\delta_0\Lambda_0^2+ A'\left(u_0+\frac12
v_0 A\Lambda_0^{z}\right)\Lambda_0^{z+2} \; ,
\end{equation}
and
\begin{eqnarray*}
D_1&=&\Lambda_0^{4-d}\left(u_0+ v_0A\Lambda_0^{z}\right)\left(\frac{2\pi
Z_\omega}{Z}\right)^{\frac{d-2}{z}}\times \\
&&
\times\left[B'-A'\left(\frac{2\pi Z_\omega}{Z}\right)\right]\;,
\end{eqnarray*}
and
\begin{eqnarray*}
D_2&=& \frac{A'A\Lambda_0^{6-2d}}{2}\left(\frac{2\pi
Z_\omega}{Z}\right)^{\frac{2(d+z)-4}{z}}+ \\
&&+B' \Lambda_0^{6-2d}\Bigg[\frac{B}{2}\left(\frac{2\pi Z_\omega}{Z}\right)^{\frac{2d-4}{z}}-
A\left(\frac{2\pi Z_\omega}{Z}\right)^{\frac{2d+z-4}{z}}\Bigg]
\; ,
\end{eqnarray*}
where
\begin{equation}
A=120v_d\frac{Z}{\pi Z_\omega}\frac{1}{(d+z-2)^2}\;,
\end{equation}
$A'=96\,A/120$
and
\begin{equation}
B=\frac{120v_d}{d(d-2)},
\qquad
B'=\frac{96v_d}{d(d-2)}
\;.
\end{equation}
\end{appendix}
|
\section{Introduction}
The study of food webs has recently attracted the interest of complex systems scientists as one of the clearest example of a network structure whose property can only be understood by looking at the system as a whole.
A food web is the collection of the predation relations in an environment and assumes naturally the form of a network that is a mathematical object composed by vertices (the biological species) and their edges (the predation relations). Such a structure is very widespread and can be recovered with similar statistical properties in a variety of other situations from WWW\cite{AB99} and the Internet\cite{CMP00} to protein interactions\cite{AJB} and social systems\cite{ASSO}. Despite this similarity, food webs represent one of the most interesting cases of study for their particular topology. For example vertices can be divided in classes thanks to their biological meaning (i.e. prey/predators). Also, the structure is naturally layerized when considering the minimum distance of species from the external resources. All these properties make these structures extremely interesting for testing models and algorithms related to complex networks. Similarly, some of the ideas developed in the area of computer science turned out to be useful in the case of food webs, signalling some biological meaning hidden in this topology\cite{Mercedes}.
Traditionally, the most important quantities in a food web are the number of vertices $N$ and the number of edges $L$. Since the maximum number of edges you may have in such a system grows as $N^2$ (precisely $N(N-1)$ for a directed graph) one also considers the density of edges $L/N^2$, a quantity known in ecology as the {\em directed connectance}.
The edges are directed and each of them follows the convention (based on the flux of nutrients) of drawing the edge from prey to predator. In this work we will consider only {\em trophic webs}, where all species which have exactly the same predators and prey are merged.
Another characterization can be obtained by considering all the species that have no predators. They are usually
indicated as the {\em top} (T) species. Similarly, the species with no prey are called {\em basal} species (B). All the others form the {\em intermediate} (I) class.
All the species are ultimately sustained by the transformation into biomass of external resources like water, mineral sunlight by means of the basal species. It is then customary to describe this situation with the introduction of an external node, called zero node, which points to all basal species i.e. to that nodes with only out-edges. Given this structure, it is easy to define layers of species given by the distance (i.e. the minimum path) towards the zero node of external resources. The distance is measured as the number of edges the biomass has to travel. As in the Internet\cite{loops} some loops can be present in the system and they account for the stability and resilience of these structures\cite{nat}.
Food webs have been modelled in many ways, and the different models have been validated with the experimental data available.
Here we focus on one recent and successful model on which we create a suitable population dynamics by means of Lotka-Volterra equations whose parameters are
self-determined by the topology created by the model itself.
\section{Niche Model}
There are many static models of food webs which reproduce the features of real ecosystems such as fractions of top, basal and intermediate species, number of food chains, average chain length, and connectance \cite{connectance}.
The simplest way is to create suitable graphs \cite{bollobas, erdosreny, newmanrand}, where (given the linkage density and the number of nodes), directed edges are assigned to randomly chosen pairs of nodes. The agreement between real and simulated food webs is not very good. This is not surprising, since this simple model has many unrealistic features such as the assumption that every species can in principle be the predator of every other species.
Recently William and Martinez introduced another static model, called ``niche model" \cite{niche}. The authors found a remarkable agreement between real webs and the synthetic ones generated by the model (much better than the one measured for cascade model \cite{cascade}). This is particularly true when considering features such as cycles and species similarities.
The external parameters of the model (i.e. the quantities fixed from the beginning) are the number of species $S$ and the directed connectance $C=L/S^2$. To every node is assigned a uniformly distributed number $n_{i}$ into the interval $[0,1]$, the niche space. A species $i$ is characterized by its niche parameter and its list of prey. Prey are chosen for all species according to the following rule: a species $i$ preys on all species $j$ with niche parameters $n_j$ inside a segment of length $r_i$ centered in a position chosen randomly inside the interval $[r_i/2,n_i]$, with $r_i=xn_i$ and $x$ a random variable with probability density function
\begin{equation}
p_x(x)=\beta(x,1,b)=b(1-x)^{(b-1)} \label{beta}
\end{equation}
Choosing $b=(1/2C)-1$ is possible to generate graphs with the desired size and connectance\footnote{In the niche model species with no prey and predators are eliminated and species with the same list of prey and predators, that is trophically identical species, are merged.}.
Niche model estimates the central tendency of empirical data remarkably well \cite{niche}. Its topological and analytical properties have been widely studied \cite{twodegree, analytic} and it has been shown that the predictions of the model are robust with respect to the specific form of the $p_{x}(x)$ chosen\cite{patterns, robustpatterns}.
\section{Population Dynamics}
Given the network structure, we want to define a population dynamics for the individuals of the species described in this food web.
To each node $i$ we associate a population i.e. a function of time $N_{i}(t)$ which represents the density of individuals of the same trophic species per unit of area.
To describe population dynamics we use the generalized Lotka-Volterra equations :
\begin{equation}
\frac{dN_i}{dt} = r_i N_i \left( \frac{K_i - \sum_{j=1}^S \alpha_{ij}N_j}{K_i} \right) \label{lveq}
\end{equation}
where $r_{i}$ is the intrinsic growth rate of species $i$, $K_{i}$ his carrying capacity and $\alpha_{ij}$ represents the effect species $j$ has on the population of species $i$. Pulling the carrying capacity into the interaction term the equations became
\begin{equation}
\frac{dN_i}{dt} = r_i N_i \left( 1 - \sum_{j=1}^S \overline{\alpha_{ij}}N_j \right) \label{lvgeneral}
\end{equation}
This doesn't actually change the equations, but only how the interaction $\overline{\alpha_{ij}}=\alpha_{ij}/K_i$ is defined. For simplicity all self-interacting terms $\overline{\alpha_{ii}}$ are set to 1.
One can represent both the populations and the growth rates as rows of numbers (vectors) and the interaction term $\alpha$ as a matrix, called also \emph{competition kernel}. Let us suppose that we have only one type of external resource $R$ produced with a constant rate $y$ (\emph{renewability} of resources) and let us also suppose that each basal species consumes a fraction $X$ at a rate $c_{i}$. The equation for the resources is:
\begin{equation}
{dR(t) \over dt} = R(y-\sum_{i=1}^Sc_iN_i) \label{res}
\end{equation}
where the first term considers the renewability and the second one gives the total rate of consume\footnote{here is assumed that $c_{i}\neq 0$ only for basal species.}.
If we consider the equilibrium conditions for equations that is \ref{lveq} and \ref{res}, $dN_i/dt=0$ and $dR/dt=0$, we find a relationship between the ecological parameters given by
\begin{equation}
y=\sum_{ij}^Sc_i\alpha_{ij}^{-1}K_j \label{constraint}
\end{equation}
This gives a fundamental constraint on all the parameters (except the $r_i$), especially on the competition kernel which must be invertible.
Some authors use as competition kernel a function of the distances between species in niche space \cite{kernel, gaussian}. The problem with such a choice is that the topology of the food web is not included in the equations; a simple way to incorporate it is to use a combination of projections weights, an issue treated in the next section.
\section{Projections Graphs}
An adjacency matrix $A$ of $S$ rows and columns represents an aggregated food web with $S$ trophic species. The elements $a_{ij}$ is taken $1$ if species $j$ predates on species $i$ and $0$ otherwise.
The ecosystem can also be represented as a bipartite graph \cite{caldaproj} where two classes of nodes are present: predators (top and intermediate species) and prey (basal and intermediate species) and each directed edge is always disposed between nodes belonging to different classes. Such a graph can be projected into the predators-network, where two predators are connected with an edge weighted proportionally to the numbers of prey they have in common and, correspondingly, into the prey-network, where two prey are connected according to the number of predators they share.
The projection graphs are two undirected, weighted graphs whose sizes are the number of predators and the number of prey in the food web respectively.
The corresponding adjacency matrices $A^{pred}$ and $A^{prey}$ are symmetric and we define their elements in the following way:
\begin{eqnarray}
a_{ij}^{pred}=\frac{\sum_{k=1}^S\delta_{a_{ki},1}\delta_{a_{kj},1}}{S(B+I)} \label{pred}\\
a_{ij}^{prey}=\frac{\sum_{k=1}^S\delta_{a_{ik},1}\delta_{a_{jk},1}}{S(T+I)} \label{prey}
\end{eqnarray}
We choose to normalize the predators weights over all possible prey, and the prey weights over all possible predators. Note that $a_{ij}^{pred}=a_{ji}^{pred}$ and $a_{ij}^{prey}=a_{ji}^{prey}$.
Since the change in the size of the graphs (without loss of generality) we relabel the vertices in the two graphs and keep the information about the species they represent \cite{cohen}.
The weights of the projections' graphs are a measure of the interspecific competition for resources, giving information on how two species compete (or are object of competition) in the predation interaction. A more significant meaning of this quantity should be derived by the analysis of an original weighted food web where the strength of the predation is also considered. For this reason, considering equation \ref{lveq} we propose the following competition kernel that joins in a self-organized way\cite{self} the topology with the dynamics:
\begin{equation}
\alpha_{ij}= a_{ij}^{pred}-a_{ij}^{prey} \label{projkernel}
\end{equation}
This means that the influence of population $i$ on population $j$ is
negative if species $i$ and $j$ share some prey and positive if they share some predators. The competition between two species increases with respect to
the number of prey species they share and vice versa.
Using the elements defined in eq. \ref{projkernel},
it is possible to simulate a population dynamics on both model and empirical food webs.
\section{Results}
The stability of empirical and model food webs has been tested
numerically following this steps:
\begin{enumerate}
\item \emph{Food web adjacency matrix}. We generate it using niche model, where the input parameters are connectance $C$ and size $S$. In real cases this is done considering empirical food webs data.
\item \emph{Competition kernel}.
We derive the competition kernel $\alpha_{ij}(C)$ from equation \ref{projkernel}.
\item \emph{Ecological parameters}. We fix the parameters $r_i$, $c_i$ and $K_i$:
\begin{itemize}
\item As a first approximation the intrinsic growths $r_i$ are set all equals for all species. This means that in the ideal condition of no competition and infinite resources, all the populations should grow at the same rate. Varying these parameters one can simulate the different lifetimes of species and their reproduction strategies.
\item Setting the carrying capacities $K_i$ of the species means that, in the ideal condition of no interspecific competition ($\alpha_{ij}=0$), the maximum number of individuals per unit of space which are sustainable from the external environment is fixed for every species.
\item the consuming rates $c_i$ has been set to 0.1 only for basal species. This is quite realistic because basal species are, by definition, the species which directly feed on the external environment.
\item At this point using equation \ref{constraint} we fix the renewability of resources $y$ which depends basicly on the topology of the graph. To avoid indefinite growth, for simulated food webs, we fix the maximum renewability $y_M=10$ and generate graphs with the same connectance and size until we obtain the desired renewability ($0<y\leq y_M$).
\end{itemize}
\item \emph{Integration of equations}. Once we have the competition kernel and the desired $y$ we solve the equations \ref{lveq} using the 4th order Runge-Kutta algorithm
\item \emph{Stability}. The results have been collected in the steady state.
\end{enumerate}
\subsection{Projections Topological Properties}
We measured topological properties of projections graphs for both empirical
and model food webs.
The data we used have been selected to be the largest and highest-quality empirical trophic food webs present in literature. They represent a wide range of ecosystems, from freshwater habitat (Skipwith Pond SWP, Little Rock Lake LRL, Bridge Brooke Lake BBL) to freshwater-marine interface (Chesapeake Bay CPB, Ythan Estuary YE) to terrestrial habitats (Coachella Valley CDE, Saint Martin Island SMI)
The measure of all topological features (except for averaged weights)
is made by considering the binary matrices of projections where all
weights different from 0 are put to 1.
Looking at the empirical graphs we find a very symmetric topology (Fig.\ref{skipwithpred}, \ref{skipwithprey}).
Some projections show the formation of isolated communities both in predators and in prey graphs (Fig. \ref{chesapeakepred}, \ref{chesapeakeprey}) and sometimes we find only isolated nodes i.e. specialists in predation or in being a prey. This possibility of community formation is not considered in directed food webs. In fact, in both empirical and model graphs, there are no isolated nodes or clusters (in niche model isolated species are removed).
We can measure this feature by considering the matrices of path lengths whose elements are given by
\begin{equation}
d_{ij}^{pred/prey}= min\{\sum_{k,l\in P_{ij}}a_{kl}^{pred/prey}\} \label{paths}
\end{equation}
where $P_{ij}$ is a path connecting node $i$ and $j$. Putting $d_{ij}=0$ when the distance between $i$ and $j$ is infinite i.e. when the nodes are in different clusters. We compute average path length using two different normalizations:
\begin{equation}
l_G=\frac{1}{n(n-1)/2}\sum_{i>j}d_{ij}\qquad l_R=\frac{1}{[n(n-1)/2]-(l_0/2)}\sum_{i>j}d_{ij} \label{averagelgt}
\end{equation}
where $l_0$ is the number of zeros in the matrix of path lengths. With this definitions $l_G$ is the average path length over all possible paths and $l_R$ is the average path length over all real paths. We have $0\leq l_G/l_R \leq 1$ being 0 when all nodes are isolated and 1 when all nodes are in the same cluster. Together with this we measure other topological quantities such as clustering, diameter and average weight, for both projections (Table \ref{tableproj}).
One of the first evidences is that the formation of communities is always common to both projections. Connectance varies from $0.24$ of CDE to $0.92$ of SWP, indicating that projections graphs are strongly connected in comparison to the original food webs. Average weights, the only quantities measured considering projections as weighted graphs, vary from $0.04$ to $0.21$ for predators and from $0.05$ to $0.22$ for prey. It seems that $\langle w \rangle$ is independent from original connectance and takes high values compared to model projections. Furthermore we can say that projections of food webs present all the characteristics of small world networks i.e. small diameter and large clustering \cite{smallworld}.
We then compute the same quantities for projections of graphs derived from niche model and random graphs in function of the mean degree of original directed graphs (Fig. \ref{4graph}). We expect averaged topological properties of random projections to be the same for predators' and prey's graphs because in and out degree distributions of random digraphs have the same form \cite{newmanrand}. Comparing the curves with empirical data we find good agreement for clustering and average path lengths. The model represents the formation of competitive communities better than random projections while it overestimates diameters and underestimates average weights. We see how empirical $D$ is better described through random projections, and how $\langle w \rangle$ of empirical projections is larger than model and random averaged weights. We can explain this trend considering the feeding rule of niche model which assigns prey from a single portion of niche space. This reduces the probability of sharing different resources to a single, well distributed interval, augmenting the diameter and reducing weights of projections graphs. On the other side, empirical food webs are not strictly interval and do exhibit a strong bias towards contiguity of prey \cite{intervality}. This result suggest that empirical observed niches, once mapped onto a single dimension, should be composed of various intervals along niche space.
\subsection{Dynamical properties}
We tested the stability of these Lotka-Volterra systems using both empirical and model food webs. We remarked that, among all possible behaviors, using coefficients \ref{projkernel} as competition kernel, the system reaches quite always the steady state for small $S$ and $C$. Otherwise the system is no longer stable and populations sizes go to infinity when complexity, measured as $SC$, grows. We tested steady states changing initial conditions, and ecological parameters.
The only stable empirical food web is Chesapeake Bay (CPB) which is the one with the lowest complexity. In Fig \ref{dyn1} we present population dynamics obtained setting initial population's densities to 1 for all species and changing the intrinsic growth of the species. We remarked that the $r_i$ are a measure of the speed of convergence of populations sizes. Augmenting the intrinsic growths the steady state is reached faster. Populations' sizes at the steady state should be equal to the carrying capacities, but the presence of competition kernel changes the real $K_i$ of every species. We remarked, however, that this parameters give the size's order of stable populations and when $K_i=K_P$ mean population is always $K_p$. We obtain the same result changing the initial conditions.
To conclude our analysis we simulated population dynamics on niche model. We fixed S and C and generated 100 realizations of population dynamics on model graphs for a relevant number of points. We counted how many times the system reaches the steady state and how many times it diverges. Again we find that, using expression \ref{projkernel} as competition kernel, these are the only two observed behaviors. The probability of reaching a steady state decreases with complexity (Fig. \ref{stable1} and \ref{stable2}). Population dynamics on graphs from niche model is stable only for small S and C which is no longer true for empirical food webs. Simulating dynamics on graphs with more than 50 nodes is a hard computational task. However it is evident that the model is stable for $SC<2$ and that stability decreases linearly between $SC=2$ and $SC=5$. This results are robust under the change of initial conditions and ecological parameters.
With this we show that it is possible to avoid chaotic behaviors of Lotka-Volterra systems using weights of projections' graphs as competition kernel. When a fixed food web topology gives raise to stable populations, its complexity must be small. This confirms recent findings of theoretical ecology that complex food webs are not necessarily stable.
|
\section{Introduction}
Since powerful radio sources reside in massive ellipticals (Best et
al. 1998), finding radio galaxies at high redshifts is important to
understand the radio evolution of galaxies. It had been
noticed in the early 80's that the fraction of radio sources that can
be optically identified is lower by a factor of 3 or more for steep
spectrum radio sources ($\alpha > 1; S_\nu \propto \nu^{-\alpha}$;
Tielens, Miley \& Wills, 1979, Blumenthal \& Miley, 1979; Gopal-Krishna \& Steppe, 1981),
suggesting that the radio sources with steeper spectra at decimeter wavelengths
are more distant compared to the ones with normal spectra ($ 1 > \alpha >
0.5$). Over the years, this correlation has been exploited to search
for radio sources at high redshifts (Rottgering et al., 1994; De Breuck et al. 2000,
2002; Klamer et al. 2006). Since the radio emission does not suffer from dust absorption,
selecting candidate high redshift radio galaxies (HzRGs) at radio
frequencies provides an optically unbiased sample. The most distant
radio galaxy at z = 5.19 was discovered using the spectral index -
redshift correlation (van Breugel et al. 1999). Until today, about 45
radio galaxies are known beyond redshift of 3, and nearly all of
them were discovered through the radio spectral index - redshift
correlation. Thus, this is the most efficient method to find HzRGs.
In addition to constraining models of formation and evolution of
massive galaxies, HzRGs can also be used to study the environment
at those epochs. Deep radio polarimetric observations of several HzRGs
have shown large rotation measures (RM) of the order of thousands of
rad m$^{-2}$ (Pentericci et al. 2000; Carilli et al. 1997; Athreya et
al. 1998; Kronberg et al. 2008). Pentericci et al. (2000) also found that the fraction of
radio galaxies with extreme Faraday rotation is increasing with
redshift which is consistent with the hypothesis that the environment
is denser at high redshift. Since at moderate redshifts large RMs are known to occur
in radio
galaxies residing near the center of clusters of galaxies,
the HzRGs are an excellent tool to study the (proto) cluster
environment at high redshifts. It is therefore important to identify
and study as many HzRGs as possible.
A major programme to search for HzRGs using the radio spectral index -
redshift correlation was carried out in the last decade using
different radio surveys at 1400 MHz and lower frequencies (eg:
Rottgering et al. 1994, De Breuck et al. 2000, 2002, 2004, 2006; Klamer et al. 2006; see also review
by Miley and De
Breuck, 2008). The searches were mostly limited by the sensitivity
limit of these shallow, wide-area radio surveys. This bias has led to the
detection of only the brightest HzRGs which are at the top end of
the radio luminosity function (See section 2), which is about three
orders of magnitude brighter than the luminosities at the lower end of
FR-II population (Fanaroff \& Riley, 1974). Since steep spectrum
radio sources are preferentially selected at low radio frequencies,
deeper observations at these frequencies are needed to discover HzRGs
that are 10 to 100 times less luminous than most of the known HzRGs, and these
could be the high-redshift counterparts of more typical FR-II radio
galaxies. The 150 MHz band of GMRT (Giant Metrewave Radio Telescope,
India, http://www.ncra.tifr.res.in) with its large field of view
(half-power beam width of 3 degrees), high angular resolution ($\sim
20^{''}$) and better sensitivity ($\sim$ 1 mJy from a full synthesis
observation) is well suited for this kind of work.
For example, a steep spectrum radio source ($\alpha > 1$)
at the completeness limit of WENSS at
327 MHz (Rengelink et al. 1997), will have a flux density $>$ 65 mJy
at 150 MHz. Our experience with GMRT at 150 MHz shows that it is
possible to reliably detect sources more than 10 times fainter than this
value.
The large field of view of GMRT also enables us to detect close to
thousand radio sources in a single pointing.
Encouraged by this, we have started a programme to
observe several carefully chosen deep fields at 150 MHz with GMRT with
an aim to detect steep spectrum radio sources to flux density levels much
fainter than that of known high-redshift radio sources (see Section 2).
As the sample at low radio frequencies goes deeper, the
available surveys at higher radio frequencies such as FIRST
won't be deep enough to get the spectral index estimates, particularly for
steep spectrum sources. For example,
from the present work, it is seen that if we use only the FIRST catalogue,
37\% of the sources below 18 mJy at 150 MHz (which corresponds to 1 mJy at
1.4 GHz for a spectral index of 1.3) do not have counter parts in FIRST
as against 9\% for stronger sources. This limitation can be addressed
by deeper surveys of this region at higher radio frequencies.
Therefore, we have chosen the well known deep fields because significant amount of
data already exist for them at higher radio frequencies and in the optical and
infrared bands. These deep observations will allow us to estimate the
spectral index for faint sources which are below the detection limit
of NVSS or FIRST. The deep optical data will help to estimate the redshifts
and eliminate nearby and known objects among the steep spectrum sources.
Wherever needed, additional deep observations with the
GMRT at 610 MHz will be obtained for a better estimate of radio
spectra. In addition to the steep spectrum radio sources, this kind of
survey will also help us to understand the evolution (LogN-LogS) and
spectral index properties of faint radio sources, to detect low-power
FRI sources, relic emission and large angular size radio sources.
In this paper we present deep 150 MHz radio observations of the
LBDS-Lynx field with the GMRT with the primary aim of detecting steep
spectrum radio sources which are candidate HzRGs. The Leiden-Berkeley
Deep Survey (LBDS) in the Lynx area (Windhorst et al. 1984) was
carried out in the mid eighties to better understand the nature of
faint radio sources and their cosmological evolution. Multi-band
optical data were obtained with the Kitt Peak Mayall 4m telescope and
subsequently deep radio observations were carried out at 327 MHz and
1412 MHz with WSRT, complimented by 1400 MHz and 4860 MHz imaging with the VLA
(Windhorst et al 1984, 1985, Oort, 1987, Oort et al. 1988). These
radio observations showed a moderate increase in the number of radio
sources below $\sim$ 5 mJy at 1.4 GHz, which were predominantly
identified with blue galaxies, a population responsible for the upturn in the
source counts at faint flux density levels (Windhorst et al. 1985). The
spectral index studies of radio sources selected at 327 MHz in this region using the 327 MHz and
1400 MHz data showed that the median spectral index was flatter for
fainter sources selected at 327 MHz (Oort et al. 1988). The sources
with the steepest spectra ($\alpha > 1$) were mostly unidentified in
the optical 4m telescope images, once again indicating that they were
likely to be very distant.
The sources we detected at 150 MHz with the GMRT were compared with
the deep observations of this region at 327, 610 and 1412 MHz with
the WSRT and at 4860 MHz with the VLA. We have also used the available
catalogues such as WENSS at 327 MHz and NVSS and FIRST at 1400 MHz to
estimate spectral indices.
We identify about 150 steep spectrum radio
sources from this field. About two-third of them are not detected to
the limit of SDSS, hence are strong candidate HzRGs. In Section 2,
we present supporting arguments for reviving the search for HzRGs using
the steep spectrum technique. Observations, data analysis and the
data obtained are presented in Section 3. Results and discussions are
presented in Section 4 and concluding remarks in Section 5.
\section{Re-visit of Ultra-steep spectrum technique for finding
HzRGs}
The method to find HzRGs using the correlation between steep radio spectrum
and high redshift of the radio galaxy is the most efficient technique
to discover HzRGs (e.g: Pedani, 2003; de Breuck et al. 2000). Till date
about 45 radio galaxies have been discovered
with z $>$ 3 using this technique. Miley \& de Breuck (2008) in their
review have provided a table of all known HzRGs with redshift more than 2.
We have independently surveyed the literature for an update on all HzRGs
with redshift more than 3, which is provided in Table 1. A total of 47 galaxies are listed in the table,
which is arranged as follows: Column 1: Source Name;
Column 2: redshift; Column 3: radio spectral index (in a few cases where
the spectral index was not known, we have assumed a value of 1); Column 4: observed
flux density at 1.4 GHz; Column 5: estimated flux density at 150 MHz using the spectral
index and the 1.4 GHz flux density, given in Columns 3 and 4; Column 6: Reference.
\begin{table}
\caption{Known HzRGs till date, discovered mostly using ultra steep spectrum technique.}
{\footnotesize
\begin{tabular}{l l l l l l }
Source Name & z & $\alpha$ &S$_{1400}$&S$_{150}$& Ref\\
& & & mJy & mJy & \\
& & & & & \\
B30032+412 &3.658 &1.16 & 102 & 1330 & Jar01\\
NVSSJ012142+132058&3.516 &1.37 & 55.4 & 1150 & deB01\\
B20140+32 &4.413 &1.17 & 91.5 & 1220 & Raw91\\
NVSSJ020510+224250&3.5061&1.37 & 58.4 & 1212 & deB00\\
NVSSJ021308-322338&3.976 &0.90 & 30.4 & 223 & deB06\\
NVSSJ023111+360027&3.079 &1.30 & 44.3 & 788 & deB00\\
PMNJ0253-2708 &3.16 &1.03 & 273.3& 2673 & McC96\\
RCJ0311+0507 &4.514 &1.31 & 473.0 & 8597 & Kop06\\
PKS0315-257 &3.1307&1.05 & 424.0 & 4334 & McC90\\
NVSSJ033953-232136&3.49 &1.00$^*$ & 36.5 & 334 & Cac00\\
NVSSJ034642+303949&3.72 &1.42 & 34.9 & 809 & deB02\\
4C+60.07 &3.791 &1.45 & 156.8 & 3885 & Cha96\\
WNJ0617+5012 &3.153 &1.37 & 26.5 & 550 & deB00\\
4C+41.17 &3.792 &1.51 & 235 & 6650 & Cha96\\
WNJ0747+3654 &2.992 &1.41 & 36.8 & 835 & deB00\\
NVSSJ075806+501104&2.996 &1.07 & 98.0 & 1047 & Cru06\\
NVSSJ083609+543325&3.341 &1.02 & 62.7 & 600 & Cru07\\
B20902+34 &3.382 &0.84 & 313 & 2010 & Lil88\\
TNJ0924-2201 &5.19 &1.63 & 73.3 & 2705 & van99\\
NVSSJ094724-210505&3.377 &1.00$^*$ & 10.9 & 100 & Bro06\\
NVSSJ095438-210425&3.431 &1.00$^*$ & 51.3 & 469 & Bro06\\
NVSSJ095751-213321&3.126 &1.00$^*$ & 62.3 & 570 & Bro06\\
NVSSJ104906-125819&3.697 &1.51 & 108.1 & 3059 & Bor07\\
6CB105034+544035 &3.083 &1.34 & 66.4 & 1290 & deB00\\
NVSSJ105917-303658&3.263 &1.06 & 64.7 & 676 & Bry09\\
NVSSJ111223-294807&3.09 &1.40 & 101.0 & 2240 & deB00\\
B21121+31B &3.2174&1.46 & 74.0 & 1875 & deB00\\
TNJ1123-2154 &4.109 &1.57 & 49.6 & 1603 & deB00\\
4C+39.37 &3.22 &1.47 & 255.3 & 6612 & Raw90\\
4C+03.24 &3.5699&1.28 & 256 & 4354 & Rot97\\
8C1435+635 &4.261 &1.31 & 451 & 8197 & Lac94\\
TNJ1338-1942 &4.11 &1.31 & 122.9 & 2234 & deB00\\
J163912+405236 &4.88 &0.75 & 21.8 & 114.7 & Jar09\\
WN B1702+6042 &3.223 &1.24 & 53.2 & 828 & Vil99\\
7C1814+670 &4.05 &1.09 & 236.1 & 2637 & Lac99\\
4C+72.26 &3.536 &1.22 & 259 & 3857 & deB01\\
TXS1911+636 &3.59 &1.42 & 23.3 & 540 & deB01\\
MRC2005-134 &3.837 &1.42 & 115.4 & 2676 & deB01\\
TNJ2009-3040 &3.158 &1.36 & 65.3 & 1326 & Bor07\\
4C+18.61 &3.056 &0.87 & 575.3 & 3948 & Ste99\\
4C+19.71 &3.594 &1.02 & 306 & 2927 & Ste99\\
NVSSJ230123-364656&3.22 &1.38 & 20.3 & 431 & deB06\\
WNJ2313+4053 &2.99 &1.50 & 11.3 & 313 & deB00\\
NVSSJ231402-372925&3.45 &1.22 & 129.9 & 1934 & deB06\\
NVSSJ231727-352606&3.874 &1.19 & 59.2 & 825 & deB06\\
NVSSJ232100-360223&3.32 &1.65 & 15.1 & 583 & deB06\\
B3J2330+3927 &3.087 &1.21 & 148.4 & 2162 & Per05\\
\end{tabular}
}
\\
$^*$The radio spectral index measurements were not available; so a spectral
index of 1 was assumed. \\ References:
{\sf Jar01:} Jarvis et al. 2001;
{\sf deB01:} de Breuck et al. 2001;
{\sf Raw91:} Rawlings et al. 1991;
{\sf deB00:} de Breuck et al. 2000;
{\sf deB06:} de Breuck et al. 2006;
{\sf McC96:} McCarthy et al. 1996;
{\sf Kop06:} Kopylov et al. 2006;
{\sf McC90:} McCarthy et al. 1990;
{\sf Cac00:} Caccianiga et al, 2000;
{\sf deB02:} de Breuck et al. 2002;
{\sf Cha96:} Chambers et al. 1996;
{\sf Cru06:} Cruz et al. 2006;
{\sf Cru07:} Cruz et al. 2007;
{\sf Lil88:} Lilly, 1988;
{\sf van99:} van Breugel et al. 1999;
{\sf Bro06:} Brookes et al. 2006;
{\sf Bor07:} Bornancini et al. 2007;
{\sf Bry09:} Bryant et al. 2009;
{\sf Raw90:} Rawlings et al. 1990;
{\sf Rot97:} Rottgering et al. 1997;
{\sf Lac94:} Lacy et al. 1994;
{\sf Jar09:} Jarvis et al. 2009;
{\sf Vil99:} Vilani et al. 1999;
{\sf Lac99:} Lacy et al. 1999;
{\sf Ste99:} Stern et al. 1999;
{\sf deB06:} de Breuck et al. 2006;
{\sf Per05:} Perez-Torres et al. 2005.
\end{table}
In order to understand whether the known HzRGs represent typical FRII
radio sources at high-redshifts, or the highest luminosity sources in
that category, we have computed the expected flux density at 150 MHz
corresponding to the FRI/FRII break luminosity (assuming standard
cosmological parameters of $\Omega_m$ of 0.27, $\Omega_{vac}$ of 0.73
and H$_o$ of 71 kms$^{-1}$Mpc$^{-1}$; Bennett et al. 2003). We assume here that the FRI/FRII
break luminosity does not evolve significantly with redshift. At
a redshift of 3, the FRI/FRII break luminosity corresponds to a flux
density of $\sim$ 4 mJy at 150 MHz (taking a spectral index of 1 for
the K-correction). The corresponding values at redshift of 4 and 5 will be
$\sim$ 1.9 and $\sim$ 1.1 mJy, respectively.
\begin{figure}
\includegraphics[width=59mm,angle=270]{figs/fig1a.eps}
\caption{The 150 MHz flux densities of known HzRGs, extrapolated using
the available spectral index and flux density measurements.
The dotted line at the bottom indicates
the observed 150 MHz flux density corresponding to the rest-frame FRI/FRII break
luminosity. The dashed horizontal line is the GMRT detection limit from the
present 150 MHz observation. It is clear from this figure that a large number
'normal population' of FRIIs, that are 10 to 100 times less luminous than the known HzRGs are
yet to be discovered. }
\end{figure}
In Figure 1, we plot the observed flux densities at 150 MHz of all known
HzRGs, computed using the available flux density measurements and the spectral
index. The dotted line corresponds to the 150 MHz flux density at the
FRI/FRII break luminosity computed for different redshifts, assuming $\Lambda$
of 0.73 and $\Omega_M$ of 0.27. It is clear from the figure that
nearly all of the known HzRGs are two to three orders of magnitude more
luminous than the FRI/FRII break luminosity. This is largely due to
selection effects because nearly all the searches for steep spectrum
radio sources use sky surveys such as WENSS and SUMSS (Mauch et al. 2003)
at low radio
frequencies and NVSS at higher radio frequencies. This selection is
biased towards stronger radio sources. The median flux density at 1400 MHz of the
known HzRGs is $\sim$ 70 mJy and the corresponding flux density at 150 MHz is
$\sim$ 1.3 Jy (90\% of them are stronger than $\sim$ 0.5 Jy), whereas
the FRI/FRII break luminosity is more than two orders of magnitude
fainter.
The above argument implies that the known HzRGs represent the tip of
the ice-berg in flux density. There are, potentially, a large number
of HzRGs yet to be discovered which are expected to be 10 to 100 times less luminous
than the known HzRGs (see Fig. 1). Although initially it was believed
that there is a 'cut-off' in the radio luminosity function at
high-redshift, subsequent to the discovery of several radio galaxies
at z $>$ 4, this evidence became suspect (Jarvis et al 2001). The radio luminosity
function (RLF) of
HzRGs indicate that at luminosities 10 to 100 times lower than the
known HzRGs, we could expect at least a 10 fold increase in space density
of the radio sources beyond the redshift of 3 (Waddington et al. 2001).
Systematic efforts are needed to find this population of radio
galaxies, which are not at the brightest end of the radio luminosity
function. As mentioned earlier, the most efficient method to find
high-redshift radio galaxies is through the steep spectrum selection.
This large gap, can thus be filled by searching for steep spectrum sources
using deep radio observations at low frequencies. The present
observation of the LBDS field at 150 MHz with the GMRT has an rms noise of
$\sim$ 0.7 mJy/beam, and the source catalog has about 750 sources
having a flux density of $\gtrsim$ 4 mJy. The sky coverage is
about 7 degree$^2$ for half-power beamwidth and
15 degree$^2$ to the 20\% of the peak primary
beam response in a given pointing. A source at the threshold of detection in the FIRST survey
will have a flux density of $>$ 10 mJy at 150 MHz for a spectral index
steeper than 1. Compared to the 150 MHz flux densities of known
HzRGs, this value is close to two orders of magnitude fainter. In the
present case of LBDS-Lynx field, deep observations exist at 327,
610 and 1400 MHz, which allowed us to obtain reliable spectral index
estimates for sources as faint as 7 mJy at 150 MHz.
Using our observations, we have obtained about 100 radio sources with
spectral index steeper than 1 having no optical counterparts in the
SDSS. The median 150 MHz flux density for these sources is $\sim$ 110 mJy,
which is an order of magnitude fainter than the median flux density at 150 MHz
for known HzRGs.
\begin{figure}
\includegraphics[width=83mm,angle=0]{figs/lbds-fov.eps}
\caption{A portion of the 150 MHz image of LBDS field obtained with the GMRT.
The rms noise in the image is $\sim$ 0.7 mJy/beam and the resolution is
$\sim 19^{''} \times 15^{''}$ with position angle of 27$^\circ$. }
\end{figure}
\section{Observation and data analysis}
\subsection{GMRT 150 MHz observations}
The 150 MHz observations of the LBDS-Lynx field centered at
RA=08h41m46s and DEC=+44d46$'$50${''}$ (J2000) were carried out with the Giant
Metrewave Radio Telescope (GMRT) on December 11, 2006 using a
bandwidth of 16 MHz, between local midnight to morning in order to reduce the
radio frequency interference (RFI). The GMRT consists of 30
antennas, each of 45 meter diameter located 90 km from Pune, India and
operates at five frequency bands from 150 MHz to 1450 MHz (Swarup et
al. 1991). For determining the flux density scale, standard flux
calibrators (3C147 and 3C286) were observed and their flux densities at
150 MHz was estimated using the Perley-Taylor scale (Perley \& Taylor 1991). From the estimate
of median spectral index of the sources in the field (0.78), which was
close to the expected value (e.g: Oort, Steemers \& Windhorst, 1988),
we infer that the error in the flux density scale
is $\lesssim$ 10\% at 150 MHz. The data were recorded in the spectral
line mode with 256 channels of 62.5 kHz bandwidth per channel at 150
MHz to minimise the effect of bandwidth smearing.
A phase calibrator was observed for five minutes for every 30
minutes of on source observation. The data were analysed using {\tt
AIPS++} (v1.9;build \#1556; see Sirothia 2008, Sirothia et al. 2009).
While calibrating the data, bad data
points were flagged at various stages. The data for antennas with
relatively large errors in antenna-based gain solutions were examined
and flagged over certain time ranges. Some baselines were flagged,
based on closure errors on the bandpass calibrator. Channel and
time-based flagging of the data points corrupted by radio frequency
interference (RFI) was done by applying a median filter with a
$6\sigma$ threshold. Residual errors above $5\sigma$ were also
flagged after a few rounds of imaging and self-calibration.
The final image was made after several rounds of phase self-calibration,
and one round of amplitude-and-phase self-calibration, where the data were normalized
by the median gain found for the entire data. The primary
beam correction was applied to the final image. The final rms
noise achieved at 150 MHz near the center of the field in
a source free region
was $\sim$ 0.7 mJy/beam with a resolution of $\sim 19^{''} \times 15^{''}$ at position
angle of 27$^\circ$. A portion of the image is presented in Figure 2.
\begin{figure}
\includegraphics[width=50mm,height=85mm,angle=270]{figs/area.eps}
\caption{Coverage of different existing deep observations of LBDS field
at 327, 610, 1412 and 1465 MHz with respect to the GMRT field of view.
One set of observations at 1412 MHz cover a much smaller area but are
deeper compared to another set of observation at 1412 MHz.
WENSS, NVSS and FIRST cover the full area of our 150 MHz coverage. They are not indicated
in this figure (Also see Table 2)}
\end{figure}
\begin{table}
\caption{Summary of radio observations and number of radio sources
available in the LBDS-Lynx area, within
20\% of the primary beam of GMRT at 150 MHz (Figure 3). }
\begin{tabular}{lllll}
& & & & \\
Freq. & Telescope & 1 $\sigma$ noise & No. of & Ref. \\
(MHz) & & (mJy) & Sources& \\
& & & & \\
153 & GMRT & 0.7 & 765 & This paper \\
327 & WSRT & 0.9 & 394 & Oort88 \\
327 & WSRT-WENSS & 4 & 302 & Reng97\\
610 & WSRT & 0.3 & 437 & WindPvt \\
1400 & VLA-FIRST & 0.2 & 1327 & Beck95\\
1400 & VLA-NVSS & 0.7 & 789 & Cond98\\
1412 & WSRT & 0.2 & 231 & Wind84\\
1412 & WSRT & 0.02 & 349 & Oort87\\
1462 & VLA & 0.045 & 139 & Wind85\\
4860 & VLA & 0.015 & 59 & Donn87\\
\end{tabular}
References: ~ Oort88: Oort et al. 1988, Reng97: Rengelink et al. 1997,
WindPvt: Windhorst, unpublished data, Beck95: Becker et al. 1995,
Cond98: Condon et al. 1998, Wind84: Windhorst et al. 1984,
Oort87: Oort 1987, Wind85: Windhorst et al. 1985, Donn87: Donnelly et al. 1987.
\end{table}
\subsection{Ancillary radio data}
Deep radio observations of this field are available at 327, 610 and
1412 MHz using the WSRT and at 1400 and 4860 MHz with the VLA (See
Table 2). The 1412 MHz observations with the WSRT 3 km array have the
east-west beam of $\sim ~12.5^{''}$ and rms noise of 0.12 to 0.2
mJy/beam (Windhorst, Heerde \& Katgert, 1984) from a 12 hour
observation. The resolution is comparable to that of GMRT at 150 MHz.
The second set of observations cover a smaller area of $\sim$ 0.7
degree$^2$, but reach a deeper flux density limit of 0.3 mJy (Oort and
Windhorst, 1985). A third set of deep observations at 1412 MHz of the
region has detected 349 sources down to 0.1 mJy (Oort, 1987). A total of 580
sources are found in the combined 1412 MHz catalogue. Deep observations
with the VLA at 1462 MHz were also carried out which yielded a little over
100 sources above a flux density limit of 0.23 mJy (Windhorst et
al. 1985). The resolution of this VLA image was $\sim$ 15$^{''}$,
comparable to the WSRT resolution at 1412 MHz. The LBDS region was also mapped at
327 MHz using the WSRT with a resolution of about an arc-min. Above a 5
$\sigma$ limit of 4.5 mJy, a total of 384 sources were found (Oort,
Seemers \& Windhorst, 1988). The 610 MHz catalog (unpublished) has
$\sim$ 400 sources stronger than 1.3 mJy. The VLA and WSRT
observations at 5 GHz were also used, though they cover a small
part of the area. Table 2 gives the summary of all available data, and
Figure 3 shows the areal overlap with the GMRT
observations. However, all these observations combined cover less than
50\% of the GMRT 150 MHz area. We also use the WENSS data at 327 MHz and NVSS
and FIRST data at 1400 MHz for the entire field.
\section{Results and discussion}
\subsection{The 150 MHz source catalog}
A catalogue of sources out to the 20\% of the peak primary beam response
was created with the peak source brightness greater than 6
times the local rms noise value. The details of the source extraction
criterion are given in Sirothia et al. (2009). The final catalogue
contains about 750 sources above a flux density limit of $\sim$ 4 mJy. Since the
rms is not uniform across the image (rms is higher away from the phase
center and also near a few bright sources), the actual completeness
limit of this catalog is higher than this limiting value. The median flux
density of the catalogue is $\sim$ 26 mJy. The majority (about two-third)
of the sources are unresolved within the resolution of the
observation. About 18\% of the sources are significantly extended
(larger than about 0.5 arcmin). Close inspection of these sources
suggests a range of morphology, such as FR-II, FR-I, one sided jets, relic
types, etc. Further details on these sources and their properties is
beyond the scope of this paper. The mean position offset at 150 MHz with
respect to the FIRST position is about -0.019$\pm$3.27 arc-sec in RA
and -1.61$\pm$3.12 arc-sec in DEC. A portion of the table is
presented in Table 3 (the full catalogue is available online).
Column 1: Source Name; Columns 2-4: RA; Columns 5-7: DEC; Columns
8-17: The flux density at 150 MHz(GMRT), 327 MHz(WSRT), 327 MHz(WENSS), 610 MHz(WSRT),
1412 MHz(WSRT, Windhorst et al. 1984), 1412 MHz(WSRT, Oort et al. 1987), 1400 MHz(NVSS), 1400 MHz(FIRST), 1462 MHz(VLA) and 4860 MHz(VLA) respectively;
Column 18: Spectral index from the fit.
The source count
reported from this field goes an order of magnitude deeper than earlier
150 MHz surveys (McGilchrist et al. 1990). In Fig. 4 we plot the GMRT
differential source count and for comparisons with previous surveys
we also plot the source count from 7C survey for the same region (Hales
et al. 2007 and references therein). We believe that our 150 MHz sources
are dominated by a cosmological population of AGNs.
\begin{figure}
\includegraphics[width=58mm,angle=270]{figs/logn-logs.eps}
\caption{LogN-LogS of the 150 MHz sources. We present sources extending
to an order of magnitude deeper than previous surveys. At higher
flux densities we are consistent with the
7C survey at this frequency (Hales et al. 2007). The source counts for 7C survey are shown for the
same region as that covered by the GMRT.}
\end{figure}
\begin{table*}
\caption{Sample table of GMRT 150 MHz sources (The complete table is available in online version).
The spectral index is from the
blind straightline fit, which may not be valid for a small fraction ($\sim$ 3\%) of sources.}
\scriptsize
\begin{tabular}{@{\extracolsep{-5pt}}llllllllllllllllll}
NAME & \multicolumn{3}{c}{RA (J2000)} & \multicolumn{3}{c}{DEC (J2000)} & S$_{150}$ & S$_{327}$ & S$_{WENSS}$ & S$_{610}$ & S$_{1412^a}$ & S$_{1412^b}$ & S$_{NVSS}$ & S$_{FIRST}$ & S$_{1462}$ & S$_{4860}$ & $\alpha$\\
& hh & mm & ss.s & dd & mm & ss.s & mJy & mJy & mJy &mJy & mJy & mJy & mJy & mJy & mJy & mJy & \\
(1) & (2)& (3)& (4) &(5) &(6) &(7) & (8) & (9) & (10) & (11) & (12)& (13) & (14) & (15) & (16) & (17) & (18) \\
& & & & & & & & & & & & & & & & & \\
GMRTJ083035+454329 & 08 & 30 & 35.7 & 45 & 43 & 29.6 & 11342.3 & $-$ & 3049.0 & $-$ & $-$ & $-$ & 614.4 & 608.3 & $-$ & $-$ & 1.48 \\
GMRTJ083752+445024 & 08 & 37 & 52.9 & 44 & 50 & 24.6 & 11158.2 & 4561.5 & 4662.0 & 2890.0 & 1519.2 & $-$ & 1528.9 & 1139.8 & $-$ & $-$ & 1.01 \\
GMRTJ084639+462141 & 08 & 46 & 39.7 & 46 & 21 & 41.9 & 3201.8 & 1299.4 & 1303.0 & $-$ & $-$ & $-$ & 362.3 & 380.0 & $-$ & $-$ & 1.03 \\
GMRTJ084737+461405 & 08 & 47 & 37.6 & 46 & 14 & 5.3 & 2819.6 & 1028.8 & 1043.0 & $-$ & $-$ & $-$ & 262.1 & 217.3 & $-$ & $-$ & 1.16 \\
GMRTJ084034+465112 & 08 & 40 & 34.3 & 46 & 51 & 12.8 & 2225.0 & $-$ & 928.0 & $-$ & $-$ & $-$ & 287.8 & 297.6 & $-$ & $-$ & 0.94 \\
GMRTJ084530+461035 & 08 & 45 & 30.6 & 46 & 10 & 35.5 & 47.1 & 28.0 & 19.0 & $-$ & $-$ & $-$ & 7.8 & 7.9 & $-$ & $-$ & 0.80 \\
\end{tabular}
\end{table*}
\subsection{Spectral index estimates}
The counterparts for the 150 MHz sources at higher radio frequencies were
searched within a 20${''}$ radius from the 150 MHz position. In addition to
the published deep observations of the LBDS field at 327, 610, 1412, 1462
and 4860 MHz, we used the WENSS catalog at 325 MHz and the NVSS and
FIRST catalogs at 1400 MHz to obtain the spectral index of sources found
at 150 MHz. Since the aim of the project is to search for steep
spectrum sources, the spectral indices were computed using the GMRT 150
MHz catalog as the primary catalog. The spectral index was computed if
a counterpart was found at 610 MHz or at any of the higher frequencies.
If the
counterpart was seen only at 327 MHz, the spectral index was not
computed, since the two frequencies are quite close by and the error
in the spectral index would be large. Nonetheless, if the source was detected at
327 MHz either in WENSS or at the deep observation of this field at
327 MHz, this value of flux was also used to fit the spectrum
along with other available measurements. The typical error on the
spectral index estimate was about 0.1, or better. A small fraction of
the sources (3\%) do not fit well with a straight spectrum and these are
mostly sources with spectral turnover. Analysis of these sources is
beyond the scope of this paper. The spectral index distribution is
presented in Fig. 6. The full table is available online.
\begin{figure}
\includegraphics[width=59mm,angle=270]{figs/fig5.eps}
\caption{Histogram of spectral index distribution of all sources with spectral index
measurements. The median value of spectral index is 0.78.}
\end{figure}
A total of 639 sources out of 765 (83\%) have spectral index determined.
The remaining 17\% sources are mostly
weak radio sources with a median flux density of $\sim$ 9 mJy, or fall
in the regions where deep observations at higher frequencies do not
exist. The median spectral index of the sample is 0.78 (Fig. 5).
This is in good agreement with the similar measurements available in
the literature (e.g., Oort, Steemers and Windhorst, 1988; Gruppioni et
al. 1997). About 15 sources have spectra index flatter than 0.1, 100 sources
have spectra flatter than 0.5, 157 have a spectral index steeper than 1,
and about 20 have spectral index steeper than 1.3. The steep spectrum sources are of
primary interest in this work, and we discuss these sources
in the next section. Our sample of steep spectrum sources consists of
sources with spectra steeper than 1 (and not 1.3 as used by several
authors earlier) for the following reasons; (a) Despite the fact that
most previous searches are limited to sources with spectral index
steeper than 1.3, the median spectral index of known HzRGs is 1.31.
(b) Though Klamer et al. (2006) have shown that the spectra for the
majority of HzRGs are straight over a large frequency range, evidence
for spectral curvature cannot be ruled out completely (Bornancini et
al. 2007). This means that at frequencies as low as 150 MHz, a higher
cut-off in the spectral index will translate into even higher cutoff
at the rest-frame of high-redshift objects. (c) From the
$P-\alpha$ relation (Mangalam \& Gopal-Krishna, 1995; Blundell et al. 1999), sources which are less
powerful (at low radio frequencies) can have marginally flatter
spectra. Since we are aiming to detect HzRGs that are 10 to 100 times
less luminous than the known HzRGs, we could miss HzRGs if we adopt a
steep spectral index cut-off of 1.3. Finally, our catalogue
will be cross-checked against SDSS for possible optical counterparts,
which would naturally eliminate any nearby objects that have steeper spectral index
for any other reason (e.g: synchrotron losses in galaxy clusters).
Therefore, we adopt a spectral index cutoff of 1, which yields a sample
of 157 radio sources selected at 150 MHz, which are then compared with
available deep optical observations.
\begin{figure}
\includegraphics[width=59mm,angle=270]{figs/fig6.eps}
\caption{The spectral index distribution of sources detected
at 150 MHz (S$_\nu \propto \nu^{-\alpha}$). The spectral index was computed whenever a
counterpart was found at 610 MHz or higher frequencies, irrespective
of detection at 325 MHz. The dashed line shows the
spectral index limit corresponding to five sigma limit
of FIRST. There is a weak trend indicating that the fainter
sources tend to have flatter spectra.}
\end{figure}
\begin{figure*}
\vbox{
\hbox{
\includegraphics[width=40mm,angle=270]{figs/fitsrc1.eps}
\includegraphics[width=40mm,angle=270]{figs/fitsrc2.eps}
\includegraphics[width=40mm,angle=270]{figs/fitsrc3.eps}
}
\vspace{0.2in}
\hbox{
\includegraphics[width=40mm,angle=270]{figs/fitsrc4.eps}
\includegraphics[width=40mm,angle=270]{figs/fitsrc5.eps}
\includegraphics[width=40mm,angle=270]{figs/fitsrc6.eps}
}
\vspace{0.2in}
\hbox{
\includegraphics[width=40mm,angle=270]{figs/fitsrc7.eps}
\includegraphics[width=40mm,angle=270]{figs/fitsrc8.eps}
\includegraphics[width=40mm,angle=270]{figs/fitsrc9.eps}
}
}
\caption{Radio spectra of a few steep spectrum sources without an optical counterpart in SDSS.
Note that for majority of the sources, the spectrum is straight over the frequency range
from 150 to 1400 MHz.}
\end{figure*}
\subsection{Optical Counterparts in the Sloan Digital Sky Survey (SDSS)}
The position accuracy at 150 MHz is not good enough to search for
optical counterparts in SDSS. We have cross matched the sample of 157
sources, with spectral index steeper than 1 with VLA FIRST survey
(Becker et al 1995). The position accuracy in FIRST survey is better
than an arcsecond. If the counterpart is seen in FIRST, we have taken
the position from FIRST survey to obtain better accuracy. When multiple
components were seen in FIRST, the FIRST images were inspected and the
position of the core (or the centroid in case of a clear compact double
lobe structure) was chosen. In a couple of cases, the radio structures were such
that there was one unresolved compact component and a second component
which is extended. In such cases, we have taken the position of the
unresolved component. In this procedure, it is possible that we
have missed likely counterparts in some cases. Among the 157 sources, 8
sources did not have counterparts in FIRST and hence the 150 MHz
position was used to cross match with SDSS.
We positionally matched each steep spectrum source position with the
photometric object catalog (PhotoObjAll) from the most recent (DR7)
release of the Sloan Digital Sky Survey (SDSS; Abazajian et al. 2009).
59 radio sources from this sample had at least one SDSS primary object
within 6 arcsec (38\% identification). The remaining 98 sources had no optical counterpart
in SDSS within 6${''}$ radius and these are listed in Table 4.
This radio source optical identification fraction is similar to that
found by Ivezic et al. (2002) (30\%) by cross-matching FIRST survey
sources with SDSS.
\subsubsection{Photometric redshifts}
The SDSS Catalog Archive Server (CAS) includes photometric redshift estimates
for most galaxies with flux measurements in multiple SDSS filters. Two
independent methods have been used to estimate the photometric redshift 1. A
template fitting method (Csabai et al. 2003) with outputs listed in
the table photoz of the CAS and 2. A neural network based method (Collister \&
Lahav 2004) with outputs listed in table photoz2 of the CAS. The neural network
based technique suffers from larger uncertainties and biases for
faint galaxies, because training examples with known spectroscopic
redshifts are not numerous enough for these objects. We, therefore,
chose to use only the photometric redshifts obtained by the template
fitting method. 44 of our 157 steep spectrum radio sources have a total
of 57 potential optical counterparts with listed photometric redhifts. Note that
the number of optical counterparts is greater than the number of steep
spectrum sources, because a few radio sources have more than one galaxy
with photometric redshift lying within the 6 arcsec search radius. The
photometric redshifts range from 0.02 to 0.75. The redshift histogram
for these galaxies is shown in Fig 8.
\begin{figure}
\includegraphics[width=85mm,angle=0]{figs/photoz.eps}
\caption{The distribution of photometric redshifts of the steep spectrum
radio sources with optical counterparts in the SDSS.}
\end{figure}
\begin{figure*}
\vbox{
\hbox{
\includegraphics[width=45mm,angle=0]{figs/kntr1.150.eps}
\includegraphics[width=45mm,angle=0]{figs/kntr1.1400.eps}
\includegraphics[width=40mm,angle=0]{figs/kntr2.150.eps}
\includegraphics[width=40mm,angle=0]{figs/kntr2.1400.eps}
}
\vspace{0.2in}
\hbox{
\hbox{
\includegraphics[width=35mm,angle=0]{figs/kntr3.150.eps}
\includegraphics[width=35mm,angle=0]{figs/kntr3.1400.eps}
\includegraphics[width=49mm, angle=0, viewport=-30 -40 650 600, clip]{figs/kntr4.150.eps}
\includegraphics[width=52mm, angle=0, viewport=-30 -10 700 600, clip]{figs/kntr4.1400.eps}
}
}
\vspace{0.1in}
\hbox{
\includegraphics[width=45mm,angle=270]{figs/kntr5.150.eps}
\hspace{0.2in}
\includegraphics[width=45mm,angle=270]{figs/kntr5.1400.eps}
}
\vspace{0.1in}
\hbox{
\includegraphics[width=45mm,angle=270]{figs/kntr6.150.eps}
\hspace{0.5in}
\includegraphics[width=45mm,angle=270]{figs/kntr6.1400.eps}
}
}
\caption{A few extended sources among the steep spectrum source catalogue.
The GMRT 153 MHz and VLA 1400 MHz (FIRST) image pairs are arranged as follows:
Top line GMRTJ083349+432020 and GMRTJ083405+453443; second line
GMRTJ083507+453628 and GMRTJ083833+461651; third line GMRTJ084105+445045
and the image at the bottom is GMRTJ084711+454015 }
\end{figure*}
\begin{table}
\caption{Steep spectrum sources which do not have counterparts
in SDSS (see section 4.3).
The coordinates given are from the GMRT 150 MHz observations and the flux
density (S$_{150}$) is in mJy.}
\begin{tabular}{lllrl}
& & & & \\
NAME & RA & DEC & S$_{150}$ & $\alpha$ \\
& & & & \\
GMRTJ082947+442036 & 08 29 47.7 & 44 20 36.5 & 975.8 & 1.18 \\
GMRTJ083132+450742 & 08 31 32.4 & 45 07 42.8 & 55.9 & 0.99 \\
GMRTJ083156+435728 & 08 31 56.3 & 43 57 28.1 & 1054.0 & 1.14 \\
GMRTJ083212+451934 & 08 32 12.3 & 45 19 34.2 & 90.8 & 1.06 \\
GMRTJ083245+450716 & 08 32 46.0 & 45 07 16.2 & 75.5 & 1.27 \\
GMRTJ083349+432020 & 08 33 49.3 & 43 20 20.4 & 503.1 & 1.37 \\
GMRTJ083405+453443 & 08 34 05.3 & 45 34 43.0 & 1776.4 & 1.08 \\
GMRTJ083421+441057 & 08 34 21.2 & 44 10 57.7 & 905.8 & 1.17 \\
GMRTJ083507+453628 & 08 35 07.7 & 45 36 28.6 & 251.2 & 1.25 \\
GMRTJ083521+453305 & 08 35 22.0 & 45 33 05.3 & 416.8 & 1.06 \\
GMRTJ083527+440950 & 08 35 27.3 & 44 09 50.3 & 596.5 & 1.21 \\
GMRTJ083527+440004 & 08 35 27.4 & 44 00 04.0 & 94.2 & 1.08 \\
GMRTJ083548+450944 & 08 35 48.8 & 45 09 44.5 & 12.4 & 1.05 \\
GMRTJ083620+444826 & 08 36 20.7 & 44 48 26.3 & 17.2 & 1.06 \\
GMRTJ083628+430053 & 08 36 29.0 & 43 00 53.0 & 64.2 & 1.19 \\
GMRTJ083632+450605 & 08 36 32.8 & 45 06 05.4 & 41.3 & 1.10 \\
GMRTJ083701+435554 & 08 37 01.9 & 43 55 54.5 & 12.6 & 1.13 \\
GMRTJ083714+462908 & 08 37 14.8 & 46 29 08.9 & 18.2 & 1.05 \\
GMRTJ083757+443443 & 08 37 57.0 & 44 34 43.4 & 378.6 & 1.06 \\
GMRTJ083808+441910 & 08 38 08.5 & 44 19 10.1 & 39.6 & 1.29 \\
GMRTJ083814+424248 & 08 38 14.0 & 42 42 48.2 & 20.8 & 1.01 \\
GMRTJ083816+460632 & 08 38 16.5 & 46 06 32.7 & 39.3 & 1.04 \\
GMRTJ083825+432522 & 08 38 25.6 & 43 25 22.2 & 315.1 & 1.37 \\
GMRTJ083831+443808 & 08 38 31.5 & 44 38 08.2 & 21.2 & 1.51 \\
GMRTJ083833+461651 & 08 38 33.8 & 46 16 51.6 & 53.3 & 1.53 \\
GMRTJ083840+460825 & 08 38 40.5 & 46 08 25.6 & 618.0 & 1.27 \\
GMRTJ083859+424414 & 08 38 59.3 & 42 44 14.4 & 60.5 & 1.35 \\
GMRTJ083934+463534 & 08 39 34.3 & 46 35 34.0 & 924.1 & 1.21 \\
GMRTJ083957+432301 & 08 39 57.7 & 43 23 01.1 & 225.8 & 1.10 \\
GMRTJ084005+424657 & 08 40 05.8 & 42 46 57.8 & 24.2 & 1.20 \\
GMRTJ084012+432511 & 08 40 12.3 & 43 25 11.8 & 61.9 & 1.07 \\
GMRTJ084013+432319 & 08 40 13.5 & 43 23 19.8 & 66.8 & 1.20 \\
GMRTJ084021+431928 & 08 40 21.9 & 43 19 28.4 & 62.1 & 1.16 \\
GMRTJ084036+443132 & 08 40 36.6 & 44 31 32.2 & 132.7 & 1.02 \\
GMRTJ084045+460817 & 08 40 45.0 & 46 08 17.2 & 45.5 & 1.10 \\
GMRTJ084046+460719 & 08 40 46.3 & 46 07 19.9 & 197.1 & 1.01 \\
GMRTJ084105+445045 & 08 41 05.7 & 44 50 45.8 & 855.8 & 1.14 \\
GMRTJ084157+434105 & 08 41 57.5 & 43 41 05.6 & 86.2 & 1.02 \\
GMRTJ084211+445404 & 08 42 11.7 & 44 54 04.8 & 147.4 & 1.21 \\
GMRTJ084213+440637 & 08 42 13.8 & 44 06 37.7 & 9.2 & 1.00 \\
GMRTJ084220+424651 & 08 42 20.3 & 42 46 51.6 & 158.4 & 1.07 \\
GMRTJ084223+425532 & 08 42 23.8 & 42 55 32.1 & 100.0 & 1.19 \\
GMRTJ084242+453312 & 08 42 42.3 & 45 33 12.8 & 80.1 & 0.99 \\
GMRTJ084251+464505 & 08 42 51.5 & 46 45 05.1 & 198.5 & 1.04 \\
GMRTJ084300+453423 & 08 43 00.7 & 45 34 23.8 & 234.3 & 1.15 \\
GMRTJ084306+445552 & 08 43 06.8 & 44 55 52.1 & 191.7 & 1.16 \\
GMRTJ084311+433104 & 08 43 11.0 & 43 31 04.7 & 118.3 & 1.06 \\
GMRTJ084315+435630 & 08 43 15.8 & 43 56 30.9 & 124.5 & 1.00 \\
GMRTJ084322+461433 & 08 43 22.0 & 46 14 33.0 & 169.2 & 1.08 \\
GMRTJ084332+461703 & 08 43 32.4 & 46 17 03.3 & 192.8 & 1.03 \\
GMRTJ084402+462616 & 08 44 02.0 & 46 26 16.5 & 62.6 & 1.01 \\
GMRTJ084412+451559 & 08 44 12.4 & 45 15 59.7 & 38.1 & 1.04 \\
GMRTJ084415+430136 & 08 44 15.8 & 43 01 36.3 & 49.2 & 1.31 \\
GMRTJ084420+445802 & 08 44 20.4 & 44 58 02.3 & 29.0 & 0.99 \\
GMRTJ084437+442558 & 08 44 37.2 & 44 25 58.5 & 84.0 & 1.53 \\
GMRTJ084505+442543 & 08 45 05.5 & 44 25 43.4 & 28.0 & 1.05 \\
GMRTJ084515+431304 & 08 45 15.1 & 43 13 04.7 & 129.1 & 0.99 \\
GMRTJ084519+434953 & 08 45 19.9 & 43 49 53.1 & 342.7 & 1.04 \\
GMRTJ084522+434350 & 08 45 22.6 & 43 43 50.0 & 219.9 & 0.99 \\
GMRTJ084527+465014 & 08 45 27.6 & 46 50 14.6 & 1439.6 & 1.06 \\
GMRTJ084528+440732 & 08 45 28.7 & 44 07 32.4 & 141.8 & 1.48 \\
\end{tabular}
\end{table}
\begin{table}
\noindent {\bf Table 4. Contd}\\
\begin{tabular}{lllrl}
& & & & \\
NAME & RA & DEC & S$_{150}$ & $\alpha$ \\
& & & & \\
GMRTJ084533+455835 & 08 45 33.2 & 45 58 35.0 & 23.9 & 1.06 \\
GMRTJ084542+461535 & 08 45 42.9 & 46 15 35.9 & 753.3 & 1.11 \\
GMRTJ084559+425705 & 08 45 59.6 & 42 57 05.9 & 191.1 & 1.06 \\
GMRTJ084616+463234 & 08 46 16.2 & 46 32 34.3 & 150.3 & 0.99 \\
GMRTJ084627+453128 & 08 46 27.2 & 45 31 28.7 & 233.9 & 1.15 \\
GMRTJ084635+455030 & 08 46 35.5 & 45 50 30.2 & 19.4 & 1.01 \\
GMRTJ084644+432903 & 08 46 44.7 & 43 29 03.1 & 371.8 & 1.06 \\
GMRTJ084654+455935 & 08 46 54.4 & 45 59 35.7 & 219.7 & 1.11 \\
GMRTJ084711+454015 & 08 47 11.5 & 45 40 15.9 & 1704.3 & 1.04 \\
GMRTJ084728+430551 & 08 47 28.1 & 43 05 51.9 & 124.0 & 1.09 \\
GMRTJ084730+440822 & 08 47 30.5 & 44 08 22.5 & 63.6 & 1.11 \\
GMRTJ084737+461405 & 08 47 37.5 & 46 14 05.2 & 2819.6 & 1.16 \\
GMRTJ084742+451726 & 08 47 42.4 & 45 17 26.8 & 11.0 & 1.07 \\
GMRTJ084743+431711 & 08 47 43.7 & 43 17 11.2 & 17.4 & 1.01 \\
GMRTJ084747+451016 & 08 47 47.7 & 45 10 16.7 & 196.9 & 1.06 \\
GMRTJ084748+441228 & 08 47 48.2 & 44 12 28.7 & 50.3 & 1.08 \\
GMRTJ084755+445359 & 08 47 55.7 & 44 53 59.3 & 143.0 & 1.07 \\
GMRTJ084804+425728 & 08 48 04.3 & 42 57 28.0 & 90.4 & 1.16 \\
GMRTJ084811+460025 & 08 48 11.7 & 46 00 25.3 & 37.5 & 1.61 \\
GMRTJ084824+432240 & 08 48 24.5 & 43 22 40.2 & 132.3 & 1.23 \\
GMRTJ084842+455357 & 08 48 42.3 & 45 53 57.4 & 175.5 & 1.04 \\
GMRTJ084901+445049 & 08 49 01.2 & 44 50 49.3 & 18.8 & 1.88 \\
GMRTJ084938+442311 & 08 49 38.6 & 44 23 11.5 & 686.3 & 1.09 \\
GMRTJ084943+453750 & 08 49 43.1 & 45 37 50.5 & 12.8 & 1.28 \\
GMRTJ085018+432248 & 08 50 18.2 & 43 22 48.3 & 153.7 & 1.19 \\
GMRTJ085019+451435 & 08 50 19.4 & 45 14 35.1 & 18.7 & 1.34 \\
GMRTJ085028+445522 & 08 50 28.3 & 44 55 22.3 & 108.6 & 1.07 \\
GMRTJ085038+445637 & 08 50 38.0 & 44 56 37.5 & 41.9 & 1.45 \\
GMRTJ085044+454559 & 08 50 44.6 & 45 45 59.0 & 37.5 & 1.11 \\
GMRTJ085052+460440 & 08 50 52.9 & 46 04 40.0 & 110.6 & 1.27 \\
GMRTJ085132+441952 & 08 51 32.1 & 44 19 52.4 & 774.6 & 1.15 \\
GMRTJ085157+453448 & 08 51 57.9 & 45 34 48.8 & 66.5 & 1.08 \\
GMRTJ085202+435054 & 08 52 02.6 & 43 50 54.4 & 24.1 & 1.13 \\
GMRTJ085308+441109 & 08 53 08.3 & 44 11 09.0 & 92.6 & 1.09 \\
GMRTJ085322+450357 & 08 53 22.3 & 45 03 57.3 & 907.7 & 1.01 \\
GMRTJ085330+452256 & 08 53 30.3 & 45 22 56.5 & 162.1 & 1.13 \\
GMRTJ085406+445052 & 08 54 06.3 & 44 50 52.5 & 17.1 & 1.09 \\
\end{tabular}
\end{table}
\subsection{Sample of steep spectrum sources}
We present the sample of steep spectrum radio sources which do not have SDSS
counterparts in Table 4.
The median 150 MHz flux density for these sources is $\sim$ 100 mJy, which is
more than an order of magnitude fainter than the median flux density at 150
MHz for known HzRGs (section 2). If we further divide the sample into
two parts, one with 1 $< \alpha < 1.3$ and $\alpha > 1.3$, we find
that the fraction of sources without identification in SDSS is 63\%
for sources with 1 $< \alpha < 1.3$ and 68\% for sources with $\alpha
> 1.3$. Although this difference is not statistically significant, the
trend of increased fraction of sources without identification in SDSS
for sources with steeper spectra goes in the expected direction.
A few examples of the radio spectra among these sources are
given in Figure 7. One of the sources with
steepest spectrum ($\alpha$ = 1.53; GMRT084437+442558) has measurements at 150, 327, 1412 and 4860 MHz.
We have inspected the spectra where three or more frequency measurements
are available for signatures of spectral curvature. There is no clear indication
of spectral curvature in most of the sources. Since the frequency range is
just one order of magnitude, higher frequency radio data are needed to
further investigate the question of spectral curvature. The absence of spectral curvature in these
data is consistent with the results of Klamer et al (2006), also
argued that the spectral curvature is unlikely to be the
reason for steepening of the radio spectra at high-redshifts.
One of the steep spectrum source without SDSS counterpart (GMRTJ084533+455835),
is unresolved at 150 MHz, but resolves into a clear compact FRII source
of about 8 arcsec size in VLA FIRST. Using the 150 MHz flux density and FRI/FRII
break luminosity, we estimate that this source should be at a redshift of $\sim$
2 or higher (for its luminosity to be above the FRI/FRII break).
\subsection{Notes on individual sources:}
Among the 98 steep spectrum sources that do not have counterparts in
SDSS, 68 are unresolved sources and have only one component in 1.4 GHz FIRST as
well. Among the remaining 30 sources, we discuss those which show
significant structure, many of which are unlikely to be at high-redshift (Figure 9),
but remained unidentified in SDSS either due to confusion in getting
the correct radio component and position or due to dust obscuration
in the host galaxy.
\noindent {\bf GMRTJ083349+432020:} This is a well defined double
radio source at 150 MHz as well as in FIRST. The lobes appear to be of
FR-II type, however they are not compact, and seem to have 'diffused
out'. Since the core is not visible, the SDSS identification is
difficult. The spectral index of this source is 1.37. From the relaxed
morphology of the lobes, this could be an example of 'dead AGN', as
highlighted recently using 74 MHz and NVSS data (Dwarakanath \& Kale,
2009).
\noindent {\bf GMRTJ083405+453443:} The 150 MHz map does not reveal any
structure. However, the FIRST image shows interesting 'one-sided jet'
morphology. There is a compact component appearing like a core at the north-east end
and extended emission towards the south-west resembling a lobe, but lacks clear hotspot near the edge.
The plume like structure to the right is also unusual. This is an interesting
source worth further high-resolution observations to understand the unusual morphology.
\noindent {\bf GMRTJ083507+453628:} The structure at 150 MHz just
resolves into two lobes, which are mostly resolved out in FIRST. This
indicates that the lobes do not have compact hotspots. This
steep spectrum is suggestive of the lobes belonging to 'dead AGN'.
\noindent {\bf GMRTJ083833+461651:} This is one of the steepest spectrum source in the
sample with the spectral index of 1.53. The 150 MHz images does not resolve two lobes
clearly. The FIRST image shows three structures. However, the central emitting region
does not appear like a core. It is possible that the two emitting regions on either
side are a pair of lobes, and the central component is an unrelated source.
\noindent {\bf GMRTJ084105+445045:} An asymmetric triple with a bright core at 150 MHz?
The FIRST image resolves this into a linear double and a likely unrelated source.
At first sight, this double radio source resembles the 'double-double' morphology,
however closer inspection reveals that the inner pair of lobes is more relaxed than
the outer pair, which is inconsistent with the standard double-double
morphology (Schoenmakers et al. 2000).
\noindent {\bf GMRTJ084711+454015:} Appears to be a double at 150 MHz.
The lobes on either side showed extended diffuse emission.
The FIRST image clearly resolves the eastern lobe into a diffuse
'relic' type morphology and the west lobe into a head-tail looking structure.
There is no known cluster in this region. The spectral index of diffuse region alone is 1.2
\section{Concluding Remarks}
The radio spectral-index redshift correlation is the most efficient
method used to detect high-redshift radio galaxies. Up to now, about 45
radio galaxies beyond redshift of 3 have been discovered using this
method. We have shown that this known high-z population is just the tip of the
iceberg, in the sense that they are the most luminous objects in this
class. Radio sources which are 10 to 100 times less luminous than these are yet
to be discovered, and these are essential to fully understand the cosmological evolution
of radio galaxies. We have initiated a major
programme to search for steep spectrum radio sources with the GMRT with the
aim to detect high-redshift radio galaxies of moderate luminosity.
The fields for this programme are carefully chosen such that extensive
data already exist at higher radio frequencies and deep optical imaging
and/or spectroscopy is also available for most of the fields.
Here we have presented the results from deep 150 MHz low frequency
radio observations with the Giant Metrewave Radio Telescope (GMRT), India
reaching an rms noise of of $\sim$ 0.7 mJy/beam and a resolution of $\sim$
20$^{''}$. Further, several deep observations exist, mainly at 327, 610 and
1412 MHz for this field. The radio spectral index analysis of the
sources in the field was done using these deep observations and also
using the WENSS at 325 MHz and the NVSS and FIRST at 1400 MHz. We have
demonstrated that this GMRT survey can search for high-redshift radio
galaxies more than an order of magnitude fainter in luminosity
compared to most of the known HzRGs. We provide a sample of about 100
candidate HzRGs with spectral index steeper than 1 and no optical
counterpart to the SDSS sensitivity limits. A significant fraction of the sources are
compact, and are strong candidates for high-redshift radio galaxies.
These sources will need to be followed up at optical and
near-infrared bands to estimate their redshifts.
\section*{Acknowledgments }
We thank the referee for insightful suggestions and Gopal-Krishna for comments on the manuscript.
GMRT is operated by the National Centre for Radio Astrophysics of the Tata
Institute of Fundamental Research. This research has made use of the
NASA/IPAC Extragalactic Database (NED) which is operated by the Jet
Propulsion Laboratory, California Institute of Technology, under
contract with the National Aeronautics and Space Administration.
Funding for the SDSS and SDSS-II has been provided by the Alfred
P. Sloan Foundation, the Participating Institutions, the National
Science Foundation, the U.S. Department of Energy, the National
Aeronautics and Space Administration, the Japanese Monbukagakusho, the
Max Planck Society, and the Higher Education Funding Council for
England. The SDSS Web Site is http://www.sdss.org/.
The SDSS is managed by the Astrophysical Research Consortium for the
Participating Institutions. The Participating Institutions are the
American Museum of Natural History, Astrophysical Institute Potsdam,
University of Basel, University of Cambridge, Case Western Reserve
University, University of Chicago, Drexel University, Fermilab, the
Institute for Advanced Study, the Japan Participation Group, Johns
Hopkins University, the Joint Institute for Nuclear Astrophysics, the
Kavli Institute for Particle Astrophysics and Cosmology, the Korean
Scientist Group, the Chinese Academy of Sciences (LAMOST), Los Alamos
National Laboratory, the Max-Planck-Institute for Astronomy (MPIA),
the Max-Planck-Institute for Astrophysics (MPA), New Mexico State
University, Ohio State University, University of Pittsburgh,
University of Portsmouth, Princeton University, the United States
Naval Observatory, and the University of Washington.
|
\section{Introduction}
Let $R$ be a complete local hypersurface over an algebraically closed field.
Suppose that the characteristic of $k$ is zero and that $R$ has finite Cohen-Macaulay (CM) representation type, namely, there exist only finitely many isomorphism classes of indecomposable maximal Cohen-Macaulay (MCM) $R$-modules.
Then $R$ is isomorphic to the residue ring $k[[x_0,x_1,x_2,\ldots ,x_d]]/(f)$ where $f$ is one of the following polynomials:
$$
\begin{array}{ll}
(A_n) (n\ge 1) & x_0^2+x_1^{n+1}+x_2^2+\cdots +x_d^2, \\
(D_n) (n\ge 4) & x_0^2x_1+x_1^{n-1}+x_2^2+\cdots +x_d^2, \\
(E_6) & x_0^3+x_1^4+x_2^2+\cdots +x_d^2, \\
(E_7) & x_0^3+x_0x_1^3+x_2^2+\cdots +x_d^2, \\
(E_8) & x_0^3+x_1^5+x_2^2+\cdots +x_d^2.
\end{array}
$$
In this case, all objects and morphisms in the category $\mathrm{CM}(R)$ of MCM $R$-modules have been classified completely, that is, the Auslander-Reiten quiver of the stable category $\underline{\mathrm{CM}}(R)$ of $\mathrm{CM}(R)$ has been obtained.
For the details, see \cite{BGS,K,Y}.
Now, assume that $k$ has characteristic different from two, and that $R$ has countable CM representation type, namely, there exist infinitely but only countably many isomorphism classes of indecomposable MCM $R$-modules.
Then $R$ is isomorphic to $k[[x_0,x_1,x_2,\ldots ,x_d]]/(f)$ where $f$ is either of the following \cite[(1.6)]{GK}:
$$
\begin{array}{ll}
(A_\infty^d) & x_0^2+x_2^2+\cdots +x_d^2, \\
(D_\infty^d) & x_0^2x_1+x_2^2+\cdots +x_d^2.
\end{array}
$$
In this case, all objects in $\mathrm{CM}(R)$ have been classified completely \cite{BGS,BD,K}, but morphisms in $\mathrm{CM}(R)$ have not.
The purpose of this paper is to investigate the relationships among objects in $\mathrm{CM}(R)$ by focusing on the objects that are locally free on the punctured spectrum of $R$.
Modules locally free on the punctured spectrum have recently been studied in relation to whose nonfree loci; see \cite{res,stcm}.
We also use nonfree loci to get our results.
Let us introduce here some notation.
We denote by $\mathcal{P}(R)$ the full subcategory of $\mathrm{CM}(R)$ consisting of all modules that are locally free on the punctured spectrum of $R$.
Let $\mathcal{M}(R)$ be the set of nonisomorphic indecomposable MCM $R$-modules that are {\em not} locally free on the punctured spectrum of $R$, and let $\mathcal{V}(M)$ be the nonfree locus of a finitely generated $R$-module $M$.
(Recall that the nonfree locus of $M$ is defined as the set of prime ideals $\mathfrak p$ of $R$ such that the $R_\mathfrak p$-module $M_\mathfrak p$ is nonfree.)
Let $F_1\overset{\partial}{\to}F_0\to M\to 0$ be part of a minimal free resolution of $M$.
Then the (first) syzygy of $M$ is by definition the image of the map $\partial$ and denoted by $\Omega M$.
(Note that it is uniquely determined up to isomorphism.)
The main theorem of this paper is the following.
\begin{theorem}\label{thm0}
Let $k$ be an algebraically closed field of characteristic different from $2$.
Let $R$ be a complete local hypersurface over k of countable CM representation type.
Then, as we stated above, $R$ is (isomorphic to) a residue ring $k[[x_0,x_1,x_2,\ldots ,x_d]]/(f)$, where $f$ is either $({A}^d_{\infty})$ or $({D}^d_{\infty})$.
Let $\mathfrak p_R = (x_0, x_2, \ldots , x_d)$ and $\mathfrak m_R = (x_0, x_1, x_2, \ldots , x_d)$ be ideals of $R$.
The following hold.
\begin{enumerate}[\rm (1)]
\item
There exists an $R$-module $X_R$ such that
\begin{enumerate}[\rm (a)]
\item
$\mathcal{M}(R) = \{ X_R, \Omega(X_R)\}$,
\item
$\mathcal{V}(X_R) = \{ \mathfrak p_R, \mathfrak m_R\} = \mathcal{V}(\Omega(X_R))$.
\end{enumerate}
\item
For each indecomposable $R$-module $M \in \mathcal{P}(R)$, there is an exact sequence
$$
0 \to L \to M\oplus R^n \to N \to 0
$$
of $R$-modules with $L, N \in \mathcal{M}(R)$ and $n\ge 0$.
\end{enumerate}
\end{theorem}
The first assertion of this theorem especially says that there exist at most two indecomposable MCM $R$-modules which are {\em not} locally free on the punctured spectrum of $R$.
The second assertion especially says that the MCM $R$-modules which are locally free on the punctured spectrum of $R$ are dominated by the MCM $R$-modules which are {\em not} locally free on the punctured spectrum of $R$.
On the other hand, the structure of the MCM modules locally free on the punctured spectrum has been clarified by Schreyer \cite{Sc}:
\begin{theorem}[Schreyer]
Let $n$ be a positive integer.
The Auslander-Reiten quiver of the stable category of $\mathcal{P}(R)$ has the following form.
\begin{enumerate}[\rm (1)]
\item
When $f=(A^{2n-1}_\infty)$:
\begin{picture}(400,40)
\put (54,25){$
\xymatrix{
\bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \cdots
}
$}
\end{picture}
\item
When $f=(D^{2n-1}_\infty)$:
\begin{picture}(400,63)
\put (54,48){$
\xymatrix{
\bullet \ar@{.}[d] \ar[r] & \ar[ld] \bullet \ar@{.}[d] \ar[r] & \ar[ld] \bullet \ar@{.}[d] \ar[r] & \ar[ld] \bullet \ar@{.}[d] \ar[r] & \ar[ld] \bullet \ar@{.}[d] \ar[r] & \ar[ld] \bullet \ar@{.}[d] \ar[r] & \ar[ld] \bullet \ar@{.}[d] \ar[r] & \ar[ld] \cdots \\
\bullet \ar[r] & \ar[lu] \bullet \ar[r] & \ar[lu] \bullet \ar[r] & \ar[lu] \bullet \ar[r] & \ar[lu] \bullet \ar[r] & \ar[lu] \bullet \ar[r] & \ar[lu] \bullet \ar[r] & \ar[lu] \cdots
}
$}
\end{picture}
\item
When $f=(A^{2n}_\infty)$:
\begin{picture}(400,40)
\put (54,25){$
\xymatrix{
\cdots \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) & \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \cdots
}
$}
\end{picture}
\item
When $f=(D^{2n}_\infty)$:
\begin{picture}(400,40)
\put (54,25){$
\xymatrix{
\bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \bullet \ar@{.}@(dr,dl) \ar@<.5ex>[r] & \ar@<.5ex>[l] \cdots
}
$}
\end{picture}
\end{enumerate}
\end{theorem}
Consequently, combining our Theorem \ref{thm0} with this result due to Schreyer, we get a new understanding of the structure of the category of MCM modules over hypersurfaces of countable CM representation type.
We give two applications of our Theorem \ref{thm0}.
Using Theorem \ref{thm0}, we are able to calculate the dimension of the triangulated category $\underline{\mathrm{CM}}(R)$ in the sense of Rouquier \cite{R}, and the Grothendieck groups of $\mathrm{CM}(R)$ and $\underline{\mathrm{CM}}(R)$.
\begin{corollary}\label{cor3}
With the notation of Theorem \ref{thm0} the following hold.
\begin{enumerate}[\rm (1)]
\item
The triangulated category $\underline{\mathrm{CM}}(R)$ has dimension $1$.
\item
For $m\ge1$ one has:
\begin{align*}
K_0(\mathrm{CM}(R)) & =\langle[R],[X_R]\rangle\cong
\begin{cases}
\Bbb Z & \text{if }f=(A^1_\infty),\\
\Bbb Z^2 & \text{if }f = (A^{2m}_{\infty})\text{ or }f=(D^{2m-1}_{\infty}),\\
\Bbb Z \oplus \Bbb Z/2\Bbb Z & \text{if }f = (A^{2m+1}_{\infty})\text{ or } f = (D^{2m}_{\infty}),
\end{cases}
\\
K_0(\underline{\mathrm{CM}}(R)) & =\langle[\underline{X_R}]\rangle\cong
\begin{cases}
\Bbb Z & \text{if }f = (A_\infty^{2m})\text{ or }f=(D^{2m-1}_{\infty}),\\
\Bbb Z/2\Bbb Z & \text{if }f = (A^{2m-1}_{\infty})\text{ or }f = (D^{2m}_{\infty}).
\end{cases}
\end{align*}
\end{enumerate}
\end{corollary}
\begin{ac}
This research was conducted in large part while the first and second authors visited Shinshu University in August 2009, greatly supported by 2009 Discretionary Expense of the President of Shinshu University.
The authors are indebted to Yuji Yoshino for his useful and helpful suggestions.
The authors would like to thank Takuma Aihara, Igor Burban and Osamu Iyama very much for their valuable comments.
\end{ac}
\section{One and two dimensional cases}
In this section, we prove Theorem \ref{thm0} in the cases where the ring $R$ is of dimension $1$ and $2$.
The following proposition includes Theorem \ref{thm0} in the $1$-dimensional case.
\begin{proposition}\label{prop1}
Let $S$ be $2$-dimensional regular local ring and let $x_0,x_1$ be a regular system of parameters of $S$.
\begin{enumerate}[\rm (1)]
\item
Let $R = S/(x_0^2)$.
Then $R/(x_0)\cong\Omega(R/(x_0))$, and the following statements hold.
\begin{enumerate}[\rm (a)]
\item
For every indecomposable MCM $R$-module $M$, there is an exact sequence $0 \to L \to M \to N \to 0$ of $R$-modules with $L,N \in \{ 0, R/(x_0) \}$.
\item
One has $\mathcal{M}(R) = \{ R/(x_0) \}$.
\item
One has $\mathcal{V}(R/(x_0)) = \{ (x_0) , (x_0,x_1) \}$.
\end{enumerate}
\item
Let $R = S/(x_0^2x_1)$.
Then $R/(x_0x_1)\cong\Omega(R/(x_0))$, and the following statements hold.
\begin{enumerate}[\rm (a)]
\item
For every indecomposable MCM $R$-module $M$, there is an exact sequence $0 \to L \to M \oplus R^{n} \to N \to 0$ with $L,N \in \{ 0, R/(x_0), R/(x_0x_1) \}$ and $n=0,1$.
\item
One has $\mathcal{M}(R) = \{ R/(x_0), R/(x_0x_1) \}$.
\item
One has $\mathcal{V}(R/(x_0)) = \{ (x_0) , (x_0,x_1) \} = \mathcal{V}(R/(x_0x_1))$
\end{enumerate}
\end{enumerate}
\end{proposition}
\begin{proof}
(1) By \cite[(4.1)]{BGS}, all the nonisomorphic indecomposable MCM $R$-modules are $R$, $R/(x_0)$ and $\mathrm{Cok}\,\varphi_n\ (n=1,2,\dots)$, where $\varphi_n =
\begin{pmatrix}
x_0&x_1^n\\
0&-x_0
\end{pmatrix}$.
We have the following short exact sequence of complexes.
$$
\begin{CD}
@. 0 @. 0 @. 0 \\
@. @VVV @VVV @VVV \\
(\cdots @>{x_0}>> R @>{x_0}>> R @>{x_0}>> R @>>> 0) \\
@. @V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV \\
(\cdots @>{\varphi_n}>> R^2 @>{\varphi_n}>> R^2 @>{\varphi_n}>> R^2 @>>> 0) \\
@. @V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV \\
(\cdots @>{-x_0}>> R @>{-x_0}>> R @>{-x_0}>> R @>>> 0) \\
@. @VVV @VVV @VVV \\
@. 0 @. 0 @. 0
\end{CD}
$$
Taking the long exact sequence of the homology modules, we get an exact sequence $0 \to R/(x_0) \to \mathrm{Cok}\,\varphi_n \to R/(x_0) \to 0$.
Also, the first row makes an exact sequence $0 \to R/(x_0) \to R \to R/(x_0) \to 0$.
We have $\mathcal{V}(R/(x_0)) = \{ \mathfrak p , \mathfrak m \}=\mathrm{Spec}\,R$, where $\mathfrak p=(x_0)$ and $\mathfrak m=(x_0,x_1)$ (cf. \cite[(1.15(4))]{stcm}).
In particular, $R/(x_0)$ belongs to $\mathcal{M}(R)$.
There is an equality $\begin{pmatrix}
-\frac{x_0}{x_1^n}&-1\\
\frac{1}{x_1^n}&0
\end{pmatrix}
\begin{pmatrix}
x_0&x_1^n\\
0&-x_0
\end{pmatrix}
\begin{pmatrix}
1&0\\
-\frac{x_0}{x_1^n}&1
\end{pmatrix}
=
\begin{pmatrix}
0&0\\
0&1
\end{pmatrix}$ of matrices over $R_{\mathfrak p}$.
Hence $(\mathrm{Cok}\,\varphi_n)_{\mathfrak p}\cong R_{\mathfrak p}$ and $\mathrm{Cok}\,\varphi_n\notin\mathcal{M}(R)$.
Therefore $\mathcal{M}(R) = \{ R/(x_0) \}$.
(2) By \cite[(4.2)]{BGS}, all the nonisomorphic indecomposable MCM $R$-modules are $R$, $R/(x_0)$, $R/(x_0x_1)$, $R/(x_0^2)$, $R/(x_1)$, $\mathrm{Cok}\,\varphi^{+}_n$, $\mathrm{Cok}\,\varphi^{-}_n$, $\mathrm{Cok}\,\psi^{+}_n$ and $\mathrm{Cok}\,\psi^{-}_n\ (n=1,2,\dots)$, where $\varphi^{+}_n =
\begin{pmatrix}
x_0&x_1^n\\
0&-x_0
\end{pmatrix},\ \varphi^{-}_n =
\begin{pmatrix}
x_0x_1&x_1^{n+1}\\
0&-x_0x_1
\end{pmatrix},\ \psi^{+}_n =
\begin{pmatrix}
x_0x_1&x_1^n\\
0&-x_0
\end{pmatrix}$ and $\psi^{-}_n =
\begin{pmatrix}
x_0&x_1^n\\
0&-x_0x_1
\end{pmatrix}$.
Setting $x_1^0=1$, for $n\ge 0$ we have commutative diagrams
$$
\begin{CD}
0 @. 0 @. 0 @. 0 \\
@VVV @VVV @VVV @VVV \\
R @>{x_0}>> R @>{x_0x_1}>> R @>{x_0}>> R \\
@V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV \\
R^2 @>{\varphi_n^+}>> R^2 @>{\varphi_n^-}>> R^2 @>{\varphi_n^+}>> R^2 \\
@V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV \\
R @>{-x_0}>> R @>{-x_0x_1}>> R @>{-x_0}>> R \\
@VVV @VVV @VVV @VVV \\
0 @. 0 @. 0 @. 0
\end{CD}
\qquad\qquad
\begin{CD}
0 @. 0 @. 0 @. 0 \\
@VVV @VVV @VVV @VVV \\
R @>{x_0x_1}>> R @>{x_0}>> R @>{x_0x_1}>> R \\
@V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV \\
R^2 @>{\psi_n^+}>> R^2 @>{\psi_n^-}>> R^2 @>{\psi_n^+}>> R^2 \\
@V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV \\
R @>{-x_0}>> R @>{-x_0x_1}>> R @>{-x_0}>> R \\
@VVV @VVV @VVV @VVV \\
0 @. 0 @. 0 @. 0
\end{CD}
$$
with exact rows and columns.
Also we have equalities of matrices over $R$:
\begin{align*}
\begin{pmatrix}
x_0 & 1 \\
1 & 0
\end{pmatrix}
\begin{pmatrix}
x_0 & 1 \\
0 & -x_0
\end{pmatrix}
\begin{pmatrix}
1 & 0 \\
-x_0 & 1
\end{pmatrix}
& =
\begin{pmatrix}
x_0^2 & 0 \\
0 & 1
\end{pmatrix}
,\\
\begin{pmatrix}
1 & 0 \\
x_0 & 1
\end{pmatrix}
\begin{pmatrix}
x_0x_1 & x_1 \\
0 & -x_0x_1
\end{pmatrix}
\begin{pmatrix}
0 & 1 \\
1 & -x_0
\end{pmatrix}
& =
\begin{pmatrix}
x_1 & 0 \\
0 & 0
\end{pmatrix}
.
\end{align*}
Hence there are exact sequences
\begin{align*}
& 0 \to R/(x_0x_1) \to R \to R/(x_0) \to 0, \\
& 0 \to R/(x_0) \to R/(x_0^2) \to R/(x_0) \to 0, \\
& 0 \to R/(x_0x_1) \to R/(x_1) \oplus R \to R/(x_0x_1) \to 0, \\
& 0 \to R/(x_0) \to \mathrm{Cok}\,\varphi^{+}_n \to R/(x_0) \to 0, \\
& 0 \to R/(x_0x_1) \to \mathrm{Cok}\,\varphi^{-}_n \to R/(x_0x_1) \to 0, \\
& 0 \to R/(x_0x_1) \to \mathrm{Cok}\,\psi^{+}_n \to R/(x_0) \to 0, \\
& 0 \to R/(x_0) \to \mathrm{Cok}\,\psi^{-}_n \to R/(x_0x_1) \to 0.
\end{align*}
Put $\mathfrak p=(x_0)$, $\mathfrak q=(x_1)$ and $\mathfrak m=(x_0,x_1)$.
Then we have $\mathrm{Spec}\,R=\{\mathfrak p,\mathfrak q,\mathfrak m\}$, and easily see that the equalities $\mathcal{V}(R/(x_0)) = \mathcal{V}(R/(x_0x_1)) = \{ \mathfrak p , \mathfrak m \}$ and $\mathcal{V}(R/(x_0^2)) = \mathcal{V}(R/(x_1)) = \{ \mathfrak m \}$ hold.
Therefore $R/(x_0), R/(x_0x_1) \in \mathcal{M}(R)$ and $R/(x_0^2), R/(x_1) \notin \mathcal{M}(R)$.
Since $R_{\mathfrak q}$ is a field, all $R_\mathfrak q$-modules are free.
There are equalities of matrices whose entries are in $R_{\mathfrak p}$:
\begin{align*}
\begin{pmatrix}
\frac{x_0}{x_1^n}&1\\
\frac{1}{x_1^n}&0
\end{pmatrix}
\begin{pmatrix}
x_0&x_1^n\\
0&-x_0
\end{pmatrix}
\begin{pmatrix}
1&0\\
-\frac{x_0}{x_1^n}&1
\end{pmatrix}
& =
\begin{pmatrix}
0&0\\
0&1
\end{pmatrix},
\\
\begin{pmatrix}
\frac{x_0}{x_1^n}&1\\
\frac{1}{x_1^n}&0
\end{pmatrix}
\begin{pmatrix}
x_0x_1&x_1^n\\
0&-x_0
\end{pmatrix}
\begin{pmatrix}
1&0\\
-\frac{x_0x_1}{x_1^n}&1
\end{pmatrix}
& =
\begin{pmatrix}
0&0\\
0&1
\end{pmatrix}.
\end{align*}
Hence $(\mathrm{Cok}\,\varphi^{+}_n)_{\mathfrak p} \cong R_{\mathfrak p} \cong (\mathrm{Cok}\,\psi^{+}_n)_{\mathfrak p}$.
Note that $\mathrm{Cok}\,\varphi^{-}_n$ and $\mathrm{Cok}\,\psi^{-}_n$ are the syzygies of $\mathrm{Cok}\,\varphi^{+}_n$ and $\mathrm{Cok}\,\psi^{+}_n$, respectively.
Thus $\mathrm{Cok}\,\varphi^{+}_n$, $\mathrm{Cok}\,\varphi^{-}_n$, $\mathrm{Cok}\,\psi^{+}_n$ and $\mathrm{Cok}\,\psi^{-}_n$ are not in $\mathcal{M}(R)$.
Consequently, we have $\mathcal{M}(R)$ = $\{ R/(x_0), R/(x_0x_1) \}$.
\end{proof}
Next, let us consider the case where the base ring has dimension $2$.
\begin{proposition}\label{prop2}
Let $k$ be an algebraically closed field.
\begin{enumerate}[\rm (1)]
\item
Let $R = k[[x_0,x_1,x_2]]/(x_0x_2)$.
Then $R/(x_2)\cong\Omega(R/(x_0))$, and the following hold.
\begin{enumerate}[\rm (a)]
\item
For any indecomposable MCM $R$-module $M$, there is an exact sequence $0 \to L \to M \to N \to 0$ with $L,N \in \{ 0, R/(x_0), R/(x_2) \}$.
\item
One has $\mathcal{M}(R) = \{ R/(x_0), R/(x_2) \}$.
\item
One has $\mathcal{V}(R/(x_0)) = \{ (x_0,x_2) , (x_0,x_1,x_2) \} = \mathcal{V}(R/(x_2))$.
\end{enumerate}
\item
Let $R = k[[x_0,x_1,x_2]]/(x_0^2x_1-x_2^2)$.
Then $(x_0,x_2)\cong\Omega(x_0,x_2)$, and the following hold.
\begin{enumerate}[\rm (a)]
\item
For any indecomposable MCM $R$-module $M$, there is an exact sequence $0 \to L \to M \oplus R^{n} \to N \to 0$ with $L,N \in \{ 0, (x_0,x_2) \}$ and $n=0,1$.
\item
One has $\mathcal{M}(R) = \{ (x_0,x_2) \}$.
\item
One has $\mathcal{V}((x_0,x_2)) = \{ (x_0,x_2) , (x_0,x_1,x_2) \}$.
\end{enumerate}
\end{enumerate}
\end{proposition}
\begin{proof}
(1) By \cite[(5.3)]{BD}, all indecomposable MCM $R$-modules are $R$, $R/(x_0)$, $R/(x_2)$, $\mathrm{Cok}\,\varphi^{+}_n$ and $\mathrm{Cok}\,\varphi^{-}_n\ (n=1,2,\dots)$, where $\varphi^{+}_n =
\begin{pmatrix}
x_2&x_1^n\\
0&x_0
\end{pmatrix},\ \varphi^{-}_n =
\begin{pmatrix}
x_0&-x_1^n\\
0&x_2
\end{pmatrix}$.
We have a commutative diagram with exact rows and columns:
$$
\begin{CD}
0 @. 0 @. 0 @. 0 \\
@VVV @VVV @VVV @VVV \\
R @>{x_2}>> R @>{x_0}>> R @>{x_2}>> R \\
@V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)
}VV \\
R^2 @>{\varphi_n^+}>> R^2 @>{\varphi_n^-}>> R^2 @>{\varphi_n^+}>> R^2 \\
@V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 1
\end{smallmatrix}
\right)
}VV \\
R @>{x_0}>> R @>{x_2}>> R @>{x_0}>> R \\
@VVV @VVV @VVV @VVV \\
0 @. 0 @. 0 @. 0
\end{CD}
$$
There are exact sequences $0 \to R/(x_2) \to R \to R/(x_0) \to 0$, $0 \to R/(x_2) \to \mathrm{Cok}\,\varphi^{+}_n \to R/(x_0) \to 0$ and $0 \to R/(x_0) \to \mathrm{Cok}\,\varphi^{-}_n \to R/(x_2) \to 0$.
We have $\mathcal{V}(R/(x_0)) = \{ (x_0,x_2), (x_0,x_1,x_2) \}=\mathcal{V}(R/(x_2))$, which implies $R/(x_0),R/(x_2)\in\mathcal{M}(R)$.
Let $\mathfrak p$ be any nonmaximal prime ideal.
Then one of $x_0,x_1,x_2$ is not in $\mathfrak p$, and we have:
\begin{align*}
\begin{pmatrix}
x_0&-x_1^n\\
0&\frac{1}{x_0}
\end{pmatrix}
\begin{pmatrix}
x_2&x_1^n\\
0&x_0
\end{pmatrix}
=
\begin{pmatrix}
0&0\\
0&1
\end{pmatrix}
& \quad\text{if } x_0 \notin \mathfrak p,\\
\begin{pmatrix}
\frac{x_0}{x_1^n}&-1\\
1&0
\end{pmatrix}
\begin{pmatrix}
x_2&x_1^n\\
0&x_0
\end{pmatrix}
\begin{pmatrix}
-x_1^n&x_0\\
x_2&\frac{1}{x_1^n}
\end{pmatrix}
=
\begin{pmatrix}
0&0\\
0&1
\end{pmatrix}
& \quad\text{if } x_1 \notin \mathfrak p,\text{ and}\\
\begin{pmatrix}
0&1\\
1&0
\end{pmatrix}
\begin{pmatrix}
x_2&x_1^n\\
0&x_0
\end{pmatrix}
\begin{pmatrix}
-x_1^n&\frac{1}{x_2}\\
x_2&0
\end{pmatrix}
=
\begin{pmatrix}
0&0\\
0&1
\end{pmatrix}
& \quad\text{if } x_2 \notin \mathfrak p
\end{align*}
over $R_\mathfrak p$.
In each case $(\mathrm{Cok}\,\varphi^{+}_n)_{\mathfrak p}$ is isomorphic to $R_{\mathfrak p}$.
Since $\mathrm{Cok}\,\varphi^{-}_n$ is the syzygy of $\mathrm{Cok}\,\varphi^{+}_n$, we see that $\mathrm{Cok}\,\varphi^{+}_n$ and $\mathrm{Cok}\,\varphi^{-}_n$ are not in $\mathcal{M}(R)$.
Thus $\mathcal{M}(R) = \{ R/(x_0), R/(x_2) \}$ holds.
(2) By \cite[(5.7)]{BD} the following three assertions hold.
\begin{enumerate}[(i)]
\item
All indecomposable MCM $R$-modules are $R$, $\mathrm{Cok}\,\alpha^{+}$, $\mathrm{Cok}\,\alpha^{-}$, $\mathrm{Cok}\,\beta^{+}$, $\mathrm{Cok}\,\beta^{-}$, $\mathrm{Cok}\,\varphi^{+}_n$, $\mathrm{Cok}\,\varphi^{-}_n$, $\mathrm{Cok}\,\psi^{+}_n$ and $\mathrm{Cok}\,\psi^{-}_n\ (n=1,2,\dots)$, where $\alpha^{+} = \begin{pmatrix}
x_2&x_0x_1\\
x_0&x_2
\end{pmatrix},\ \alpha^{-} = \begin{pmatrix}
-x_2&x_0x_1\\
x_0&-x_2
\end{pmatrix}, \beta^{+} = \begin{pmatrix}
x_0^2&x_2\\
x_2&x_1
\end{pmatrix},\ \beta^{-} = \begin{pmatrix}
x_1&-x_2\\
-x_2&x_0^2
\end{pmatrix}$ and
\begin{align*}
& \varphi^{+}_n =
\begin{pmatrix}
x_2&x_0x_1&0&-x_1^{n+1}\\
x_0&x_2&x_1^n&0\\
0&0&x_2&x_0x_1\\
0&0&x_0&x_2
\end{pmatrix},\
\varphi^{-}_n =
\begin{pmatrix}
-x_2&x_0x_1&0&-x_1^{n+1}\\
x_0&-x_2&x_1^n&0\\
0&0&-x_2&x_0x_1\\
0&0&x_0&-x_2
\end{pmatrix},\\
& \psi^{+}_n =
\begin{pmatrix}
x_2&x_0x_1&-x_1^n&0\\
x_0&x_2&0&x_1^n\\
0&0&x_2&x_0x_1\\
0&0&x_0&x_2
\end{pmatrix},\
\psi^{-}_n =
\begin{pmatrix}
-x_2&x_0x_1&-x_1^n&0\\
x_0&-x_2&0&x_1^n\\
0&0&-x_2&x_0x_1\\
0&0&x_0&-x_2
\end{pmatrix}.
\end{align*}
\item
There are isomorphisms $\mathrm{Cok}\,\alpha^{+} \cong \mathrm{Cok}\,\alpha^{-}$, $\mathrm{Cok}\,\beta^{+} \cong \mathrm{Cok}\,\beta^{-}$, $\mathrm{Cok}\,\varphi^{+}_n \cong \mathrm{Cok}\,\varphi^{-}_n$ and $\mathrm{Cok}\,\psi^{+}_n \cong \mathrm{Cok}\,\psi^{-}_n$.
\item
The $R$-modules $\mathrm{Cok}\,\beta^{+},\ \mathrm{Cok}\,\varphi ^{+}_n,\ \mathrm{Cok}\,\psi^{+}_n$ are locally free on the punctured spectrum.
\end{enumerate}
We can easily check that the sequence $R^2\xrightarrow{\alpha^+}R^2\xrightarrow{(x_0,-x_2)}R\to R/(x_0,x_2)\to0$ is exact.
Hence $\mathrm{Cok}\,\alpha^{+}$ is isomorphic to the prime ideal $\mathfrak p:=(x_0,x_2)$ of $R$.
For an ideal $I$ of $R$, denote by $\mathrm{V}(I)$ the set of prime ideals of $R$ containing $I$.
It is easy to see that $\mathcal{V}(\mathfrak p)$ is contained in $\mathrm{V}(\mathfrak p)$, while $\mathfrak p$ belongs to $\mathcal{V}(\mathfrak p)$ because the local ring $R_\mathfrak p$ is not regular.
Thus we obtain $\mathcal{V}(\mathfrak p)=\mathrm{V}(\mathfrak p)=\{\mathfrak p,\mathfrak m\}$, where $\mathfrak m=(x_0,x_1,x_2)$ is the maximal ideal of $R$, and we have $\mathcal{M}(R)=\{\mathfrak p\}$.
Setting $x_1^0=1$, for $n\ge 0$ we have commutative diagrams
$$
\begin{CD}
0 @. 0 @. 0 @. 0 \\
@VVV @VVV @VVV @VVV \\
R^2 @>{\alpha^+}>> R^2 @>{\alpha^-}>> R^2 @>{\alpha^+}>> R^2 \\
@V{
\left(
\begin{smallmatrix}
1 & 0 \\
0 & 1 \\
0 & 0 \\
0 & 0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 & 0 \\
0 & 1 \\
0 & 0 \\
0 & 0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 & 0 \\
0 & 1 \\
0 & 0 \\
0 & 0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 & 0 \\
0 & 1 \\
0 & 0 \\
0 & 0
\end{smallmatrix}
\right)
}VV \\
R^4 @>{\varphi_n^+}>> R^4 @>{\varphi_n^-}>> R^4 @>{\varphi_n^+}>> R^4 \\
@V{
\left(
\begin{smallmatrix}
0 & 0 & 1 & 0 \\
0 & 0 & 0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 0 & 1 & 0 \\
0 & 0 & 0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 0 & 1 & 0 \\
0 & 0 & 0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 0 & 1 & 0 \\
0 & 0 & 0 & 1
\end{smallmatrix}
\right)
}VV \\
R^2 @>{\alpha^+}>> R^2 @>{\alpha^-}>> R^2 @>{\alpha^+}>> R^2 \\
@VVV @VVV @VVV @VVV \\
0 @. 0 @. 0 @. 0
\end{CD}
\qquad\qquad
\begin{CD}
0 @. 0 @. 0 @. 0 \\
@VVV @VVV @VVV @VVV \\
R^2 @>{\alpha^+}>> R^2 @>{\alpha^-}>> R^2 @>{\alpha^+}>> R^2 \\
@V{
\left(
\begin{smallmatrix}
1 & 0 \\
0 & 1 \\
0 & 0 \\
0 & 0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 & 0 \\
0 & 1 \\
0 & 0 \\
0 & 0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 & 0 \\
0 & 1 \\
0 & 0 \\
0 & 0
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
1 & 0 \\
0 & 1 \\
0 & 0 \\
0 & 0
\end{smallmatrix}
\right)
}VV \\
R^4 @>{\psi_n^+}>> R^4 @>{\psi_n^-}>> R^4 @>{\psi_n^+}>> R^4 \\
@V{
\left(
\begin{smallmatrix}
0 & 0 & 1 & 0 \\
0 & 0 & 0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 0 & 1 & 0 \\
0 & 0 & 0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 0 & 1 & 0 \\
0 & 0 & 0 & 1
\end{smallmatrix}
\right)
}VV @V{
\left(
\begin{smallmatrix}
0 & 0 & 1 & 0 \\
0 & 0 & 0 & 1
\end{smallmatrix}
\right)
}VV \\
R^2 @>{\alpha^+}>> R^2 @>{\alpha^-}>> R^2 @>{\alpha^+}>> R^2 \\
@VVV @VVV @VVV @VVV \\
0 @. 0 @. 0 @. 0
\end{CD}
$$
with exact rows and columns.
There is an equality of matrices over $R$
$$
\begin{pmatrix}
0 & -x_0 & 0 & 1 \\
-1 & 0 & 0 & 0 \\
x_0 & -x_2 & 1 & 0 \\
0 & 1 & 0 & 0
\end{pmatrix}
\begin{pmatrix}
x_2 & x_0x_1 & 0 & -x_1 \\
x_0 & x_2 & 1 & 0 \\
0 & 0 & x_2 & x_0x_1 \\
0 & 0 & x_0 & x_2
\end{pmatrix}
\begin{pmatrix}
-1 & 0 & 0 & 0 \\
0 & 0 & 1 & 0 \\
x_0 & 0 & -x_2 & 1 \\
0 & 1 & x_0 & 0
\end{pmatrix}
=
\begin{pmatrix}
x_0^2 & x_2 & 0 & 0 \\
x_2 & x_1 & 0 & 0 \\
0 & 0 & 0 & 0 \\
0 & 0 & 0 & 1
\end{pmatrix},
$$
which gives an isomorphism $\mathrm{Cok}\,\varphi_0^+\cong\mathrm{Cok}\,\beta^+\oplus R$.
Consequently, we obtain exact sequences $0 \to \mathrm{Cok}\,\alpha^+ \to \mathrm{Cok}\,\beta^+\oplus R \to \mathrm{Cok}\,\alpha^+ \to 0$, $0 \to \mathrm{Cok}\,\alpha^{+} \to \mathrm{Cok}\,\varphi^{+}_n \to \mathrm{Cok}\,\alpha^{+} \to 0$ and $0 \to \mathrm{Cok}\,\alpha^{+} \to \mathrm{Cok}\,\psi^{+}_n \to \mathrm{Cok}\,\alpha^{+} \to 0$.
This completes the proof of the second assertion of the proposition.
\end{proof}
\section{General case}
In this section, we give a proof of Theorem \ref{thm0} in the general case.
Throughout this section, we assume that $k$ is an algebraically closed field of characteristic different from two.
First of all, let us recall the definition of a matrix factorization.
\begin{definition}
Let $S=k[[x_0,x_1,\dots,x_n]]$ and $0\ne f\in(x_0,x_1,\dots,x_n)S$.
A pair of square matrices $(A,B)$ with entries being in $S$ is called a {\it matrix factorization} of $f$ if it satisfies $AB=BA=fE$, where $E$ is an identity matrix.
\end{definition}
Let $\mathrm{MF} (f)$ be the category of matrix factorizations of $f$.
Eisenbud \cite[\S 6]{E} (see also \cite[Chap. 7]{Y}) proved that taking the cokernel induces a category equivalence $\mathrm{MF} (f)/\langle(1,f)\rangle\overset{\cong}{\longrightarrow}\mathrm{CM} (S/(f))$, where $\langle(1,f)\rangle$ denotes the ideal of the category $\mathrm{MF} (f)$ generated by $(1,f)$.
We identify $\mathrm{CM} (S/(f))$ with $\mathrm{MF} (f)/\langle(1,f)\rangle$.
The following lemma is called Kn\"orrer's periodicity (cf. \cite[(3.1)]{K}, \cite[Chap. 12]{Y}) and it will play a key role in this section.
\begin{lemma}\label{knorrer}
Let $S=k[[x_0,x_1,\dots,x_n]]$ and $T=k[[x_0,x_1,\dots,x_n,y,z]]$ be formal power series rings and $f \in (x_0,x_1,\dots,x_n)S$ a nonzero element.
The functor $\mathrm{MF} (f) \to \mathrm{MF} (f+yz)$ given by $(A,B)\mapsto\left(\left(
\begin{smallmatrix}
A & yE\\
zE & -B
\end{smallmatrix}
\right),\left(
\begin{smallmatrix}
B & yE\\
zE & -A
\end{smallmatrix}
\right)\right)$ induces a triangle equivalence between the stable categories $F: \underline{\mathrm{CM}} (S/(f))\overset{\cong}{\longrightarrow}\underline{\mathrm{CM}} (T/(f+yz))$.
\end{lemma}
Here we recall some basic properties of the stable category of MCM modules.
Let $R$ be a Henselian Gorenstein local ring.
For MCM $R$-modules $M$ and $N$, we have $M\cong N$ in $\underline{\mathrm{CM}}(R)$ if and only if $M\oplus R^m\cong N\oplus R^n$ in $\mathrm{CM}(R)$ for some $m,n\ge 0$.
Since $R$ is Henselian, for any object $M\in\underline{\mathrm{CM}}(R)$ there exists a unique object $M_0\in\mathrm{CM}(R)$ such that $M_0$ has no nonzero free summand and that $M_0\cong M$ in $\underline{\mathrm{CM}}(R)$.
If $M$ is an indecomposable object of $\underline{\mathrm{CM}}(R)$, then $M_0$ is a nonfree indecomposable MCM $R$-module.
\begin{proposition}\label{prop3}
Let $S,T,f$ and $F$ be as in Lemma \ref{knorrer}.
Put $R=T/(f+yz)$ and $R'=S/(f)$.
For a nonfree MCM $R'$-module $M$, the following statements hold.
\begin{enumerate}[\rm (1)]
\item
One has an inclusion $\mathcal{V}_{R}(FM)\subseteq\mathrm{V}_R(y,z)$.
\item
One has a bijection $\Phi:\mathcal{V}_{R}(FM) \rightarrow \mathcal{V}_{R'}(M)$ which sends $\mathfrak P$ to $\mathfrak P/(y,z)$.
\end{enumerate}
\end{proposition}
\begin{proof}
(1) Let $(A,B) \in \mathrm{MF} (f)$ be a matrix factorization corresponding to $M$.
Then $\mathrm{Cok}\left(
\begin{smallmatrix}
A & yE\\
zE & -B
\end{smallmatrix}
\right)$ is isomorphic to $FM$ up to free summand, and $FM$ is a nonfree MCM $R$-module.
Let $\mathfrak P\in\mathrm{Spec}\,R$ with $y\notin\mathfrak P$.
Then there is an equality $\begin{pmatrix}
\frac{1}{y}B&E\\
\frac{1}{y}E&0
\end{pmatrix}
\begin{pmatrix}
A & yE\\
zE & -B
\end{pmatrix}
\begin{pmatrix}
yE&0\\
-A&E
\end{pmatrix}
=
\begin{pmatrix}
0&0\\
0&E
\end{pmatrix}$ of matrices over $R_\mathfrak P$.
Hence $(FM)_{\mathfrak P}$ is a free $R_{\mathfrak P}$-module, which implies $\mathcal{V}_{R}(FM)\subseteq\mathrm{V}_{R}(y)$.
Similarly, $\mathcal{V}_{R}(FM)\subseteq\mathrm{V}_{R}(z)$ holds.
(2) The assignment $\mathfrak P\mapsto\mathfrak P/(y,z)$ makes a bijection $\mathrm{V}_R(y,z)\to\mathrm{Spec}\,R'$.
For $\mathfrak P \in \mathrm{V}_{R}(y,z)$, set $\mathfrak p = \mathfrak P /(y,z) \in \mathrm{Spec}\, R'$.
It is enough to show that $\mathfrak P\in\mathcal{V}_{R}(FM)$ if and only if $\mathfrak p\in\mathcal{V}_{R'}(M)$.
There are a ring isomorphism $R/(y,z) \cong R'$ and an $R'$-module isomorphism $FM/(y,z)FM \cong M \oplus \Omega_{R'} M$ up to free summand.
We have $R'_\mathfrak p$-module isomorphisms $(FM)_{\mathfrak P}/(y,z)(FM)_{\mathfrak P}\cong \left( FM/(y,z)FM \right) _{\mathfrak P}\cong M_{\mathfrak P} \oplus (\Omega_{R'} M)_{\mathfrak P} \cong M_\mathfrak p \oplus (\Omega_{R'} M)_\mathfrak p \cong M_\mathfrak p \oplus \Omega_{R'_\mathfrak p} M_\mathfrak p$ up to free $R'_\mathfrak p$-summand.
Since $y,z$ is an $R$-regular sequence and $FM$ is a (nonzero) MCM $R$-module, the sequence $y,z$ is $FM$-regular.
Hence $(FM)_{\mathfrak P}$ is $R_{\mathfrak P}$-free if and only if $M_\mathfrak p \oplus \Omega_{R'_\mathfrak p} M_\mathfrak p$ is $R'_\mathfrak p$-free.
Therefore $\mathfrak P \in \mathcal{V}_{R}(FM)$ if and only if $\mathfrak p \in \mathcal{V}_{R'}(M)$.
\end{proof}
Note that in general a nonfree finitely generated module $M$ over a commutative local ring $(R,\mathfrak m)$ is locally free on the punctured spectrum of $R$ if and only if the equality $\mathcal{V}_R(M)=\{\mathfrak m\}$ holds.
Thus Proposition \ref{prop3} yields the corollary below.
\begin{corollary}\label{cor1}
With the notation of Proposition \ref{prop3}, $M$ is locally free on the punctured spectrum of $R'$ if and only if $FM$ is locally free on the punctured spectrum of $R$.
\end{corollary}
Now we can prove Theorem \ref{thm0} in the general case.
\begin{tpf}
Let $R = k[[x_0,x_1,x_2,\ldots ,x_d]]/(f)$, where $f$ is either $({A}^d_{\infty})$ or $({D}^d_{\infty})$.
We induce on $d=\dim R$.
We may assume $d \geq 3$ thanks to Propositions \ref{prop1} and \ref{prop2}.
Put $R' = k[[x_0,x_1,x_2,\ldots ,x_{d-2}]]/(f')$ such that $f=f'+x_{d-1}^2+x_d^2$.
Then $f'$ is either $({A}^{d-2}_{\infty})$ or $({D}^{d-2}_{\infty})$.
As the characteristic of $k$ is not $2$, the ring $R$ is isomorphic to $k[[x_0,x_1,x_2,\ldots ,x_{d-2},y,z]]/(f'+yz)$, where $y,z$ are indeterminates over $k[[x_0,x_1,x_2,\dots,x_{d-2}]]$.
Let $F:\underline{\mathrm{CM}} (R') \to \underline{\mathrm{CM}} (R)$ be the equivalence in Lemma \ref{knorrer}.
(1) There is a nonfree indecomposable MCM $R'$-module $X_{R'}$ with $\mathcal{M}(R') = \{ X_{R'}, \Omega_{R'}(X_{R'})\}$ and $\mathcal{V}_{R'}(X_{R'}) = \{ (x_0, x_2, \dots , x_{d-2})R', (x_0, x_1, x_2,\dots , x_{d-2})R'\} = \mathcal{V}_{R'}(\Omega _{R'}(X_{R'}))$ by the induction hypothesis.
Let $X_R$ be the nonfree direct summand of $FX_{R'}$.
Recall that the shift functor on the triangulated category $\underline{\mathrm{CM}} (R)$ (respectively, $\underline{\mathrm{CM}} (R')$) is the cosyzygy functor $\Omega^{-1}_R$ (respectively, $\Omega^{-1}_{R'}$).
Hence $F$ commutes with the syzygy functor, and we have $F(\Omega _{R'}(X_{R'})) \cong \Omega _R (FX_{R'}) \cong \Omega _R (X_R)$ in $\underline{\mathrm{CM}} (R)$.
It follows from this that $\Omega _R(X_R)$ is the nonfree direct summand of $F(\Omega _{R'}(X_{R'}))$.
By Proposition \ref{prop3} and Corollary \ref{cor1}, we see that $\mathcal{M}(R) = \{ X_R, \Omega_R (X_R) \}$ and that $\mathcal{V}_R(X_{R}) = \{ (x_0, x_2, \dots , x_{d})R, (x_0, x_1, x_2,\dots , x_{d})R\} = \mathcal{V}_R(\Omega _{R}(X_{R}))$.
(2) Let $M\in\mathcal{P}(R)$ be an indecomposable $R$-module.
Then there exists $M'\in\mathrm{CM}(R')$ such that $FM'$ is isomorphic to $M$ up to free summand.
Since the $R$-module $M$ is indecomposable, $M'$ can be chosen as an indecomposable $R'$-module.
Corollary \ref{cor1} implies that $M'$ is in $\mathcal{P}(R')$.
By induction hypothesis, there is an exact sequence $0 \to L' \to M'\oplus R'^n \to N' \to 0$, where $L', N' \in \mathcal{M}(R')$ and $n\ge 0$.
Then we obtain an exact triangle $L' \to M' \to N' \to \Omega _{R'}^{-1} L'$ in $\underline{\mathrm{CM}} (R')$, which gives an exact triangle $FL' \to M \to FN' \to \Omega _R^{-1}FL'$ in $\underline{\mathrm{CM}} (R)$.
Let $L$ and $N$ be the nonfree direct summands of $FL'$ and $FN'$, respectively.
Then we have an exact triangle $L\to M\to N \to \Omega_R^{-1}L$, which gives a short exact sequence $0 \to L \to M \oplus R^m \to N \to 0$ of $R$-modules.
Since $L'$ and $N'$ belong to $\mathcal{M}(R')$, the modules $L$ and $N$ are in $\mathcal{M}(R)$ by Corollary \ref{cor1}.
\qed
\end{tpf}
\section{Applications}
In this section, we give some applications of Theorem \ref{thm0}.
First, we calculate the dimension of the triangulated category $\underline{\mathrm{CM}}(R)$.
For the definition of the dimension of a triangulated category, see \cite[(3.2)]{R}.
\begin{proposition}
The dimension of $\underline{\mathrm{CM}}(R)$ is equal to $1$.
\end{proposition}
\begin{proof}
Theorem \ref{thm0}(2) especially says that the dimension of $\underline{\mathrm{CM}}(R)$ is at most $1$.
Note that our hypersurface $R$ is not of finite CM representation type and that every MCM $R$-module is isomorphic to its second syzygy up to free summand.
Hence the dimension of $\underline{\mathrm{CM}}(R)$ is nonzero.
Now the conclusion follows.
\end{proof}
Next, let us consider the Grothendieck group $K_0(\mathrm{CM}(R))$ of $\mathrm{CM}(R)$.
Applying Proposition \ref{prop1} and \ref{prop2}, we can calculate the Grothendieck group $K_0(\mathrm{CM}(R))$ of $\mathrm{CM}(R)$ for the hypersurfaces $R$ of types $(A_\infty^1)$, $(D_\infty^1)$, $(A_\infty^2)$ and $(D_\infty^2)$.
\begin{proposition}\label{1755}
With the notation of Theorem \ref{thm0}, we have
$$
K_0(\mathrm{CM}(R))\cong
\begin{cases}
\Bbb Z & \text{if }f = (A^1_{\infty}),\\
\Bbb Z^2 & \text{if }f = (D^1_{\infty})\text{ or }f = (A^2_{\infty}),\\
\Bbb Z \oplus \Bbb Z/2\Bbb Z & \text{if }f = (D^2_{\infty}).
\end{cases}
$$
\end{proposition}
\begin{proof}
Let $R$ be as in Theorem \ref{thm0} and $d=1,2$.
By Propositions \ref{prop1} and \ref{prop2}, there is an epimorphism $\Bbb Z^2 \to K_0(\mathrm{CM}(R))$.
Indeed, sending the canonical basis $\{\binom{1}{0},\binom{0}{1}\}$ of $\Bbb Z^2$ to
\begin{align*}
\{[R],[R/(x_0)]\} \quad & \text{if }f=(A_\infty^1),(D_\infty^1)\text{ or }(A_\infty^2),\\
\{[R],[(x_0,x_2)]\} \quad & \text{if }f=(D_\infty^2)
\end{align*}
makes such a surjection.
We get an exact sequence $0 \to \Bbb Z^{2-r} \to \Bbb Z^2 \to K_0(\mathrm{CM}(R)) \to 0$, where $r$ is the rank of $K_0(\mathrm{CM}(R))$.
Let $Q$ be the total quotient ring of $R$.
For a commutative ring $A$, denote by $\mathrm{mod}\,A$ the category of finitely generated $A$-modules.
Define homomorphisms $a:K_0(\mathrm{CM}(R))\to K_0(\mathrm{mod}\,R)$ and $b:K_0(\mathrm{mod}\,R) \to K_0(\mathrm{mod}\,Q)$ by $a([M])=[M]$ and $b([N])=[N\otimes_RQ]$ for $M\in\mathrm{CM}(R)$ and $N\in\mathrm{mod}\,R$.
For $N\in\mathrm{mod}\,R$, there is an exact sequence $0 \to X_n \to X_{n-1} \to \cdots \to X_0 \to N \to 0$ with $X_i\in\mathrm{CM}(R)$ for $0\le i\le n$.
(For instance, take a free resolution of $N$.)
Then in $K_0(\mathrm{mod}\,R)$ the equality $[N]=\sum_{i=0}^n(-1)^i[X_i]$ holds, which shows that $a$ is surjective.
Clearly, $b$ is also.
Taking the composition, we have a surjection $K_0(\mathrm{CM}(R))\to K_0(\mathrm{mod}\,Q)$.
As $Q$ is Artinian, every finitely generated $Q$-module has finite length, and $K_0(\mathrm{mod}\,Q)$ is the free $\Bbb Z$-module with basis $\{ [Q/\mathfrak M]\mid\mathfrak M\text{ is a maximal ideal of }Q\}$ (cf. \cite[(1.7)]{ARS}).
When $f=(A^1_{\infty})$, the ring $Q$ has a unique maximal ideal.
Hence $K_0(\mathrm{mod}\,Q)\cong\Bbb Z$.
Since there is an exact sequence $0\to R/(x_0)\to R\to R/(x_0)\to 0$, the equality $[R]=2[R/(x_0)]$ holds in $K_0(\mathrm{CM}(R))$.
Therefore we have $r=1$.
There is a commutative diagram
$$
\begin{CD}
0 @>>> \Bbb Z @>{
\left(
\begin{smallmatrix}
1 \\
-2
\end{smallmatrix}
\right)}>> \Bbb Z^2 @>>> K_0(\mathrm{CM}(R)) @>>> 0 \\
@. @| @V{\cong}V{
\left(
\begin{smallmatrix}
1 & 0 \\
2 & 1
\end{smallmatrix}
\right)
}V @VVV \\
0 @>>> \Bbb Z @>{
\left(
\begin{smallmatrix}
1 \\
0
\end{smallmatrix}
\right)}>> \Bbb Z^2 @>>> \Bbb Z @>>> 0
\end{CD}
$$
with exact rows.
Thus, the $\Bbb Z$-module $K_0(\mathrm{CM}(R))$ is isomorphic to $\Bbb Z$.
When either $f=(D^1_{\infty})$ or $f=(A^2_{\infty})$, the ring $Q$ has two maximal ideals, and we have $K_0(\mathrm{mod}\,Q)\cong\Bbb Z^2$.
The $\Bbb Z$-module $K_0(\mathrm{CM}(R))$ is also isomorphic to $\Bbb Z^2$.
When $f=(D^2_{\infty})$, the ring $Q$ is a field.
Hence $K_0(\mathrm{mod}\,Q)\cong\Bbb Z$.
With the notation of the proof of Proposition \ref{prop2}(2), we have isomorphisms $\mathrm{Cok}\,\alpha^-\cong\mathrm{Cok}\,\alpha^+\cong(x_0,x_2)$ and an exact sequence $0\to\mathrm{Cok}\,\alpha^-\to R^2\to\mathrm{Cok}\,\alpha^+\to 0$.
Thus $K_0(\mathrm{CM}(R))$ has a relation $2[R]=2[(x_0,x_2)]$, and we see that $r=1$.
There is a commutative diagram
$$
\begin{CD}
0 @>>> \Bbb Z @>{
\left(
\begin{smallmatrix}
2 \\
-2
\end{smallmatrix}
\right)}>> \Bbb Z^2 @>>> K_0(\mathrm{CM}(R)) @>>> 0 \\
@. @| @V{\cong}V{
\left(
\begin{smallmatrix}
1 & 1 \\
0 & -1
\end{smallmatrix}
\right)
}V @VVV \\
0 @>>> \Bbb Z @>{
\left(
\begin{smallmatrix}
0 \\
2
\end{smallmatrix}
\right)}>> \Bbb Z^2 @>>> \Bbb Z\oplus\Bbb Z/2\Bbb Z @>>> 0
\end{CD}
$$
with exact rows.
Therefore $K_0(\mathrm{CM}(R))$ is isomorphic to $\Bbb Z \oplus \Bbb Z/2\Bbb Z$.
\end{proof}
Before moving to the case of higher dimension, we verify that the following relasionship exists between the Grothendieck group of $\mathrm{CM}(R)$ for a general Gorenstein complete local domain $R$ and that of $\underline{\mathrm{CM}}(R)$.
\begin{lemma}\label{k0}
Let $R$ be a complete Gorenstein local domain.
Then we have an exact sequence
$$
0 \to \langle [R] \rangle \xrightarrow{f} K_0(\mathrm{CM}(R)) \xrightarrow{g} K_0(\underline{\mathrm{CM}}(R)) \to 0
$$
of $\Bbb Z$-modules, where $f,g$ are natural maps.
Moreover, $\langle [R] \rangle \cong \Bbb Z$ holds.
\end{lemma}
\begin{proof}
It is trivial that $f$ is an injective map.
As an exact sequence $0 \to X \to Y \to Z \to 0$ of MCM $R$-modules induces an exact triangle $\underline{X} \to \underline{Y} \to \underline{Z} \to \underline{X}[1]$ in $\underline{\mathrm{CM}}(R)$, we have a well-defined map $g$.
Clearly, $g$ is surjective and $gf=0$.
Let $\underline{X} \to \underline{Y} \to \underline{Z} \to \underline{X}[1]$ be an exact triangle in $\underline{\mathrm{CM}}(R)$.
Then we have an exact sequence $0 \to X \to Y \oplus R^n \to Z \to 0$ of MCM $R$-modules, which gives an equality $[X] - [Y] + [Z] = n[R]$ in $K_0(\mathrm{CM}(R))$.
Thus we have the exact sequence in the lemma.
As to the last statement, we have a map $K_0(\mathrm{CM}(R))\to\Bbb Z$ given by $[M]\mapsto \mathrm{rank}_R\,M$, where $\mathrm{rank}_R\,M$ denotes the rank of $M$.
The restriction of this map to $\langle [R] \rangle$ is the inverse map of the natural surjection $\Bbb Z\to\langle [R] \rangle$.
\end{proof}
The Grothendieck group of $\underline{\mathrm{CM}}(R)$ for a hypersurface $R$ of countable CM representation type is described as follows.
\begin{proposition}\label{prop4}
With the notation of Theorem \ref{thm0}, for $m \geq 1$ we have
$$
K_0(\underline{\mathrm{CM}}(R))\cong
\begin{cases}
\Bbb Z/2\Bbb Z & \text{if }f = (A_{\infty}^{2m-1})\text{ or }f = (D_{\infty}^{2m}),\\
\Bbb Z & \text{if }f = (D_{\infty}^{2m-1})\text{ or }f = (A_{\infty}^{2m}).
\end{cases}
$$
\end{proposition}
\begin{proof}
By virtue of Kn\"{o}rrer's periodicity (Lemma \ref{knorrer}), we have only to deal with the cases $\dim R=1,2$.
Using Propositions \ref{prop1} and \ref{prop2}, we obtain the assertion.
\end{proof}
\begin{proposition}\label{prop5}
With the notation of Theorem \ref{thm0}, for $m\ge1$ we have
$$
K_0(\mathrm{CM}(R))\cong
\begin{cases}
\Bbb Z \oplus \Bbb Z/2\Bbb Z & \text{if }f = (A^{2m+1}_{\infty})\text{ or } f = (D^{2m+2}_{\infty}),\\
\Bbb Z^2 & \text{if }f = (D^{2m+1}_{\infty})\text{ or }f = (A^{2m+2}_{\infty}).
\end{cases}
$$
\end{proposition}
\begin{proof}
Let $R$ be as in Theorem \ref{thm0}.
As $m\ge1$, the ring $R$ is an integral domain.
Tensoring the quotient field $Q$ of $R$ induces a surjection $K_0(\mathrm{CM}(R))\to K_0(\mathrm{mod}\,Q)$ (this is nothing but the composition $ba$ with the notation in the proof of Proposition \ref{1755}), and $K_0(\mathrm{mod}\,Q)\cong\Bbb Z$ as $Q$ is a field.
Hence $K_0(\mathrm{CM}(R))$ has a direct summand isomorphic to $\Bbb Z$.
We see from Theorem \ref{thm0}(2) that $K_0(\mathrm{CM}(R))\cong\Bbb Z\oplus\Bbb Z/e\Bbb Z$ for some $e\ge0$.
Let $F:\underline{\mathrm{CM}}(R')\to\underline{\mathrm{CM}}(R)$ be the $m$-th power of the triangle equivalence given in Lemma \ref{knorrer}, where $R'$ is the corresponding hypersurface of dimension $1$ or $2$.
When $f$ is either $(D^{2m+1}_{\infty})$ or $(A^{2m+2}_{\infty})$, Proposition \ref{prop4} and Lemma \ref{k0} imply the $\Bbb Z$-isomorphism $K_0(\mathrm{CM}(R)) \cong \Bbb Z^2$.
Let $f=(A^{2m+1}_{\infty})$.
Then, since there is an exact sequence $0\to F((x_0),(x_0))\to R^{2^m}\to F((x_0),(x_0))\to 0$, the equality $2^m[R]=2[F((x_0),(x_0))]$ holds in $K_0(\mathrm{CM}(R))$.
This gives a commutative diagram
$$
\begin{CD}
0 @>>> \Bbb Z @>{
\left(
\begin{smallmatrix}
2^m \\
-2
\end{smallmatrix}
\right)}>> \Bbb Z^2 @>>> K_0(\mathrm{CM}(R)) @>>> 0 \\
@. @| @V{\cong}V{
\left(
\begin{smallmatrix}
1 & 2^{m-1} \\
0 & -1
\end{smallmatrix}
\right)
}V @VVV \\
0 @>>> \Bbb Z @>{
\left(
\begin{smallmatrix}
0 \\
2
\end{smallmatrix}
\right)}>> \Bbb Z^2 @>>> \Bbb Z\oplus\Bbb Z/2\Bbb Z @>>> 0
\end{CD}
$$
with exact rows.
Thus, the $\Bbb Z$-module $K_0(\mathrm{CM}(R))$ is isomorphic to $\Bbb Z\oplus\Bbb Z/2\Bbb Z$.
Let $f=(D^{2m+2}_{\infty})$.
Then, with the notation of the proof of Proposition \ref{prop2}(2), we have an isomorphism $F(\alpha^-,\alpha^+) \cong F(\alpha^+,\alpha^-)$ and an exact sequence $0\to F(\alpha^-,\alpha^+)\to R^{2^{m+1}}\to F(\alpha^+,\alpha^-)\to 0$.
Hence $K_0(\mathrm{CM}(R))$ has a relation $2^{m+1}[R]=2[F(\alpha^+,\alpha^-)]$.
A commutative diagram
$$
\begin{CD}
0 @>>> \Bbb Z @>{
\left(
\begin{smallmatrix}
2^{m+1} \\
-2
\end{smallmatrix}
\right)}>> \Bbb Z^2 @>>> K_0(\mathrm{CM}(R)) @>>> 0 \\
@. @| @V{\cong}V{
\left(
\begin{smallmatrix}
1 & 2^m \\
0 & -1
\end{smallmatrix}
\right)
}V @VVV \\
0 @>>> \Bbb Z @>{
\left(
\begin{smallmatrix}
0 \\
2
\end{smallmatrix}
\right)}>> \Bbb Z^2 @>>> \Bbb Z\oplus\Bbb Z/2\Bbb Z @>>> 0
\end{CD}
$$
with exact rows exists, which shows that $K_0(\mathrm{CM}(R))$ is isomorphic to $\Bbb Z \oplus \Bbb Z/2\Bbb Z$.
\end{proof}
|
\section*{Introduction and main results}
Let $\Delta$ be a regular completely non-holonomic distribution on a connected smooth manifold $M$,
i.e. a vector subbundle of $TM$. This paper concerns the Lie algebra $\op{sym}(\Delta)$ of its symmetries.
In \cite{T} N.\,Tanaka introduced a graded nilpotent Lie algebra (GNLA) $\m_x$ and its
algebraic prolongation $\g_x=\hat\m_x$ at every point $x\in M$ to majorize the symmetry
algebra of $\Delta$ (precise definitions follow in the next section).
One of the purposes of this paper is to prove
\begin{theorem}\label{Thm1}
The Lie algebra $\op{sym}(\Delta)$ of the symmetries of $\Delta$ satisfies:
$$
\dim\sym(\Delta)\le\op{sup}_M\dim\g_x.
$$
\end{theorem}
This statement was proven in \cite{T} (Corollary of Theorem 8.4) for strongly regular systems,
i.e. such $\Delta$ that the GNLA $\m_x$ does not depend on $x$.
Generic distributions on high-dimensional manifolds
(for instance, for rank 2 distributions starting from dimension 8) fail to satisfy this property,
but we relax the assumption to usual regularity.
\begin{rk}\label{rk1}
In fact, the equality in the above theorem, provided that all $\g_x$ are finite-dimensional
Lie algebras, is attained only in one case: when the distribution $\Delta$ is strongly regular
and flat in the Tanaka sense (then it is locally isomorphic to the standard model of \cite{T}).
\end{rk}
\begin{rk}\label{rk2}
If we assume $\dim\g_x<\infty$ $\forall x\in M$, then the above inequality refines to
\begin{equation}\label{ineq-rk2}
\dim\sym(\Delta)\le\op{inf}_M\dim\g_x.
\end{equation}
\end{rk}
Our approach here is to elaborate upon the Tanaka theory of symmetries for distributions
and relate it to the Spencer theory of formal integrability for PDEs.
Existence of this relation is natural since both theories are abstract versions and generalizations
of the Cartan equivalence method \cite{St,SS,Y}.
\begin{cor}\label{cor}
Suppose for every $x\in M$ there exists no grading 0 derivation of the GNLA $\m_x^\CC=\oplus_{i<0}\g_i^\CC$,
which has rank 1 in $\op{gl}(\g_{-1}^\CC)$ and acts trivially on $\g_i^\CC$, $i<-1$.
Then the Lie algebra $\sym(\Delta)$ is finite-dimensional.
\end{cor}
The meaning of the condition in the theorem is that the complex characteristic variety of
(the prolongation-projection of the Lie equation corresponding to) $\Delta$ is empty.
When the distribution is strongly regular this assertion follows from \cite{T} (Corollary 2 of Theorem 11.1).
This latter theorem and corollary concern finite-dimensionality of the Lie algebra
$\g_x$ and are purely algebraic. That is the reason it is "if and only if" statement.
For just regular distributions (even strongly regular but non-flat) there are other
reasons for decrease of the size of the symmetry algebra (for instance, abundance
of the independent components of the curvature).
Our theorems imply a-priory knowledge of finite-dimensionality and size estimate for the algebra
$\op{sym}(\Delta)$. Here is one output (the derived distribution is defined in Section \ref{S1}).
\begin{theorem}\label{Thm2}
Consider a distribution $\Delta$ of rank $n$. Assume that rank of the derived distribution
$\Delta_2$ is greater than $\frac{n(n-1)}2+2$ in the case $n>2$. For $n=2$ we assume that rank of
the 2nd derived distribution $\Delta_3$ is $5$.
Then the Lie algebra $\op{sym}(\Delta)$ is finite-dimensional.
\end{theorem}
Another application of Corollary \ref{cor} is investigation of distributions
with infinite algebra of symmetries. In Section \ref{S8} we construct a model that,
together with the operation of prolongation, allows to classify
all rank 2 and 3 distributions with infinitely many symmetries.
\medskip
\textsc{Acknowledgment.}
I am grateful to Ian Anderson, Keizo Yamaguchi and Igor Zelenko for useful discussions
during the mini-workshop organized by Ian Anderson in Utah State University in November 2009.
\section{Review of Tanaka theory}\label{S1}
Given a distribution $\Delta\subset TM$, its \emph{weak derived flag}
$\{\Delta_i\}_{i>0}$ is given via the module of its sections by
$\Gamma(\Delta_{i+1})=[\Gamma(\Delta),\Gamma(\Delta_i)]$ with $\Delta_1=\Delta$.
We will assume throughout this paper that our distribution is \emph{completely non-holonomic},
i.e. $\Delta_i=TM$ for $i\ge\kappa$, and we also assume that the flag of $\Delta$ is \emph{regular},
so that the ranks of $\Delta_i$ are constant (whence $\kappa$ is constant as well).
The quotient sheaf $\m=\oplus_{i<0}\g_i$, $\g_i=\Delta_{-i}/\Delta_{-i-1}$ (we let $\Delta_0=0$),
has a natural structure of graded nilpotent Lie algebra at
any point $x\in M$. The bracket on $\m$ is induced by the commutator
of vector fields on $M$. $\Delta$ is called \emph{strongly regular} if the GNLA $\m=\m_x$ does not depend
on the point $x\in M$.
The \emph{growth vector} of $\Delta=\g_{-1}$ is the sequence of dimensions\footnote{In other sources, it is $(\dim\Delta_1,\dim\Delta_2,\dots)$.}\linebreak
$(\dim \g_{-1},\dim \g_{-2}\dots)$ depending on $x\in M$.
The Tanaka prolongation $\g=\hat\m$ is the graded Lie algebra with negative graded part $\m$
and non-negative part defined successively by
$$
\g_k=\{u\in\bigoplus\limits_{i<0}\g_{k+i}\ot\g_i^*:
u([X,Y])=[u(X),Y]+[X,u(Y)],\ X,Y\in\m\}.
$$
Since $\Delta$ is bracket-generating, the algebra $\m$ is fundamental, i.e.
$\g_{-1}$ generates the whole GNLA $\m$, and therefore the grading $k$ homomorphism $u$
is uniquely determined by the restriction $u:\g_{-1}\to\g_{k-1}$.
The space $\g=\oplus\g_i$ is naturally a graded Lie algebra, called
the {\em Tanaka algebra\/} of $\Delta$, and to indicate dependence on the point $x\in M$, we
will write $\g=\g_x$ (also the value of a vector field $Y$ at the point $x$ will be denoted by $Y_x$).
Alternatively the above symbolic prolongation can be defined via Lie algebra cohomology with coefficients:
$\g_0=H^1_0(\m,\m)$, $\g_1=H^1_1(\m,\m\oplus\g_0)$ etc, where the subscript indicates
the grading \cite{AK}. The prolongation of $\m$ is $\g=\m\oplus\g_0\oplus\g_1\oplus\dots$.
In particular, if $\g$ is finite-dimensional, then $H^1(\m,\g)=0$ in accordance with \cite{Y}.
In addition to introducing the Lie algebra $\g$, which majorizes the symmetry algebra of $\Delta$,
the paper \cite{T} contains the construction of an important ingredient to the equivalence problem
-- an absolute parallelism on the prolongation manifold of the structure.
Distribution is locally flat if the structure functions of the absolute parallelism vanish.
Then the distribution $\Delta$ is locally diffeomorphic
to the standard model on the Lie group corresponding to $\m$, see \cite{T}.
\section{Lie algebra sheaf of a distribution}\label{S2}
Instead of considering the global object (vector fields) $\op{sym}(\Delta)\subset\Dd(M)$
let us study the more general structure of local symmetries. Namely we consider the
Lie algebras sheaf (LAS) $\Ll$ of germs of vector fields preserving the distribution $\Delta$.
Regularity of the latter implies that $\Ll$ is a sub-sheaf of $\Dd_\text{loc}(M)=\Gamma_\text{loc}(TM)$.
Let $\Ll(x)$ be its stalk at $x\in M$ and $\Ll(x)^0$ the isotropy subalgebra.
Note that in terms of the evaluation map $\op{ev}_x:\Ll(x)\to T_xM$ the latter is $\op{ev}_x^{-1}(0)$.
The LAS $\Ll$ is transitive if $\op{ev}_x$ is onto and in this case all stalks $\Ll(x)$ are isomorphic.
In the opposite case (provided $\Ll$ is integrable and regular) the Frobenius theorem implies that
$M$ is foliated by the leaves of $\op{ev}_x(\Ll(x))$; the stalks are constant along it.
We define the dimension of LAS to be
$$
\dim\Ll=\op{sup}_M(\op{rank}[\op{ev}_x]+\dim\Ll(x)^0).
$$
Since globalization can only decrease objects (see the Appendix for precise conditions), we will have:
$$
\op{dim}\op{sym}(\Delta)\le\dim\Ll.
$$
There are two decreasing filtrations of the stalk $\Ll(x)$ of the LAS $\Ll$ for every $x\in M$.
The first filtration is defined in the transitive case
by the rule \cite{SS}: $\Ll(x)_*^i=\Ll(x)$ for $i<0$, $\Ll(x)_*^0=\Ll(x)^0$
and
$$
\Ll(x)_*^{i+1}=\{X\in\Ll(x)_*^i:[X,\Ll(x)]\subset\Ll(x)_*^i\}\text{ for }i\ge0.
$$
This definition works in the abstract setting and in general there are existence and realization theorems
\cite{SS}. In our case this filtration is closely related to the jets. In fact,
$\Ll(x)_*^i=\Ll(x)\cap\mu_x^{i+1}\cdot\Dd_\text{loc}(M)$, where $\mu_x\subset C^\infty(M)$ is the maximal ideal.
This latter definition of $\Ll(x)_*^i$ ($i\ge0$) works nicely in the intransitive case as well.
The second filtration in the transitive case is defined as follows \cite{T}:
$\Ll(x)^i=\op{ev}_x^{-1}(\Delta_{-i})$ for $i<0$,
$\Ll(x)^0$ as above and
$$
\Ll(x)^{i+1}=\{X\in\Ll(x)^i:[X,\Ll(x)^{-1}]\subset\Ll(x)^i\}\text{ for }i\ge0.
$$
In particular, $\Ll(x)^{-i}=\Ll(x)$ for $i\ge\kappa$.
The two filtrations are related by the following diagram:
$$
\begin{array}{cccccccc}
\Ll(x) \!&= \!&\dots \!&= \Ll(x)^{-1}_* \!&\supset \Ll(x)^0_* \!&\supset \Ll(x)^1_* \!&\supset \Ll(x)^2_* \!&\supset \dots\\
|| & & & \ \cup \!\!& \ \ || \!\!& \ \ \cap \!\!& \ \ \cap \!\!& \\
\Ll(x)^{-\kappa} \!&\supset \!&\dots \!&\supset \Ll(x)^{-1} \!&\supset \Ll(x)^0 \!&\supset \Ll(x)^1 \!&\supset \Ll(x)^2 \!&\supset \dots
\end{array}
$$
In the intransitive case the situation with the second filtration is more complicated and can be resolved as follows.
The non-positive terms keep the same value.
For $X\in\Ll(x)^0$ and $Y\in\Dd(M)$ the value $\Psi_X^1(Y)=[X,Y]_x\in T_xM$ depends on $Y_x\in T_xM$ only.
Thus $\Psi_X^1:T_xM\to T_xM$ is a linear morphism and it maps $\Delta_i$ to $\Delta_i$, $i>0$.
We define
$$
\Ll(x)^1=\{X\in\Ll(x)^0:\Psi_X^1|_{\Delta_x}=0\}.
$$
Thus $X\in\Ll(x)^1$ are characterized by the property $\Psi_X^1:\Delta_i\to\Delta_{i-1}$.
For $X\in\Ll(x)^1$ and $Y,Z\in\Dd(M)$ the value $\Psi_X^2(Y,Z)=[[X,Y],Z]_x\in T_xM$ depends on
$Y_x,Z_x\in T_xM$ provided $Y_x,Z_x\in\Delta_x$.
Indeed, since $\R$-linearity of $\Psi_X^2$ is obvious, consider multiplication by a function $f\in C^\infty(M)$:
\begin{multline*}
[[X,fY],Z]=f[[X,Y],Z]+X(f)[Y,Z]\\
-Z(f)[X,Y]-X(Z(f))Y+[X,Z](f)Y.
\end{multline*}
At the point $x$ all the terms except the first in the r.h.s.\ vanish, implying the claim. The claim
for the second argument is even easier, as the last term below vanishes at $x$:
$$
[[X,Y],fZ]=f[[X,Y],Z]+[X,Y](f)Z.
$$
Thus we have a linear morphism $\Psi_X^2:\Delta_x\otimes\Delta_x\to\Delta_x$,
which we can also treat as a map $\Delta_x\to\Delta_x^*\otimes\Delta_x$,
$Z\mapsto\Psi_X^2(\cdot,Z)$.
We define
$$
\Ll(x)^2=\{X\in\Ll(x)^1:\Psi_X^2|_{\Delta_x\ot\Delta_x}=0\}.
$$
The Jacobi identity implies that $\Psi^1_X|_{\Delta_2}=0$ for $X\in\Ll(x)^2$
and more generally $\Psi_X^1:\Delta_i\to\Delta_{i-2}$ for such $X$ and $i\ge2$.
Continuing in the same way we define $\Ll(x)^i$ for $i>0$ and we inductively get
the multi-linear maps $\Psi_X^{i+1}:\ot^{i+1}\Delta_x\to T_xM$, $X\in\Ll(x)^i$ by the formula
$$
\Psi_X^{i+1}(Y_1,\dots,Y_{i+1})=[[..[[X,Y_1],Y_2],..],Y_{i+1}]_x.
$$
In the next section we extend $\Psi_X^{i+1}$ to various arguments and show
that $\Psi_X^{i+1}:\ot^{i+1}\Delta_x\to\Delta_x$.
Then the next filtration term is
$$
\Ll(x)^{i+1}=\{X\in\Ll(x)^i:\Psi_X^{i+1}|_{\Delta_x\ot..\ot\Delta_x}=0\}.
$$
These two filtrations define the same topology on $\Ll(x)$ because
\begin{equation}\label{2filtr}
\Ll(x)^{i\cdot\kappa}\subset\Ll(x)_*^i\subset\Ll(x)^i,\ \ i\ge0.
\end{equation}
\section{Investigation of the second filtration}\label{S3}
We need first to prove the following characterization of $X\in\Ll(x)^i$, $i\ge0$.
Let $Z_{jl}$ be a frame around $x$, compatible with the weak flag, i.e. $Z_{1l}(x)$ is a basis of
$\g_{-1}=\Delta_x$, $Z_{2l}(x)$ produces a basis of $\g_{-2}=\Delta_2/\Delta_1$ etc.
Decompose a local symmetry $X=\sum f_{jl}\,Z_{jl}$.
Let us say that function $f\in C^\infty(M)$ belongs to $\mu_{\Delta,x}^{k+1}$ if
$Y_1\cdots Y_t(f)$ vanishes at $x$ for all $Y_j\in\Gamma(\Delta)$, $t\le k$
(for $k=0$ we have: $\mu_{\Delta,x}=\mu_x$).
\begin{lem}\label{L0}
If for $i\ge0$ $X\in\Ll(x)^i$, then $f_{jl}\in\mu_{\Delta,x}^{i+j}$, $1\le j\le\kappa$.
\end{lem}
\begin{proof}
$X$ is a local symmetry if for any $Y_{i_1}\in\Gamma(\Delta)$ the Lie bracket
$$
[X,Y_{i_1}]=\sum(f_{jl}[Z_{jl},Y_{i_1}]-Y_{i_1}(f_{jl})Z_{jl})
$$
is also a section of $\Delta$. This equality $\op{mod}\Delta_{j-1}$ implies the following decomposition
\begin{equation}\label{fYi1}
Y_{i_1}(f_{jl})=\sum_{s\ge j-1}\a_{i_1jl}^{st}f_{st},\quad
j>1.
\end{equation}
Differentiation of this implies
\begin{equation}\label{fYi2}
Y_{i_1}Y_{i_2}(f_{jl})=\sum_{s\ge j-2}\a_{i_1i_2jl}^{st}f_{st},\quad j>2;
\end{equation}
\begin{equation}\label{fYi3}
Y_{i_1}Y_{i_2}(f_{2l})=\sum_{s,t}\a_{i_1i_22l}^{st}f_{st}+\b_{i_1i_2l}^tY_{i_1}(f_{1t}).
\end{equation}
Proceeding we get generally for $j>1$
\begin{equation}\label{fYi5}
Y_{i_1}\cdots Y_{i_r}(f_{jl})=\sum_{s\ge j-r}\a_{i_1\dots i_rjl}^{st}\,f_{st}+
\sum_{k=1}^{r-j+1}\b_{i_1\dots i_rjl}^{q_1\dots q_kt}\, Y_{q_1}\cdots Y_{q_k}(f_{1t}),
\end{equation}
where we assume $\sum_A^B=0$ if $A>B$; we can also assume that $(q_1\dots q_k)$ is an ordered subset of
$(i_1\dots i_r)$ though it plays no role in what follows.
Consider at first the case $i=0$. Then obviously $f_{jl}(x)=0$, i.e. $f_{jl}\in\mu^1_x$.
Substituting this into (\ref{fYi1}) we get $f_{jl}\in\mu^2_{\Delta,x}$ for $j\ge2$.
Next substitution into (\ref{fYi2}) yields $f_{jl}\in\mu^3_{\Delta,x}$ for $j\ge3$,
and continuing with (\ref{fYi5}) we get $f_{jl}\in\mu^j_{\Delta,x}$ for all $j$.
Now let us look at $i=1$. By definition $f_{1l}\in\mu^2_{\Delta,x}$. Then
(\ref{fYi3}) implies $f_{2l}\in\mu^3_{\Delta,x}$. Continuing with (\ref{fYi5}) we get
$f_{jl}\in\mu^{j+1}_{\Delta,x}$.
Now the pattern is clear and the general claim
$f_{jl}\in\mu_{\Delta,x}^{i+j}$ is easily obtained by induction.
\end{proof}
\begin{lem}\label{L1/2}
For $X\in\Ll(x)^i$ and $Y_1,\dots,Y_{i+1}\in\Delta_x$ we have:
$$\Psi^{i+1}_X(Y_1,\dots,Y_{i+1})\in\Delta_x.$$
\end{lem}
\begin{proof}
Indeed Lemma \ref{L0} implies that for $X\in\Ll(x)^i$ and $Y_j\in\Gamma(\Delta)$
$$
[[..[X,Y_1],\dots Y_i],Y_{i+1}]=\sum\pm Y_{i+1}\cdots Y_1(f_{1l})Z_{1l}+\dots,
$$
where the omitted terms that are linear combinations of $f_{jl}$, $Y_{q_1}(f_{jl})$, \dots,
$Y_{q_1}\cdots Y_{q_i}(f_{jl})$ and $Y_{q_1}\cdots Y_{q_{i+1}}(f_{tl})$ for $t>1$,
all of which vanish at $x$. Thus the result evaluated at $x$ is a vector from $\Delta_x$.
\end{proof}
Recall that $\Delta_s=0$ for $s\le0$ and $\Delta_s=TM$ for $s\ge\kappa$.
The formulae of the previous section for any $X\in\Ll(x)$ yield the maps ($j,s_\nu>0$)
$$
\Psi_X^j:\Gamma(\Delta_{s_1})\ot\dots\ot\Gamma(\Delta_{s_j})
\to\Gamma(\Delta_{s_1+...+s_j}).
$$
\begin{lem}\label{L1}
If $X\in\Ll(x)^i$ and $Y_t\in\Gamma(\Delta_{s_t})$ for $1\le t\le j\le i+1$, then
$$
\Psi_X^j(Y_1,\dots,Y_j)_x\in\Delta_{s_1+\dots+s_j-i}.
$$
\end{lem}
\begin{proof}
For $i=j$ and $s_1=\dots=s_j=1$ this is just the definition as $\Delta_0=0$.
Without loss of generality we can take decomposable
$Y_t=[[..[V_{t1},V_{t2}],..],V_{ts_t}]_x$, $V_{t\a}\in\Gamma(\Delta)$, $1\le t\le j$,
$1\le\a\le s_t$. Then
as a result of the Jacobi identity $\Psi^j_X(Y_1,\dots,Y_j)$ decomposes into a linear combination
of the terms
$$
[[..[[X,V_{\a_1}],V_{\a_2}]..],V_{\a_r}],\quad r=s_1+\dots+s_j.
$$
For $r\le i$ this vanishes at $x$, i.e. the result belongs to $\Delta_0$.
For $r>i$ we deduce from Lemma \ref{L0} for $X=\sum f_{jl}Z_{jl}$:
$$
[[..[[X,V_{\a_1}],V_{\a_2}]..],V_{\a_r}]_x=\sum_{j\le r-i}\pm V_{\a_r}\cdots V_{\a_1}(f_{jl}Z_{jl})_x\in
\Delta_{r-i}.
$$
Finally for $i=j-1$ the claim follows from Lemma \ref{L1/2} and the extension argument of this proof.
\end{proof}
For $X\in\Ll(x)^i$ the multi-linear map $\Psi_X^j(Y_1,\dots,Y_j)$ of Lemma \ref{L1} is not $C^\infty(M)$-linear in $Y_t$,
but considered with the values modulo $\Delta_{s_1+\dots+s_j-i-1}$ it is.
Thus it induces the map (all spaces evaluated at $x$)
$$
\Psi_X^j:\Delta_{s_1}\ot\dots\ot\Delta_{s_j}\to\g_{i-s_1-...-s_j}
$$
with the kernel $\sum_{t_1+...+t_j<s_1+...+s_j}\!
\Delta_{t_1}\ot\dots\ot\Delta_{t_j}$, whence the map
\begin{equation}\label{Psi}
\Psi_X^j:\g_{-s_1}\ot\dots\ot\g_{-s_j}\to\g_{i-s_1-...-s_j}.
\end{equation}
\begin{rk}
For $X\in\Ll(x)^i$ and $i\ge(j-1)\kappa$ the map
$$
\Psi_X^j:\Delta_{s_1}\ot\dots\ot\Delta_{s_j}\to\Delta_{s_1+...+s_j-i},
$$
is tensorial, so in this case we do not need to quotient.
\end{rk}
\begin{lem}\label{L2}
The second filtration respects the Lie brackets:
$$
[\Ll(x)^i,\Ll(x)^j]\subset\Ll(x)^{i+j}.
$$
\end{lem}
\begin{proof}
Let $X'\in\Ll(x)^i,X''\in\Ll(x)^j$. If $i,j\le0$, then $X'_x\in\Delta_i$, $X''_x\in\Delta_j$ and
so $[X',X'']_x\in\Delta_{i+j}$, i.e. $[X',X'']\in\Ll(x)^{i+j}$.
Consider $i,j\ge0$. Then for $Y_1,\dots,Y_{i+j}\in\Gamma(\Delta)$ we have with some factors $\nu_t$:
\begin{multline*}
[[..[[X',X''],Y_1]..,Y_{i+j}]=\\
\sum_{t;\ \z\in S_{i+j}}\nu_t\,[..[..[X',Y_{\z(1)}]..,Y_{\z(t)}],[..[X'',Y_{\z(t+1)}]..,Y_{\z(i+j)}]]
\end{multline*}
If a term in the above summation has $t\ge i$, then $i+j-t\le j$ and for
$Z=[..[X',Y_{\z(1)}]..,Y_{\z(t)}]$ with $Z_x\in\Delta_{t-i}$ Lemma \ref{L1} implies:
$[[..[X'',Y_{\z(t+1)}]..,Y_{\z(i+j)}],Z]_x=0$.
If $t<i$, then a symmetric ($X'\leftrightarrows X''$) argument applies.
Let finally $i>0,j<0$, $i+j\ge0$. Then
\begin{multline*}
[[..[[X',X''],Y_1]..,Y_{i+j}]=\\
\sum_{t,\z\in S_{i+j}}\nu_t\,[..[X',Y_{\z(1)}]..,Y_{\z(t)}],Z(X'',Y_{\z(t+1)},..,Y_{\z(i+j)})]]
\end{multline*}
where $Z(..)=[..[X'',Y_{\z(t+1)}]..,Y_{\z(i+j)}]$ has $Z_x\in\Delta_{i-t}$, so that the result follows from
Lemma \ref{L1}.
If $i+j<0$, then the claim follows from Lemma \ref{L0}.
\end{proof}
\section{Formal Lie algebra of symmetries}\label{S4}
In Section \ref{S2} we introduced two compatible filtrations (\ref{2filtr}).
The corresponding \textit{formal Lie algebra} is
$$
L^\Delta_x=\lim_{i\to+\infty}\Ll(x)/\Ll(x)^i=\lim_{i\to+\infty}\Ll(x)/\Ll(x)_*^i.
$$
This algebra has two gradations (jets and weighted jets\footnote{Weighted jets play a
crucial role in Morimoto's approach to the equivalence problem \cite{M}.}) corresponding to
the above two filtrations. In fact, the first grading is
$$
\op{gr}_*(L^\Delta_x)=\oplus\,\bar g_i(x),\quad
\bar g_i(x)=\Ll(x)_*^{i-1}/\Ll(x)_*^i\ (i\ge0).
$$
(the reason for the shift of indices will be clear in the next section).
The second grading is the following
$$
\op{gr}(L^\Delta_x)=\oplus\,\mathbf{g}_i(x),\quad
\mathbf{g}_i(x)=\Ll(x)^i/\Ll(x)^{i+1}\ (i\ge-\kappa).
$$
Notice the difference in the range of indices.
Both gradings have the induced Lie bracket, so that we have two graded Lie structures (which
might be different as Lie algebras from $L^\Delta_x$).
We clearly have the inclusion of GNLAs $\oplus_{i<0}\mathbf{g}_i\subset \m_x$, and now would like to
elaborate upon (\ref{Psi}) to understand the symbols $\mathbf{g}_i$, $i\ge0$.
The Tanaka symbol space $\g_0\subset\sum_{i<0}\g_i^*\ot\g_i$ is uniquely determined by
its restriction to $\g_{-1}$ and thus can be identified with its image
$\g_0\hookrightarrow\g_{-1}^*\ot\g_{-1}$.
Similarly, $\g_1\subset(\sum_{j<0}\g_{-1}^*\ot\g_j^*\ot\g_j)\oplus(\sum_{i<-1}\g_{i-1}^*\ot\g_i)$
can be identified with its image
$\g_1\hookrightarrow\g_{-1}^*\ot\g_0\hookrightarrow\g_{-1}^*\ot\g_{-1}^*\ot\g_{-1}$.
In the general case, $\g_i\hookrightarrow\ot^{i+1}\g_{-1}^*\ot\g_{-1}$ is a
monomorphism and we identify the symbol $\g_i$ ($i\ge0$) with its image.
Now notice that by (\ref{Psi}) every
$X\in\Ll(x)^i$ induces the linear map $\Psi_X^j:\ot^j\g_{-1}\to\g_{i-j}$.
\begin{lem}\label{L3}
For $X\in\Ll(x)^i$ the element $\Psi_X^{i+1}\in\ot^{i+1}\g_{-1}^*\ot\g_{-1}$
belongs to (the image of) $\g_i$ ($i\ge0$).
\end{lem}
\begin{proof}
For $i=0$ the claim is obvious since $\Psi^1_X=\op{ad}_X\in\op{Der}_0(\m_x)$.
For $i=1$ the Jacobi identity implies the following symmetry of $\Psi_X^2$, with $Y,Z\in\Gamma(\Delta)$ so that
$[Y,Z]\in\Gamma(\Delta_2)$:
$$
\Psi_X^2(Y,Z)-\Psi_X^2(Z,Y)=\Psi_X^1([Y,Z]).
$$
Moreover this formula holds true if we understand $Y\in\g_j$, $Z\in\g_l$, $[Y,Z]\in\g_{j+l}$,
which provides us the extension $\Psi^2_X\in\sum_{i<0}\g_{i}^*\ot\g_{i+1}$
we seek. By the construction this element satisfies the Leibniz rule and so belongs to $\g_1$.
In other words, $\Psi^2_X|_{\Delta_x\ot\Delta_x}$
belongs to the image of $\g_1$ in $\g_{-1}^*\ot\g_{-1}^*\ot\g_{-1}$.
For the general $i>0$ the arguments are the same (induction), and the reason behind the
claim of Lemma \ref{L3} is that both the element $\Psi^{i+1}_X$ and the elements of $\g_i$
are constructed on the same principle using the same formula (cf. introduction of
the spaces $\g_i$ for $i\ge0$ in \cite{T,Y}).
\end{proof}
Now we are ready for our main technical result.
\begin{theorem}\label{hm3}
For all $i\in\Z$: $\mathbf{g}_i\subset\g_i$. Thus we get the monomorphism of the graded Lie
algebras:
$$
\op{gr}(L^\Delta_x)=\oplus\mathbf{g}_i\hookrightarrow\oplus\g_i=\g_x
$$
\end{theorem}
\begin{proof}
Consider the map
$$
\Ll(x)^i\ni X\mapsto \Psi_X^{i+1}\in\g_i,\quad i\ge0.
$$
Its kernel is $\Ll(x)^{i+1}$ and therefore we get the induced monomorphism $\mathbf{g}_i\to\g_i$.
Existence of this arrow for negative $i$ is obvious.
Since the Lie bracket in both cases is induced by the commutator of vector fields
and by Lemma \ref{L2} it respects the filtration, so it respects the gradation and the map is a homomorphism of Lie algebras.
\end{proof}
Notice though that the formal Lie algebra $L^\Delta_x$ is not a Lie subalgebra of $\g_x$,
as can be seen by studying the sub-maximal examples in \cite{C$_1$}.
\begin{cor}
At any point $x\in M$ we have $\dim L^\Delta_x\le\dim\g_x$.
\end{cor}
Thus for the LAS $\Ll$ we achieved our claim. To prove it for
global/lo\-cal\footnote{This means over a small neighborhood, fixed for all vector fields.}
Lie algebra $\op{sym}(\Delta)$ we shall study the first grading too.
\section{Lie equation associated to a distribution}\label{S4}
We will use here jet-spaces and the geometric theory of PDE, for which we refer the reader to
\cite{Sp,KLV,KS,KL}. Consider the Lie equation $\E=\mathfrak{Lie}(K)$ of a geometric structure $K$.
For instance, if $K$ is a tensorial field, then this is the equation for its symmetries, i.e.
vector fields $X$ satisfying $L_X(K)=0$.
In the case of our current interest $K=\Delta$ and the Lie equation, considered as the
submanifold in jets, is
$$
\E=\{j^1_xX:L_X(\Delta)_x\subset\Delta_x\,|\, x\in M\}\subset J^1(TM).
$$
As is customary in the theory of overdetermined systems, finding compatibility conditions on the solutions
of $\E$ binds to applying the prolongation-projection method. Namely the prolongation
$\E^{(k-1)}\subset J^k$ is determined by the original equations and their derivatives up to order $k$.
It can happen that some projections $\pi_{k,l}:\E^{(k-1)}\to\E^{(l-1)}$ are not surjective
(i.e. there is a differential corollary of lower order),
then we take the image as the new system of equations and apply the prolongations and projections again.
The procedure is finite due to Cartan-Kuranishi theorem \cite{Ku} and the output is formally integrable,
meaning that it possesses a formal series solution through every regular point
(or local analytic solution for analytic $\E$).
Calculation of prolongation-projection in general is a difficult task, and $\mathfrak{Lie}(\Delta)$
is not an exception. Hopefully, we can guess an equation squeezed in between $\E$ and its $k$-th
derived $\pi_{k,1}(\E^{(k-1)})\subset J^1$. This is so because of the obvious fact that if $k$-jet of $X$
preserves $\Delta$ at $x$, then $j^1_xX$ preserves its weak derived flag up to order $k$.
Thus instead of studying symmetries of $\Delta$ we can equally well study symmetries of the derived
flag $\mathcal{W}_\Delta=\{\Delta_i\}_{i>0}$. Denote by $\E_\Delta$ the corresponding Lie equation
$$
\mathfrak{Lie}(\mathcal{W}_\Delta)=
\{j^1_xX:L_X(\Delta_i)_x\subset(\Delta_i)_x\,\forall i>0\,|\,x\in M\}\subset J^1(TM).
$$
Denote by $\bar\E$ the result of the prolongation-projection. This consists of the subsets
(submanifolds with singularities)
$$
\bar\E_i=\lim_{j\to\infty}\pi_{j+1,i}(\E_\Delta^{(j)})\subset J^i(TM).
$$
Symbols of this equation are the vector spaces (we use linearity of $\bar\E$
which simplifies the general formulae)
$$
\bar g_i=\op{Ker}(\pi_{i,i-1}:\bar\E_i\to\bar\E_{i-1}).
$$
Clearly these are subspaces of the symbols $g_i$ of the original Lie equation $\E_\Delta$:
\begin{equation}\label{II-3}
\bar g_i\subset g_i\subset S^iT^*_xM\ot T_xM.
\end{equation}
In particular, $\bar g_0$ is the tangent to the orbit of the symmetry group action
(the whole $T_xM$ in the transitive case).
Ultimately the infinite jets $\bar\E_\infty$ correspond to $L^\Delta_x$
(another descriptions is this: $\bar\E_\infty$ consists of formal vector fields
preserving all differential invariants of $\Delta$).
We shall relate the Spencer symbols $\bar g_{i+1}$ to the Tanaka symbols $\g_i$, $i\ge0$.
Consider the decreasing filtration $\Ll(x)^i\cap\Ll(x)^j_*$ of the space $\Ll(x)^i$.
It produces the grading (the isomorphisms below are not natural and respect only the
linear structure)
$$
\Ll(x)^i\simeq\bigoplus_{j\ge0}\frac{\Ll(x)^i\cap\Ll(x)^j_*}{\Ll(x)^i\cap\Ll(x)^{j+1}_*}=
\bigoplus_{j=0}^{i-1}\frac{\Ll(x)^i\cap\Ll(x)^j_*}{\Ll(x)^i\cap\Ll(x)^{j+1}_*}\oplus\bigoplus_{j\ge i}\bar g_{j+1}.
$$
This implies for $i\ge0$
\begin{equation}\label{BigO1}
\mathbf{g}_i=\Ll(x)^i/\Ll(x)^{i+1}\simeq\bigoplus_{j\le i}h_{ij},
\end{equation}
where $h_{ij}=(\Ll(x)^i\cap\Ll(x)^j_*)/(\Ll(x)^i\cap\Ll(x)^{j+1}_*+\Ll(x)^{i+1}\cap\Ll(x)^j_*)$.
Note that $h_{ij}=0$ if either $i<j$ or $i\ge(j+1)\kappa$ ($j\ge0$).
Similarly the decreasing filtration $\Ll(x)^i\cap\Ll(x)^j_*$ of the space $\Ll(x)^j_*$ yields
\begin{equation}\label{BigO2}
\bar g_{j+1}=\Ll(x)^j_*/\Ll(x)^{j+1}_*\simeq\bigoplus_{i\ge j}h_{ij}.
\end{equation}
As a by-product of calculations in the previous sections,
we can interpret the vector space $h_{ij}$ as a subspace in
$$
\sum_{s_1+{}\dots{}+s_{j+1}=i+t}\g_{-s_1}^*\ot\dots\ot\g_{-s_{j+1}}^*\ot\g_{-t}\quad(t,s_\nu\ge1).
$$
\begin{theorem}\label{hm4}
The Spencer symbols $\bar g_i$ of the equation $\bar\E$ are related to the
Tanaka symbols $\mathbf{g}_i$ via (\ref{BigO1})-(\ref{BigO2}).
This yields a (noncanonical) monomorphism of $\oplus_{j>0}\bar g_j$ into $\oplus_{i\ge0}\g_i$.
\end{theorem}
The claim of the theorem follows from the inclusions $\mathbf{g}_i(x)\subset\g_i(x)$ of Theorem \ref{hm3}.
These are strict and the only case, when we have equalities for all $x$ is the Tanaka flat
distribution (for finite-dimensional $\g$)\footnote{In this case Theorem \ref{hm4} gives two gradings
on the space of symmetries of the standard model $\Delta=\Delta_{\m}$, but only $\op{gr}(L^\Delta_x)$
yields the Lie algebra structure of $\op{sym}(\Delta)$.}.
Indeed, equalities everywhere mean that the Lie equation of the GNLA $\m_x$
is formally integrable, and the prolongation-projection does not decrease its symbols.
In particular, $\bar g_0(x)=T_xM$, the pseudogroup of symmetries is transitive, so that the distribution
is strongly regular and the result follows. This justifies Remark \ref{rk1}.
\begin{cor}
For a regular distribution $\Delta$ (not necessarily strongly regular)
$$
\sum_{i>0}\dim \bar g_i\le \sum_{j\ge0}\dim\g_j.
$$
\end{cor}
This implies that if the map $j_x^\infty:\Ll(x)^0\to J^\infty_x(TM)$ is injective
(we justify the assumption in the next section), then
$$
\dim\Ll(x)\le\sum_{j=-\kappa}^\infty\dim\g_j(x),
$$
and so $\dim\Ll\le\sup_M\sum_j\dim\g_j$.
\section{Proof of Theorem \ref{Thm1} and beyond}\label{S4.5}
We can suppose that $\dim\g_x$ is finite at every point $x\in M$, because else the inequality
in Theorem \ref{Thm1} is trivial.
\begin{lem}
The function $x\mapsto\dim\g_x$ is upper semi-continuous.
\end{lem}
\begin{proof}
Indeed, $\g_x$ is obtained from $\m_x$ by certain linear algebra rules
(in \cite{T,Y} the positive grades $\g_i$ are defined successively, but this can be easily
modified to obtain $\g_+$ via $\m$ at once).
Since ranks of matrices can only drop in the limit process, the result follows.
\end{proof}
Thus $\dim\g_x$ attains a maximum in any compact domain $\bar U$.
Let $N$ be the number such that $\g_i(x)=0$ for all $i\ge N$ and $x\in\bar U$.
Then (\ref{BigO1})+(\ref{BigO2}) imply that $\bar g_i(x)=0$ for
all $i>N$ and $x\in\bar U$. In fact, the whole prolongation-projection process
is not required, but as the condition that $X$ preserves the GNLA $\m_x$ structure is obtained
from the original $\mathfrak{Lie}(\Delta)$ in a finite number of steps and
$\g_x$ is obtained via $\m_x$ by algebraic prolongation we conclude:
In finite number of steps of prolongation-projection the Lie equation becomes
of finite type at all points $x\in\bar U$.
Consequently by the results of Theorem \ref{T-app} from the Appendix,
there are no symmetries in $\bar U$ flat at some point.
Since $\bar U\subset M$ is arbitrary, we conclude the result for the whole $M$
and hence the map
$$
j_x^\infty:\op{sym}(\Delta)\to\bar\E_x^\infty
$$
is injective for every point $x\in M$.
Thus results of Section \ref{S4} imply the inequality
$$
\dim\op{sym}(\Delta)\le\dim\Ll(x)\quad \forall x\in M
$$
and consequently we prove inequality (\ref{ineq-rk2}) of Remark \ref{rk2}, which implies in
turn Theorem \ref{Thm1}.
\qed
\begin{rk}
If the distribution is not strongly regular, then the leaves on $M$ through a typical point $x$
(obtained by fixing the invariants) have codimension $r=n-\op{rank}[\op{ev}_x]>0$ and we
refine inequality (\ref{ineq-rk2}) to
\begin{equation*}
\dim\op{sym}(\Delta)\le\op{inf}_M\dim\g_x-r.
\end{equation*}
\end{rk}
\noindent{\bf Proof of Corollary \ref{cor}.}
We use Corollary 2 of Theorem 11.1 from \cite{T}, which states that if the subalgebra
$$
\h_0=\{v\in g_0:[v,g_r]=0\ \forall r<-1\}
$$
is of finite type as the subalgebra of $\op{gl}(\g_{-1})$, then $\g$ is finite-dimensional.
The subalgebra $\h_0\subset\op{gl}(\g_{-1})$ is of finite type iff its
complex characteristic variety is empty \cite{GQS}.
Thus all Tanaka algebras $\g_x$, $x\in M$, are finite-dimensional and the result follows.
\qed
\medskip
We can introduce the characteristic variety (the set of covectors satisfying the
defining relation is homogeneous and we projectivize it)
$$
\op{Char}(\Delta)=\mathbb{P}\{p\in g_{-1}^*\setminus\{0\}:\exists q\in g_{-1}\setminus\{0\},\ p\ot q\in \mathfrak{h}_0\}
\subset \mathbb{P}\Delta^*.
$$
Working over $\CC$ (this part of the theory is algebraic, and
complexification makes no problem) we obtain $\op{Char}^\CC(\Delta)\subset\mathbb{P}^\CC\Delta^*$.
The criterion of the theorem reformulates now as follows:
$$
\op{Char}^\CC(\Delta)=\emptyset\ \Longrightarrow\ \dim\op{sym}(\Delta)<\infty.
$$
If the set $\op{Char}^\CC(\Delta)$ is empty, then the complex characteristic variety of the differential
closure $\bar\E$ of the Lie equation $\E_\Delta$ is empty as well, but the reverse implication
is not generally true though it holds for a Tanaka flat distribution $\Delta$.
\begin{rk}
One should be cautious as $\op{Char}^\CC(\bar\E)$ is usually smaller than
the initial complex characteristic variety $\op{Char}^\CC(\E_\Delta)$.
\end{rk}
{\bf Example.} Consider the Tanaka flat rank 2 distribution in $\R^6$
of growth $(2,1,2,1)$ and maximal symmetry algebra $\mathfrak{p}_6$ of dimension 11.
In \cite{AK} (see also \cite{DZ}) it was shown that it corresponds to the Monge equation
$$
\E_{1,3}:\ y'=(z''')^2,
$$
namely $\Delta=\langle\D_x=\p_x+z_3^2\p_y+z_1\p_z+z_2\p_{z_1}+z_3\p_{z_2},\p_{z_3}\rangle$
in the standard jet-coordinates in mixed jets $J^{1,3}(\R,\R\times\R)\supset\E_{1,3}\simeq\R^6(x,y,z,z_1,z_2,z_3)$.
Moreover its algebra of symmetries $\op{sym}(\Delta)$ is isomorphic to the
Tanaka algebra $\g=\mathfrak{p}_6$ with $\g_0\simeq\R^3$, $\g_1\simeq\R^2$, $\g_2=0$,
and it has the following basis corresponding to elements of pure grade in $\g$:
\begin{align*}
\g_{-4}:\quad & Z_0=\p_z,\\
\g_{-3}:\quad & Z_1=x\p_z+\p_{z_1},\ \ Y_0=\p_y,\\
\g_{-2}:\quad & Z_2=\tfrac{x^2}2\p_z+x\p_{z_1}+\p_{z_2},\\
\g_{-1}:\quad & Z_3=\tfrac{x^3}{3!}\p_z+\tfrac{x^2}2\p_{z_1}+x\p_{z_2}+\p_{z_3}+2z_2\p_y,\ \ S_0=\p_x\\
\g_0:\quad & Z_4=\tfrac{x^4}{4!}\p_z+\tfrac{x^3}{3!}\p_{z_1}+\tfrac{x^2}2\p_{z_2}+x\p_{z_3}+2(xz_2-z_1)\p_y,\\
& S_1=x\p_x+\tfrac52z\p_z+\tfrac32z_1\p_{z_1}+\tfrac12z_2\p_{z_2}-\tfrac12z_3\p_{z_3},\\
& R=y\p_y+\tfrac12z\p_z+\tfrac12z_1\p_{z_1}+\tfrac12z_2\p_{z_2}+\tfrac12z_3\p_{z_3},\\
\g_1:\quad & Z_5=\tfrac{x^5}{5!}\p_z+\tfrac{x^4}{4!}\p_{z_1}+\tfrac{x^3}{3!}\p_{z_2}+\tfrac{x^2}2\p_{z_3}+
2(\tfrac{x^2}2z_2-xz_1+z)\p_y,\\
& S_2=x^2\p_x+9z_2^2\p_y+5xz\p_z+(5z+3xz_1)\p_{z_1}+\\
& \qquad\qquad\qquad\qquad\qquad +(8z_1+xz_2)\p_{z_2}+(9z_2-xz_3)\p_{z_3}.
\end{align*}
If we take the classes of these fields in $\Ll(x)^i_*/\Ll(x)^{i+1}_*$ we get:
$$
\bar g_0=\langle[S_0],[Y_0],[Z_0],[Z_1],[Z_2],[Z_3]\rangle,\ \
\bar g_1=\langle[S_1],[S_2],[R],[Z_4],[Z_5]\rangle.
$$
However this corresponds to the differential closure (via prolongation-projection) $\bar\E_\Delta$,
while the original Lie equation $\E_\Delta=\mathfrak{Lie}(\Delta)$ is bigger. In particular,
the element $[Z_5]=dz\ot\p_y\in\bar g_1\subset g_1$ is of rank 1. Further prolongation-projections
of $\E_\Delta$ yield $\bar g_2=0$ and kill this characteristic.
\section{Finite-dimensionality of the symmetry algebra}\label{S5}
Theorem \ref{Thm2} for $n=2$ was essentially established in \cite{AK}
(but in that paper we restricted to the strongly regular case)
by showing that $\h_0=0$ provided
the growth vector is $(2,1,2,\dots)$, and even in a more general case $(2,1,\dots,1,2,\dots)$.
Another proof is as follows. We can work over $\CC$ as this does not change the dimensions
of graded components. Suppose there is a nonzero element of rank 1
$\oo=p\ot\zeta\in\h_0\subset\g_{-1}^*\ot\g_{-1}$. Let $\xi$ be a complement to $\zeta$ in $\g_{-1}$.
Then $\oo$ acts trivially on $\g_{-2}=\R\cdot[\zeta,\xi]$ iff $p(\zeta)=0$. Furthermore
$\oo$ acts trivially on $\g_i$, $i<-2$, iff the operator
$\op{ad}_\zeta:\g_{i+1}\to\g_i$ is zero. Consequently $\dim\g_i=1$ for $i<-1$.
\smallskip
\noindent{\bf Proof of Theorem \ref{Thm2} for $n>2$.}
Again, working over $\CC$ and taking an element
of rank 1 $\oo=p\ot\zeta\in\h_0\subset\g_{-1}^*\ot\g_{-1}$ we observe that $\oo$ acts trivially
on $\g_{-2}$ iff $\op{ad}_\zeta|_{\op{Ann}(p)}=0$. As $\zeta$ can belong to $\op{Ann}(p)$,
this imposes $(n-2)$ restrictions.
Thus in the case the characteristic variety is non-empty, dimension of $\g_{-2}=\op{ad}(\La^2\g_{-1})$
does not exceed $\frac{n(n-1)}2-(n-2)=\frac{(n-1)(n-2)}2+1$.
Thus distributions with growth starting $(3,3,..)$, $(4,5,..)$, $(4,6,..)$ etc
have finite-dimensional symmetry algebras. \qed
\medskip
The a-priory knowledge of finite-dimensionality of the symmetry algebra can be enhanced
by estimation of its maximal size in many cases. For distributions of rank 2 it is done in \cite{DZ,AK}.
It can be also done for distributions $\Delta$ with free truncated GNLAs. Assume that
$\m_x$ is the free Lie algebra of step $k$, i.e. the only constraints are the Jacobi identity
and the requirement $\g_{-k-1}=0$. Then obviously $\g_0=\g_{-1}^*\ot\g_{-1}$. Moreover
by \cite{W} $\g_1=0$ provided $k>2$, $n>2$ or $k>3$, $n=2$. This implies
(\cite{W} concerned only homogeneous distributions, but the result holds for any $\Delta$
with truncated free $\m$ due to \cite{T} or Theorem \ref{Thm1}):
$$
\dim\op{sym}(\Delta)\le \ell_n(k)+n^2
$$
Here $\ell_n(k)=\dim\m_x$ can be calculated recursively or via the M\"obius function by \cite{Se}
$$
k\,\ell_n(k) \,=\, n^k-\!\!\!\sum_{m|k,m<k}\!\! m\,\ell_n(m)
\,=\,\sum_{m|k}\mu(m)\,n^{k/m}.
$$
For the exceptional cases we have:
$\dim\op{sym}(\Delta)\le 2n^2+n$, provided $k=2$, $n>2$
and the maximal symmetric case is given by $B_n=\mathfrak{o}(2n+1)$ \cite{Y,CN}.
For $k=3$, $n=2$ $\dim\op{sym}(\Delta)\le14$, the maximal dimension being realized only
in the case of $\op{Lie}(G_2)$ \cite{C$_1$}.
\smallskip
The estimate of Theorem \ref{Thm2} is sharp as there exist infinite-dimensional Lie algebras acting
as symmetries on distributions with growth vector beginning $(2,1,1,..),$ $(3,1,..)$, $(3,2,..)$, $(4,1,..)$,
$(4,2,..)$, $(4,3,..)$, $(4,4,..)$ etc, as discussed in the next Section.
\section{Infinite algebras of symmetries}\label{S8}
As in the rest of the paper we assume $\Delta$ totally non-holonomic. Notice however that
if the bracket-closure $\Delta_\infty\ne TM$, and $\mathcal{F}$ are the leaves of $\Delta_\infty$,
then existence of one symmetry transversal to $\mathcal{F}$ implies existence of
an infinite-dimensional space of such symmetries.
\subsection{Distributions of rank $n=2$}\label{SS81}
If a totally non-holonomic distribution with $\op{rank}(\Delta)=2$
has infinite-dimensional symmetry algebra then its growth vector starts $(2,1,1,\dots)$.
Recall that a Cauchy characteristic of a distribution is a symmetry tangent to this distribution.
The following statement is due to E.\,Cartan \cite{C$_2$} (see \cite{AK} for another proof;
alternatively the vector field $\zeta$ is the one constructed in
the beginning of Section \ref{S5}).
\begin{theorem}\label{ThDeProl}
The growth vector of a rank 2 distribution $\Delta$ is $(2,1,1,\dots)$
if and only if there exists a vector field $\zeta\in\Gamma(\Delta)$ which is a
Cauchy characteristic for the distribution $\Delta_2$.
In this, and only in this case, $\Delta$ is locally the prolongation of another
rank 2 distribution $\bar\Delta$.
\end{theorem}
Here the (geometric) prolongation of a rank 2 distribution $\bar\Delta$ on $\bar M$ is
the projectivization $M=\mathbb{P}\bar\Delta=\{l_x\subset\bar\Delta_x:x\in\bar M\}$
with natural projection $\pi:M\to\bar M$ and the distribution $\Delta=\pi^{-1}_*(l_{\pi(x)})$.
Locally in $\bar M$ for $\bar\Delta=\langle U,V\rangle$ and $t\in S^1=\R^1/2\pi\Z$ we have:
$M=\bar M\times S^1$ and $\Delta=\langle \cos t\cdot U+\sin t\cdot V,\p_t\rangle$.
If the distribution $\bar\Delta$ has a preferred (vertical) section
$V\in\Gamma(\bar\Delta)$ [this means that the symmetries preserve it],
then the (affine) prolongation $\Delta=\bar\Delta^{(1)}$
is simply $\Delta=\langle U+t\,V,\p_t\rangle$ on $M=\bar M\times\R^1$, where $t$ is the coordinate on $\R^1$.
This coincides with (Spencer) prolongation in the geometric theory of PDEs (\cite{KLV,MZ}).
When $\Delta=\bar\Delta^{(1)}$, the operation $\Delta\mapsto\bar\Delta$ is called de-prolongation.
In the case of Theorem \ref{ThDeProl},
de-prolongation is the quotient of the first derived distribution by the Cauchy characteristic
$$
(\bar M,\bar\Delta)=(M,\Delta_2)/\zeta.
$$
The algebras $\sym(\bar\Delta)$ and $\sym(\bar\Delta^{(1)})$ are isomorphic
[indeed any symmetry on $\bar M$ induces the action on $t$ and thus lifts to $M$]. Therefore
Theorem \ref{Thm2} and a sequence of de-prolongations yield the following important statement
(it gives a-posteriori transitivity of the symmetry group action; with transitivity imposed
a-priori -- as an additional assumption -- the claim follows from Theorem 7.1 of \cite{MT}).
\begin{theorem}
A regular germ of a rank 2 non-holonomic distribution $\Delta$ has infinite-dimensional symmetry
algebra if and only if $\Delta$ is equivalent to the Cartan distribution $\mathcal{C}_k$
on jet-space $J^k(\R,\R)$, $k=\dim M-2$.
\end{theorem}
Recall that distributions $\Delta$ with strong growth vector $(2,1,1,\dots,1)$ are called
Goursat distributions. Their regular points can be characterized by the condition that
the growth vector is $(2,1,1,\dots,1)$ in both the weak and the strong sense.
Near such points the normal form
$$
\mathcal{C}_k=\op{Ann}\{dy_i-y_{i+1}\,dx|0\le i<k\}\subset TJ^k(\R,\R)
$$
is provided by the von Weber - Cartan theorem \cite{W,C$_2$}, see also \cite{GKR}.
\subsection{Distributions of higher rank}\label{SS82}
We indicate a local construction to produce a distribution with infinite-dimensional
(intransitive) symmetry algebra from any distribution.
For simplicity let's start with the case, when we extend a rank 2 distribution.
Locally any such distribution can be represented as the Cartan
distribution $\C_k$ for the Monge system, i.e. an underdetermined ODE $\E\subset J^k(\R,\R^m)$
given by $(m-1)$ equations.
Then we define $M=\E\times_\R J^l(\R,\R)\subset J^{k,l}(\R,\R^m\times\R)$ and
the distribution is $\Delta=(\C_k|_\E)\times_\R\C_l$.
In local coordinates if the equations in $\E$ are
$u^i_k=\psi^i(x,u^j,\dots,u^j_{k-1},v,\dots,v_k)$, $i,j=1,\dots,m-1$
($u^i=u^i(x)$, $v=u^m(x)$ are unknowns, $u^j_1=u^j_x$, $v_1=v_x$, $v_2=v_{xx}$ etc),
then $\C_k|_\E=\langle\D_x'=\p_x+u^i_1\p_{u^i}+\dots+\psi^i\p_{u^i_{k-1}}
+v_1\p_v+\dots+v_k\p_{v_{k-1}},\p_{v_k}\rangle$ and
$\C_l=\langle\D_x''=\p_x+w_1\p_w+w_2\p_{w_1}+\dots,\p_{w_l}\rangle$, so that
$$
\Delta=\langle\D_x=\p_x+u^i_1\p_{u^i}+v_1\p_v+w_1\p_w+\dots,\p_{v_k},\p_{w_l}\rangle.
$$
It is obvious that prolongation $\hat X$ of the vector field $X=f(w)\p_w$ is a symmetry
of $\Delta$ for any function $f(w)$.
Similarly a general regular rank $n$ distribution is realized locally as the Cartan distribution
$\C_k|_\E$ of an overdetermined PDE system $\E\subset J^k(\R^s,\R^m)$.
Then the distribution $\Delta=(\C_k|_\E)\times_{\R^s}\C_l$ on
$M=\E\times_{\R^s}J^l(\R^s,\R)\subset J^{k,l}(\R^s,\R^m\times\R)$ has an infinite algebra of
symmetries.
We can shrink the 2nd factor to a symmetric equation $\mathcal{R}\subset J^l(\R^s,\R)$,
for instance taking $\mathcal{R}=J^l(\R,\R)$ given by the equations $w_{x_2}=\dots=w_{x_s}=0$,
and still have infinitely many symmetries $\hat X$, $X=f(w)\p_w$ for the distribution
$\Delta=(\C_k|_\E)\times_{\R^s}(\C_l|_\mathcal{R})$.
\begin{rk}
For $l=0$ we get extension of the distribution via the Cauchy characteristic, i.e. locally
$(M,\Delta)=(\bar M,\bar\Delta)\times(\R,\R)$.
\end{rk}
This construction allows realizing all cases not prohibited by
Theorem \ref{Thm2} as distributions with infinite-dimensional symmetry algebras.
For rank 3 distributions all infinite primitive symmetry algebras
that occur are Lie transformations\footnote{In their natural representation
$k=0$ and $k=0,1$ in the next cases to be primitive,
but the algebra does not change with $k$ due to Lie-B\"acklund theorem.} of $J^k(\R,\R^2)$,
while for rank 4 there appear new real primitive algebras from the list of \cite{MT}:
the symmetries of the Cartan distribution of $J^k(\R,\R^3)$, $J^1(\R^2,\R)$ and $J^k(\CC,\CC)_\R$.
\subsection{Distributions of rank $n=3$}
Let us describe in more details the case of $\op{rank}(\Delta)=3$.
If the growth vector is $(3,1,..)$, then the bracket
$\La^2\g_{-1}\to\g_{-2}$ has a kernel $v\in\g_{-1}$, and this corresponds to a Cauchy characteristic
vector field, so that $(M,\Delta)\simeq(\bar M,\bar\Delta)\times(\R,\R)$.
Consider the growth $(3,2,..)$. Then $\Delta$ contains a rank 2 sub-dis\-trib\-ution $\Pi$, with
$\Pi_2\subset\Delta$. If $\Pi_2=\Delta$, then $\op{sym}(\Delta)=\op{sym}(\Pi)$. The growth vector of
$\Pi$ starts $(2,1,2,..)$ and so $\op{sym}(\Pi)$ is finite-dimensional.
Thus infinite algebras $\op{sym}(\Delta)$ correspond to integrable $\Pi$. In this case
the characteristic variety is real and $\op{Char}^\CC=\op{Char}$ equals the one
point set $\mathbb{P}(\Pi^\perp)$, where $\Pi^\perp$ is the annihilator of $\Pi\subset\g_{-1}$.
Moreover let $p\in\Pi^\perp\cap\g_{-1}^*\setminus0$. Then the kernel bundle over characteristic variety
$\mathcal{K}=\{q\in\g_{-1}:p\ot q\in\h_0\}$ is either 1- or 2-dimensional subspace of $\Pi$.
\smallskip
{\bf I:} $\dim\mathcal{K}=1$. Here $\op{sym}(\Delta)$ is the symmetry of the flag
$(\mathcal{K},\Pi,\Delta)$. Moreover this triple extends (locally) to a complete flag
$\mathfrak{F}$ of subspaces $(\mathcal{K},\Pi,\Delta,[\mathcal{K},\Delta],\Delta_2,\dots)$
in $TM$, invariant under $\op{sym}(\Delta)$.
\begin{rk}
For $n=2$ the infinite symmetry algebra also leaves invariant the complete flag
$\mathfrak{F}=(\langle\zeta\rangle,\Delta,\Delta_2,\dots)$, see Section \ref{SS81}.
\end{rk}
Now $\mathcal{K}$ is the Cauchy characteristic space of the distribution $\Delta^\dag=[\mathcal{K},\Delta]$,
so we can pass to the (local) quotient $\bar M=M/\mathcal{K}$, $\bar\Delta=\Delta^\dag/\mathcal{K}$.
Any symmetry of $\Delta$ descends to a symmetry of $\bar\Delta$, and the latter contains a
preferred direction $\Pi/\mathcal{K}$. It is not necessarily the kernel direction $\bar{\mathcal{K}}$ of $\bar\Delta$, but it belongs to the corresponding 2-distribution $\bar\Pi$.
The passage $\Delta\mapsto\bar\Delta$ is de-prolongation of rank 3 distributions in the same sense as
in Section \ref{SS81}, namely this operation is inverse to the prolongation defined as follows
(this applies only for $\op{Char}(\Delta)\ne\emptyset$, so that
$\Pi\subset\Delta$ is integrable etc).
\underline{Prolongation I}$_a$. This works in the case $\bar\Delta$ has a preferred direction
$\ell=\langle Y\rangle$ different from $\bar{\mathcal{K}}$
[if not, then the Lie algebra $\op{sym}(\bar\Delta)$ shrinks to the stabilizer].
Namely $M$ is the space of all 2-planes $P\subset\bar\Delta$ through $\ell$ transversal to $\bar{\mathcal{K}}$
(this is the affine version; compact version is without transversality).
If $\pi:M\to\bar M$ is the projection, then $\Delta=\pi_*^{-1}(P)$.
Locally if $\bar\Delta=\langle X,Y,Z\rangle$ with $X\in\bar{\mathcal{K}}$ and $Y\in\ell\subset\bar\Pi$,
then $M=\bar M\times\R(t)$, $\Delta=\langle Y,\p_t,Z+tX\rangle$.
In these notations $\Pi=\langle Y,\p_t\rangle$, $\p_t\in\mathcal{K}$.
This is inverse to the above de-prolongation when $\dim\mathcal{K}=1$.
\underline{Prolongation I}$_b$. $\bar{\mathcal{K}}$ is clearly a preferred direction,
so we can prolong along it: $M$ is the space of all 2-planes $P\subset\bar\Delta$ through
$\bar{\mathcal{K}}$ different from $\bar{\Pi}$
(the affine version; compact version - no conditions).
Again $\Delta=\pi_*^{-1}(P)$ and locally $\Delta=\langle X,\p_t,Z+tY\rangle$. Here
$\Pi=\langle X,\p_t\rangle=\mathcal{K}$, so this is not inverse to the above de-prolongation.
\smallskip
{\bf II:} $\dim\mathcal{K}=2$. Here $\Pi=\mathcal{K}$ is the space of Cauchy characteristics
for rank 5 distribution $\Delta_2$, and we can reduce $(M,\Delta)$ to a rank 3 distribution
$\bar\Delta=\Delta_2/\Pi$ on a lower-dimensional manifold $\bar M=M/\Pi$.
This is de-prolongation in the same sense as in Section \ref{SS81}, while the prolongation
($\Pi$ integrable) mimics the Spencer prolongation on $J^k(\R,\R^2)$.
\underline{Prolongation II}. Here $M$ is the space of lines $L\subset\Delta$ transversal to
$\Pi$ (the affine version; compact version - no restrictions), $\pi:M\to\bar M$ the natural projection,
then $\Delta=\pi_*^{-1}(L)$.
Locally if $\bar\Delta=\langle X,Y,Z\rangle$ with $\Pi=\langle X,Y\rangle$, then $M=\bar M\times\R^2(u,v)$
and $\Delta=\langle \p_u,\p_v,Z+uX+vY\rangle$. Clearly $\mathcal{K}=\langle\p_u,\p_v\rangle$.
\smallskip
Notice that de-prolongations in both cases induce the injective map
$\pi_*:\op{sym}(\Delta)\hookrightarrow\op{sym}(\bar\Delta)$
[prolongations I$_a$ can shrink the symmetry algebra, though in a controllable way].
Thus de-prolongations can be applied until the distribution gets a Cauchy characteristic
or its length (degree of non-holonomy) becomes $\kappa=1$.
Thus we obtain the following normal form [for $n=3$ infinitely-symmetric models can have
functional moduli, but these are gone for most symmetric models in the sense of functional dimension].
\begin{theorem}
A regular germ of a rank 3 non-holonomic distribution $\Delta$ has infinite-dimensional symmetry
algebra if and only if $\Delta$ is obtained from the product $\Pi\times\R$, where
$\Pi$ is some rank 2 distribution, by the operations of prolongations (of type I or II).
\end{theorem}
Here we include the case when $\Pi$ is the trivial distribution $(\R^2,\R^2)$, in which case prolongation II
of $\Pi\times\R$ yields $J^k(\R,\R^2)$.
\smallskip
{\bf Example I.} Consider the Monge equation $\E_{m,n}=\{y_m=z_n^2\}$ as a submanifold
in the mixed jets (subscripts count the derivatives by $x$)
$$
J^{m,n}(\R,\R^2)\simeq\R^{m+n+3}(x,y,y_1,\dots,y_m,z,z_1,\dots,z_n)
$$
and let $M=\E_{m,n}\times_\R J^l(\R,\R)\subset J^{m,n,l}(\R,\R^3)$
with the natural rank 3 distribution
$\Delta=\mathcal{C}_{m,n}\times_\R\mathcal{C}_l=\langle\D_x,\p_{z_n},\p_{w_l}\rangle$,
$$
\D_x=\p_x+y_1\p_y+{}\dots{}+z_n^2\p_{y_{m-1}}+z_1\p_z+{}\dots{}
+z_n\p_{z_{n-1}}+w_1\p_w+{}\dots{}+w_n\p_{w_{l-1}}.
$$
De-prolongation I results in
$\bar M=\R^{m+n+l+2}(x,\{y_i\}_0^{m-1},\{z_i\}_0^n,\{w_i\}_0^{l-1})$
with $\bar\Delta=\langle\D_x,\p_{z_n},\p_{w_{l-1}}\rangle$.
Further de-prolongations I bring the distribution to
$\hat\Delta=\langle\D_x,\p_{z_n},\p_w\rangle=\mathcal{C}_{m,n}\times\R$ on
$\E_{m,n}\times\R(w)$.
The general symmetry of this distribution is $X_0=Y_\E+f\cdot\p_w$, where $Y_\E$ is
the general symmetry of $\E_{m,n}$, $m\le n$ (it depends on $2n+5$ parameters for $m=1$, $n>2$
or $2n+4$ parameters for $m>1$, see \cite{AK}) and $f\in C^\infty(\E_{m,n}\times\R)$.
In order for this symmetry to allow prolongation I$_a$, it shall preserve the vertical
line $\mathfrak{v}=\langle\p_{z_n}\rangle$. This is so for the first component $Y_\E$
\cite{AK}, and it holds for $f\cdot\p_w$ iff $f_{z_n}=0$. The prolongation is
$X_1=Y_\E+f\cdot\p_w+\D_x(f)\cdot\p_{w_1}$. Now this preserves $\mathfrak{v}$ iff
$f_{z_{n-1}}=0$, $f_{y_{m-1}}=0$, and the prolongation is
$X_2=Y_\E+f\cdot\p_w+\D_x(f)\cdot\p_{w_1}+\D_x^2(f)\cdot\p_{w_2}$.
Continuing in this way we obtain that the symmetries of $(M,\Delta)$ are
$X_l=Y_\E+\sum_{k\le l}\D^k_x(f)\cdot\p_{w_k}$ and $f=f(x,\{y_i\}_0^{m-l},\{z_i\}_0^{n-l},w)$
is any smooth function of $2+\max\{0,m-l+1\}+\max\{0,n-l+1\}$ arguments.
\smallskip
{\bf Example II.} For the mixed jets $J^{m,n}(\R,\R^2)=J^m(\R,\R)\times_\R J^n(\R,\R)$ with
canonical Cartan distribution $\mathcal{C}_{m,n}$ of rank 3, de-prolongation II reduces it to
$J^{0,n-m}(\R,\R^2)=\R\times J^{n-m}(\R,\R)$, $m\le n$. The infinite algebra of symmetries
for this model is described in \cite{AK}.
\smallskip
Some other results on distributions with infinitely many symmetries
(on graded nilpotent Lie groups) can be found in \cite{DR,OR}.
|
\section{Introduction}
Questions concerning the behavior of spacetime and of quantum fields at short distances are intimately connected \cite{ISH-strucQG}. In the presence of gravity, usual quantum field theory (QFT) on a fixed background cannot be a complete theory because the effect on spacetime curvature of sufficiently high energy particles cannot be ignored. This problem is already manifest if we consider the gravitational effect of the QFT vacuum. This vacuum is the sum
of ground state oscillator energies, one for each point in momentum space, and so adds up
to quite a bit, namely, the cosmological constant problem.
In addition to this ``static'' problem, the QFT vacuum is considered to be filled with virtual particle-antiparticle pairs. The energy density of such a pair is its rest mass $2m$ divided by the Compton volume
\begin{equation*}
\rho_{\rm pair}^{\rm rest} = \frac{2m}{[\hbar/(mc)]^3} \simeq m^4.
\end{equation*}
If the virtual pair interacts gravitationally, applying Newton's law gives an enhanced interaction energy density
\begin{equation*}
\rho_{\rm pair}^{\rm int}= \frac{Gm^2}{\hbar/(mc)}\ \frac{1}{[\hbar/(mc)]^3} \simeq m^6.
\end{equation*}
These energy densities induce virtual curvature of spacetime via the Einstein equations.
Thus, it is apparent from simple arguments that high energy probes would drastically affect spacetime and make the semiclassical approximation inapplicable.
There have been numerous attempts to modifying the high energy behavior of quantum fields to
overcome the problem sketched above. All of them have the feature that an ultraviolet cut-off is introduced
by some argument that leads to well behaved propagation of quantum fields at high energies. In addition, all attempts involve the introduction of a fundamental length motivated by quantum gravity, whether it is string theory, loop quantum gravity (LQG) \cite{lqg-AL, lqg-T, lqg-S,lqg-R}, non-commutative geometry \cite{DFR} or some other
approach independent of these.
In this paper we revisit the question about the high energy behavior of quantum fields. We explore not
so much a specific theory but a quantization procedure that, unlike the Schr\"odinger quantization, comes with a specified mass scale. This is the so-called polymer or background independent quantization method that originated in LQG. Its central feature for our purposes is that {\it the Hilbert space used for the quantization has an inner product, which is independent of the background metric}, hence the name.
This property does not, however, imply that the expectation values of all operators are metric independent; there could still be a dependence on the metric through the operators themselves (such as the Hamiltonian), or possibly through a special quantum state in which metric dependence arises through coefficients in a linear combination. We consider here expectation values of operators in a suitably defined vacuum state. We will see that they turn out to be metric independent. Furthermore, the mode frequency and the mass scale associated with the quantization give rise to a dimensionless parameter. Expectation values can then be expressed in terms of this parameter, thereby allowing us to probe their high energy behavior.
A similar approach to the one we employ here has been applied to black holes and cosmology \cite{HW-cosm,
HW-qbh1, HW-qbh2, HHS-cosm}, and it has been used to derive an effective scalar field theory on a Minkowski background \cite{HHS-sft} using semiclassical states. The present work may be viewed as a parallel but independent development.
The outline of this paper is as follows. In Sec. II we give the quantization procedure, together with a
definition of Fock states as well as creation and annihilation operators. This construction is new and the way it fits in the context of earlier work on the relation between Fock and polymer quantization is discussed. In Sec. III we give a computation of the expectation values of the commutator and anti-commutator for our Fock-like states. Unlike in usual QFT, these acquire a frequency dependence. The main result of his section is that the expectation values behave like the usual Fock states at low energies but show a surprising fermionic nature at high energies. Section IV contains a discussion of the results and some open directions and speculations.
\section{Classical theory and polymer variables}
Although our primary interest is the free massless scalar field theory on the
Minkowski spacetime, let us first consider the more general metric
\begin{equation*}
g_{\alpha\beta} \mathrm{d}x^{\alpha}\mathrm{d}x^{\beta} = -\mathrm{d}t^2 + q_{ab}\mathrm{d}x^a \mathrm{d}x^b ,
\end{equation*}
where $q_{ab}$ denotes
the spatial metric. The canonical phase space variables are
$(\phi,p)$ and the scalar matter Hamiltonian is given by
\begin{equation*}
H = \int_V\frac{\sqrt{q}}{2} \left( \frac{p^2}{q} +
q^{ab}\partial_a \phi\partial_b\phi + m^2\phi^2 \right)\mathrm{d}^3x,
\end{equation*}
where $ q = \det(q_{ab})$. The usual (Klein-Gordon) wave equation for the scalar field
follows from this via Hamilton's equations of motion.
Let us consider the phase space functions
\begin{equation}\label{basic-obs}
\begin{split}
\Phi_f (t,\mathbf{x}) &= \frac{1}{V} \int_V \sqrt{q} f(\mathbf{x}-\mathbf{x}') \phi(t,\mathbf{x}')\ \mathrm{d}^3x', \\
U(t,\mathbf{x}) &= \mathrm{e}^{i\lambda p(t,\mathbf{x})/\sqrt{q}}
\end{split}
\end{equation}
as the new set of variables that are to be realized as
the basic operators in the quantum theory. The function $f$ is scalar valued,
$\lambda$ is a real constant with the dimension of length
squared (in natural units), and $V$ is the volume of $3-$space.
The factors of $\sqrt{q}$ are necessary to balance density weights in
the integral and in the exponent. The variables in Eq. \eqref{basic-obs} satisfy the (equal time)
canonical Poisson bracket
\begin{equation}
\{ \Phi_f(t,\mathbf{x}), U(t,\mathbf{x}') \} = i \frac{\lambda}{V} f(\mathbf{x}-\mathbf{x}')
U(t,\mathbf{x}'). \label{basicpb}
\end{equation}
These variables may be viewed as ``dual" to
those used in the polymer quantization of a particle system
\cite{AFW} motivated by LQG,
where the exponentiated configuration variable is used as the new
variable. The dual for quantum mechanical systems has been discussed
in \cite{halvor}. We note, however, that there are important
differences with the variables used for the polymer scalar field quantization
discussed in \cite{QSD4,MV-fock, ALS-fock,Sahl-sft}, which do not use the available
background metric in their definition. The functions we employ are in a sense more conventional
in that they are just the field configuration and field translation variables.
As we will see, the essential difference to the standard quantization lies in the
realization of the field momentum operator.
A localized field may be defined by taking, for example, $f$ to be
a Gaussian, which is sharply peaked at a point $\mathbf{x}_j$. This means that
\begin{equation}\label{gauss}
f(\mathbf{x}_j-\mathbf{x}')=\mathrm{e}^{-(\mathbf{x}_j-\mathbf{x}')^2/\sigma^2} ,
\end{equation}
%
where $\sigma^2\ll 1$. We will assume
this in the following.
Since our main interest in this paper is in QFT on flat spacetime, we use the Euclidean metric $e_{ab}$ and set $q_{ab} = e_{ab}$ from this point on. The generalization to arbitrary metrics is
straightforward and is of potential use for cosmology and black hole
physics.
\section{Quantization}
The representation of the functions defined in Eq. \eqref{basic-obs} as operators on a
suitable Hilbert space is related to that used in polymer quantum mechanics \cite{AFW,
halvor}. We therefore briefly review this quantization before generalizing
it to field theory. As already mentioned, a central feature of this quantization
is that it introduces a length scale in addition to $\hbar$.
The Hilbert space is the space of almost
periodic functions, where a wave function is written as the linear
combination
\begin{equation*}
\psi(\mathbf{p}) = \sum_{j=1}^{N} c_j \mathrm{e}^{i\mathbf{p}\mathbf{x}_j} = \sum_{j=1}^{N} c_j
\langle \mathbf{p}|\mathbf{x}_j\rangle.
\end{equation*}
Here, the set of points $\{\mathbf{x}_j|j=1,\ldots,N\}$ is a subset (graph) of
$\mathbb{R}^3$. The inner product is
\begin{align*}
\langle \mathbf{x}_j|\mathbf{x}_l\rangle &= \lim_{T\to\infty} \frac{1}{(2T)^3}\int_{-T}^T\int_{-T}^T\int_{-T}^T
\mathrm{e}^{-i\mathbf{p}(\mathbf{x}_j-\mathbf{x}_l)}\ \mathrm{d}^3p\\
&= \delta_{\mathbf{x}_j,\mathbf{x}_l},
\end{align*}
in which plane waves are normalizable (the right hand side is the
generalization of the Kronecker delta to continuous indices). The configuration and translation
operators $\hat{\mathbf{x}}=-i\nabla_{\mathbf{p}}$ and $\hat{u}=\widehat{\mathrm{e}^{i\boldsymbol{\lambda}\mathbf{p}}}$ act
as
\begin{equation*}
\hat{\mathbf{x}}\mathrm{e}^{i \mathbf{p}\mathbf{x}_j} = \mathbf{x}_j \mathrm{e}^{i \mathbf{p}\mathbf{x}_j}, \ \ \ \ \ \
\hat{u}\mathrm{e}^{i\mathbf{p}\mathbf{x}_j}= \mathrm{e}^{i\mathbf{p}(\mathbf{x}_j -\boldsymbol{\lambda})}.
\end{equation*}
However, the usual momentum operator $\hat{\mathbf{p}}=-i\nabla_{\mathbf{x}}$ is not defined.
Only \emph{finite} translations can be realized and the \emph{infinitesimal} translations do not exist
in this Hilbert space.
The polymer quantum mechanics introduced just now has been applied to the harmonic
oscillator \cite{AFW}, to the potentials $1/|\mathbf{x}|$ \cite{HLW-hyd} and $1/|\mathbf{x}|^2$ \cite{KLZ}, as well as to the cosmology and black hole cases mentioned earlier.
What is missing, however, is a complete application to QFT that goes beyond
the effective approach in Ref. \cite{HHS-sft} and that would allow for a computation of the
vacuum expectation values of field operators. This is the program we initiate
in this paper.
\subsection{Field theory}
The generalization of the quantization from quantum mechanics to field theory is
straightforward. The Poisson bracket of the basic scalar field
variables in Eq. \eqref{basicpb} is realized as an operator relation on a Hilbert space
with basis states
\begin{equation}
| \mu_1,\ldots ,\mu_N\rangle,
\label{basis}
\end{equation}
where the set of real numbers $\{\mu_j|j=1,\ldots,N\}$ represent scalar field values at the
set of space points $\{\mathbf{x}_j|j=1,\ldots,N\}$. Figure \ref{H-picture} gives an illustration of a basis state with $N=7$. If two basis states are associated with the same set of points $\{\mathbf{x}_j|j=1,\ldots,N\}$, the inner product is given by
\begin{equation}
\langle \mu_1',\ldots ,\mu_N'| \mu_1,\ldots ,\mu_N\rangle=
\delta_{\mu_1',\mu_1}\cdot\ldots\cdot \delta_{\mu_N',\mu_N}
\label{ip}
\end{equation}
and it is zero otherwise.
Equation \eqref{ip} defines an inner product that is background (metric) independent in the same
way as, for example, that for Ising model spins. A difference is, however, that
for the latter there is a finite dimensional spin space at each
lattice point rather than a real number representing a field value
at each point. In contrast, the usual Fock space quantization uses
the metric dependent Klein-Gordon inner product. We note that, when
compared with LQG, the set of points on which a
state is based may be called a ``graph" and that there is no scale
associated with the set unless one specifically chooses a uniform lattice.
\begin{figure}
\vspace{0.8cm}
\begin{center}
\includegraphics {h-space.pdf}
\caption{ A pictorial depiction of the basis state $|\mu_1,\ldots,\mu_7\rangle$. }
\label{H-picture}
\end{center}
\vspace{-.5cm}
\end{figure}
If we remember that we take $f$ to be a Gaussian sharply peaked
at a point $\mathbf{x}_j$, see Eq. \eqref{gauss}, and if we set
\begin{equation*}
\hat{\Phi}_f(t,\mathbf{x}_j) \equiv \hat{\Phi}_j, \ \ \ \ \ \ \hat{U}(t,\mathbf{x}_j) \equiv \hat{U}_j,
\end{equation*}
we have the following actions
\begin{align*}
\hat{\Phi}_j| \mu_1,\cdots, \mu_n\rangle&= \mu_j|
\mu_1,\ldots ,\mu_n\rangle,\\
\widehat{U}_j| \mu_1,\ldots \mu_n\rangle &= | \mu_1,\ldots , \mu_j- \lambda/V, \ldots ,\mu_n\rangle.
\end{align*}
Noting that $f(\mathbf{x}_j-\mathbf{x}')\approx1$ for $(\mathbf{x}_j-\mathbf{x}')^2\ll\sigma^2$ and $f(\mathbf{x}_j-\mathbf{x}')\approx0$ otherwise, it is readily verified that the commutator of these operators leads to a
faithful realization of the corresponding Poisson bracket in Eq. \eqref{basicpb}. Indeed, using $\delta_{j,l}$ for $\delta_{\mathbf{x}_j,\mathbf{x}_l}$, we have
\begin{equation}\label{basic-comm}
[ \hat{\Phi}_j, \hat{U}_l] = - \frac{\lambda}{V}\delta_{j,l}\hat{U}_l, \ \ \ \ \ \ [ \hat{\Phi}_j, \hat{U}_l ^{\dagger}] = \frac{\lambda}{V} \delta_{j,l}\hat{U}_l^{\dagger}.
\end{equation}
We note that fixing the parameter $\lambda$ is equivalent to selecting a discreteness scale in field
configuration space, as may be seen from the action of $\hat{U}$
on basis states given above.
In this representation the momentum operator does not exist because of the ``point"
nature of the inner product given in Eq. (\ref{ip}) (basis states with different excitations at a
point are orthogonal, as are any states with arbitrary excitations at different points).
There is, however, an alternative $\lambda$ dependent definition of
a ``momentum" operator given by
\begin{equation}
\hat{P}_j = \frac{1}{2i\lambda}\left(\hat{U}_j - \hat{U}^{\dagger}_j \right),
\label{mom}
\end{equation}
which can be used to define the kinetic part of the Hamiltonian operator
and any other operators that involve the momentum. For example, the
canonical commutator gets modified from that in usual field theory to
\begin{equation}
[ \hat{\Phi}_j, \hat{P}_l] = \frac{i}{2V} \delta_{j,l}\left( \hat{U}_j + \hat{U}^{\dagger}_j \right).
\label{field-comm}
\end{equation}
This directly follows from Eqs. \eqref{basic-comm} and \eqref{mom}.
\subsection{Fock operators}
We would like to define analogs of the usual field theory creation and
annihilation operators on the polymer Hilbert space introduced above.
To do this, we follow as closely as possible the standard constructions, pointing
out along the way where the important differences arise. We begin with the Fourier
decomposition of the classical field variable and its conjugate momentum
\begin{equation}\label{field-mom}
\begin{split}
\phi(t,\mathbf{x}) &=\su{\mathbf{k}}\left[ a_{\mathbf{k}}f_\mathbf{k}(t,\mathbf{x}) + \overline{a}_{\mathbf{k}}\fbk{}(t,\mathbf{x}) \right],\\
p(t,\mathbf{x})&=-i\su{\mathbf{k}}\ok{}\left[ a_{\mathbf{k}}f_\mathbf{k}(t,\mathbf{x}) - \overline{a}_{\mathbf{k}}\fbk{}(t,\mathbf{x}) \right],
\end{split}
\end{equation}
where
\begin{equation*}
\fk{}(t,\mathbf{x})= \frac{\mathrm{e}^{-i\ok{}t+i\mathbf{k}\mathbf{x}}}{ \sqrt{2\omega_\mathbf{k} V} }
\end{equation*}
are the usual flat space plane waves, $\mathbf{k}=2\pi\mathbb{Z}^3/L$ with $L^3=V$ and $\ok{}=\sqrt{\mathbf{k}^2+m^2}$. Given the (Klein-Gordon) inner product
\begin{equation*}
\langle \chi,\psi\rangle=i\int_{V}\left(\dot{\chi}\overline{\psi}-\chi\dot{\overline{\psi}}\right)\mathrm{d}^3x
\end{equation*}
for two scalar fields $\chi$ and $\psi$, the plane waves satisfy
\begin{equation*}
\langle\fk{},\fk{'}\rangle=\langle\fbk{},\fbk{'}\rangle=\delta_{\mathbf{k},\mathbf{k}'},\ \ \ \ \ \ \langle\fk{},\fbk{'}\rangle=0.
\end{equation*}
Here, $\delta_{\mathbf{k},\mathbf{k}'}$ denotes the standard Kronecker delta symbol. The expansions in Eq. \eqref{field-mom} give rise to the standard expressions for the creation and annihilation operators
\begin{align*}
\hat{a}_{\mathbf{k}}&=i\int_V\fbk{}\left(\hat{p}-i\ok{}\hat{\phi}\right)\mathrm{d}^3x,\\
\hat{a}_{\mathbf{k}}^{\dag}&=-i\int_V\fk{}\left(\hat{p}+i\ok{}\hat{\phi}\right)\mathrm{d}^3x.
\end{align*}
We now use this standard procedure to {\it motivate} a definition of the corresponding operators in polymerized QFT using the momentum given in Eq. (\ref{mom}); it is important to note at the outset that
this may not be the only possible way to proceed and we follow it here simply because it will permit a simple test of the usual field theory limit. We will also see that, although the $\hat{a}_{\mathbf{k}}$ are time independent in usual field theory, the corresponding entities in the polymer theory are not, except in the usual field theory limit.
To begin, let us consider a subset of points $\{\mathbf{x}_j|j=1,\ldots,N\}$, which will represent the ``graph" on which the operators will be defined. (In general, this set can be chosen to be countably infinite, such as a uniformly spaced lattice but this is not necessary to give the basic prescription.) We define now
\begin{align*}
\Ahk{}&=\frac{iV}{\sqrt{N}} \sum_{j=1}^N \fbk{j}\left( \hat{P}_j -i\ok{} \hat{\Phi}_j \right),\\
\Ahdk{}&=-\frac{iV}{\sqrt{N}} \sum_{j=1}^N \fk{j}\left( \hat{P}_j +i\ok{} \hat{\Phi}_j \right).
\end{align*}
We note that these operators depend on $\lambda$ (because of $\hat{P}$) and on $t$, and that they are
non-local that is to say they are a sum of operators, one at each point of the graph. The factor $1/\sqrt{N}$ is necessary to produce the correct low frequency limit of vacuum expectation values, as we will see below. Also, the explicit time dependence is emphasiced by the non-trivial Heisenberg equations of motion given later.
The polymer Hamiltonian operator may be defined in the same way as in standard field theory by
\begin{equation}
\hat{H}= \su{\mathbf{k}}\ok{}\left(\Ahk{}^{\dagger}\Ahk{}+\frac{1}{2}[\Ahk{},\Ahk{}^{\dagger}]\right).
\label{hamilt}
\end{equation}
We will work with this exact form of the Hamiltonian rather than the normal ordered one
because, as we will see below, it allows us to compute explicitly a modification to the energy-momentum dispersion relation. In addition, it is useful to keep the full form in any case because
we are after all not working with the usual Fock space quantization where the heuristic argument for
normal ordering is perhaps more compelling.
Using the relations (\ref{basic-comm}), it can be seen that the commutator and anti-commutator
are given by
\begin{align*}
[ \Ahk{}, \Ahdk{}] &= \frac{1}{2N} \sum_{j=1}^N \left( \hat{U}_j +\hat{U}_j^{\dagger} \right),\\
\{\hat{A}_{\mathbf{k}}, \hat{A}_{\mathbf{k}}^{\dag}\} &= \frac{V^2}{N}\sum_{j,l=1}^N\overline{f}_{\mathbf{k}j}f_{\mathbf{k}l} \Big[2 \hat{P}_j\hat{P}_l +2\omega_{\mathbf{k}}^2\hat{\Phi}_j\hat{\Phi}_l \nonumber\\
&\quad +i\omega_{\mathbf{k}} \Big( \{\hat{P}_j,\hat{\Phi}_l \} - \{\hat{\Phi}_j,\hat{P}_l \} \Big)
\Big],
\end{align*}
where in the first expression we used the definition (\ref{mom}) of the momentum operator. We note here a substantial departure from the usual field theory commutators: the right hand sides are not constant but effectively contain an infinite series in powers of the momentum, if the operators $\hat{U}$ and $\hat{U}^{\dag}$ are viewed as arising from the classical Poisson brackets, and an infinite series in powers of the operators $\hat{\Phi}_j$. However, as we will see below, the correct field theory limits can be recovered at low energies.
\subsection{Vacuum state and expectation values}
Let us now turn to a discussion of an analog of Fock states in the polymer Hilbert space. Such states will allow us to further probe the commutation rules derived above. The basic idea we follow is that the ``vacuum'' should be a product of oscillator vacuua, one for each point $\mathbf{x}_j$ of the sample set of points $\{\mathbf{x}_j|j=1,\ldots,N\}$. Let us first focus on the case of one point, or equivalently the quantum mechanics case, and consider the state
\begin{equation}
| G_\mathbf{k}\rangle_j = \frac{1}{\pi^{1/4}}\su{\mu} \mathrm{e}^{- \ok{}V\mu^2/2}|\mu\rangle_j.
\label{Gj}
\end{equation}
This is just the normalized ground state of an oscillator of frequency $\ok{}$ at the point $\mathbf{x}_j$.
It is normalized to $1$ if we evaluate the sum in Eq. \eqref{Gj} by means of
\begin{equation*}
\su{\mu}f(\mu)=\sqrt{\ok{}V}\int_{\mathbb{R}}f(\mu)\ \mathrm{d}\mu
\end{equation*}
for an arbitrary integrable function $f$.
Using the state $| G_\mathbf{k}\rangle_j $, we define the ``polymer Fock vacuum" on the $N$ point graph
in the mode $\mathbf{k}$ by
\begin{equation*}
|0_\mathbf{k} \rangle = \bigotimes_{j=1}^N | G_\mathbf{k}\rangle_j.
\end{equation*}
The full vacuum is the tensor product
\begin{equation}
|0\rangle = \sideset{}{_\mathbf{k}}\bigotimes |0_\mathbf{k} \rangle
\end{equation}
We will now see that the states $|0_\mathbf{k}\rangle$ and $|0\rangle$ have the properties expected of the standard quantum field theory vacuum at sufficiently low frequencies by evaluating several vacuum expectation values. We first define the dimensionless parameter
\begin{equation*}
\gamma_{\mathbf{k}} \equiv \ok{}\lambda^2/V
\end{equation*}
to simplify the following expressions.
The expectation value of the number operator is
\begin{equation*}
\langle0_\mathbf{k}|A_\mathbf{k}^{\dagger}A_\mathbf{k} |0_\mathbf{k} \rangle =
\frac{1}{4}-\frac{ \mathrm{e}^{-\gamma_{\mathbf{k}} /4}}{2} +\frac{1-\mathrm{e}^{-\gamma_{\mathbf{k}}}}{4\gamma_{\mathbf{k}} }.
\end{equation*}
For the expectation values of the commutator and the anti-commutator we need
\begin{equation*}
\langle0_\mathbf{k}| U_j |0_\mathbf{k} \rangle = \langle0_\mathbf{k}| U_j^{\dag} |0_\mathbf{k} \rangle=\mathrm{e}^{-\gamma_{\mathbf{k}}/4 }.
\end{equation*}
This gives
\begin{equation}\label{comm-acomm-expec}
\begin{split}
\langle0_\mathbf{k}|[ A_\mathbf{k}, A_\mathbf{k}^{\dagger}] |0_\mathbf{k} \rangle &= \mathrm{e}^{-\gamma_{\mathbf{k}}/ 4},\\
\langle0_\mathbf{k}| \{ A_\mathbf{k}, A_\mathbf{k}^{\dagger} \} |0_\mathbf{k} \rangle &=
\frac{1}{2} +\frac{1- e^{-\gamma_{\mathbf{k}} }}{2\gamma_{\mathbf{k}} }.
\end{split}
\end{equation}
The expectation value of the Hamiltonian is then
\begin{equation}\label{H-expec}
\langle0 | \hat{H} |0 \rangle = \frac{1}{4}\su{\mathbf{k}} \ok{} \left( 1+\frac{1-\mathrm{e}^{-\gamma_{\mathbf{k}} }}{\gamma_{\mathbf{k}} }\right).
\end{equation}
Lastly, it is useful to see the expectation value of the canonical field-momentum commutator
(\ref{field-comm}), which is
\begin{equation*}
\langle0_\mathbf{k}| [ \hat{\Phi}_j, \hat{P}_l] |0_\mathbf{k}\rangle = \frac{i}{V} \delta_{j,l} \ \mathrm{e}^{-\gamma_{\mathbf{k}} /4}.
\end{equation*}
This gives a scale dependent modification to the uncertainty relation.
\begin{figure}
\vspace{0.8cm}
\begin{center}
\includegraphics {f0.pdf}
\caption{The expectation value of the commutator, the anti-commutator and the number
operator in the state $|0_\mathbf{k}\rangle$ as a function of $\ln(\gamma_{\mathbf{k}})$.}
\label{fig0}
\end{center}
\vspace{-.5cm}
\end{figure}
\begin{figure}
\vspace{0.8cm}
\begin{center}
\includegraphics {f1.pdf}
\caption{The expectation value of the commutator, the anti-commutator and the number operator in the state $|1_\mathbf{k}\rangle$ as a function of $\ln(\gamma_{\mathbf{k}})$.}
\label{fig1}
\end{center}
\vspace{-.5cm}
\end{figure}
It is instructive to consider the low energy limits of these expressions. For $\gamma_{\mathbf{k}} \ll 1 $ all the usual field theory vacuum expectation values are recovered. This justifies our definitions of the vacuum and the creation and annihilation operators. The vacuum becomes an eigenstate of the Hamiltonian in this limit.
We note that $\lambda$ dependent corrections set in when $\gamma_{\mathbf{k}}\sim 1$.
Figure 2 shows the vacuum expectation value of the commutator, the anti-commutator and the number operator as a function of $\gamma_{\mathbf{k}}$ over several orders of
magnitude. From this we see a surprising behavior at sufficiently large frequencies: {\it the
commutator vanishes and the anti-commutator tends to one half}. This apparent fermionic behavior
is also clear from Eq. (\ref{comm-acomm-expec}).
To see whether other states exhibit this behavior for large $\gamma_{\mathbf{k}}$, it is useful to consider
multiparticle states, which may be defined following the standard prescription
\begin{equation*}
|n_\mathbf{k} \rangle = \frac{1}{\sqrt{\langle n_{\mathbf{k}}|n_{\mathbf{k}}\rangle}} \left( \Ahk{}^{\dagger} \right)^n |0_\mathbf{k}\rangle.
\end{equation*}
We computed the same expectation values for up to three particle states using a symbolic algebra
procedure. The results appear in Figs. 2-5. The main features evident from each of these graphs
are that (i) they indicate the correct low energy limit, and (ii) at high energies the expectation value of the commutator
becomes negligible compared to that of the anti-commutator. We note also that the expectation value of the number operator
rises with $\gamma_{\mathbf{k}}$ especially for the two and three particle states.
%
\begin{figure}
\vspace{0.8cm}
\begin{center}
\includegraphics {f2.pdf}
\caption{The expectation value of the commutator, the anti-commutator and the number operator in the state $|2_\mathbf{k}\rangle$ as a function of $\ln(\gamma_{\mathbf{k}})$.}
\label{fig2}
\end{center}
\vspace{-.5cm}
\end{figure}
\begin{figure}
\vspace{0.8cm}
\begin{center}
\includegraphics {f3.pdf}
\caption{The expectation value of the commutator, the anti-commutator and the
number operator in the state $|3_\mathbf{k}\rangle$ as a function of $\ln(\gamma_{\mathbf{k}})$.}
\label{fig3}
\end{center}
\vspace{-.5cm}
\end{figure}
\subsection{Time evolution}
The calculations presented so far have been performed at a single time slice that is to say at a fixed time. Since we have the Heisenberg picture in mind, it is important to
ask what features of our results are invariant under a time evolution. This is can be checked by
calculating commutators with the Hamiltonian (\ref{hamilt}) such as
%
\begin{align*}
[ \hat{H}, [\Ahk{}, \Ahdk{} ] ] &= -i\frac{\ok{}^2\lambda^2}{N}
\sum_{j=1}^N\left( \fk{j}\hat{P}_j\Ahk{} + \fbk{j}\Ahdk{}\hat{P}_j\right),\\
[\hat{H},\Ahdk{}\Ahk{}]&=i\frac{\ok{}^2\lambda^2}{2N}
\sum_{j=1}^N \left(\fk{j}\hat{P}_j\Ahk{}+\fbk{j}\Ahdk{}\hat{P}_j\right),\\
[ \hat{H},\{\Ahk{},\Ahdk{} \} ] &= 0.
\end{align*}
We thus see that at the operator level, only the anti-commutator is a constant of the motion.
It is also interesting to check the expectation values of the Heisenberg equation of motion.
For example, for the vacuum state $|0_\mathbf{k}\rangle$ we find
\begin{align*}
\langle0_\mathbf{k}|[\hat{H},\Ahdk{}\Ahk{}]|0_\mathbf{k}\rangle&=0,\\
\langle0_\mathbf{k}|[\hat{H},[\Ahk{},\Ahdk{}]]|0_\mathbf{k}\rangle&=0.
\end{align*}
It therefore appears that the fermionic behavior at high energies is time independent, at least for the vacuum state.
\subsection{Modified dispersion relation}
There has been a recent surge in interest in exploring classical Lorentz violating theories which give rise to modified dispersion relations. This is partly due to the view that a quantum theory of gravity must have this feature at high energies, since such a theory must have a built in scale. Observational bounds have been put on Lorentz violating terms in the dispersion relations\cite{ted-rev}, and there have been some derivations of such effects from LQG \cite{GP-disp,Th-Sahl-disp}. It has even been suggested that there is a doubly special relativity \cite{AC-dsr, Lee-dsr} in which modified Lorentz transformations are characterized by not just the speed of light but also by an additional scale.
The quantization we have developed is necessarily Lorentz violating at high energies due
to the scale $\lambda$, so it is interesting to see if we can derive a correction to the dispersion relation.
Using the same quantization method as employed here, this has been done in an effective approach using semiclassical states \cite{HHS-sft}. However, we now show, there is an alternative derivation that does not use semiclassical states and that gives a correction
of a lower order.
From the expectation value of the Hamiltonian in Eq. (\ref{H-expec}) we can identify
the $\lambda$ modified
mode energy
%
\begin{equation}
E_\mathbf{k} = \frac{1}{2}\ok{} \left( 1 + \frac{1-
e^{-\gamma_{\mathbf{k}} } }{\gamma_{\mathbf{k}} }\right).
\end{equation}
%
Expanding the exponential and remembering that $\omega_\mathbf{k}^2 = \mathbf{k}^2 + m^2=|\mathbf{k}|^2 + m^2$, we find that for $|\mathbf{k}| \gg m$
%
\begin{equation}
E_\mathbf{k}^2 = | \mathbf{k}|^2 -
\frac{\lambda^2}{2V} |\mathbf{k}|^3 + O\left(|\mathbf{k}|^4\right).
\end{equation}
%
It is interesting that this gives a leading order correction proportional to $|\mathbf{k}|^3$.
If we take $V$ to be a particle physics mass scale $M$ and $\lambda$ to be the
Planck scale, the dispersion relation may be written
\begin{equation}
E^2_{\mathbf{k}} = |\mathbf{k}|^2 \left[1 - \frac{1}{2} \left( \frac{M}{M_P} \right)^3 \left( \frac{k}{M_P} \right) + \cdots\right],
\end{equation}
which shows a strong suppression.
\section{Discussion}
We have developed a background independent ``Fock'' quantization for the scalar field theory by the relatively straightforward prescription of following standard methods. Our starting point was a set of
classical variables which are essentially the integrated field variable and exponentiated momentum. Utilizing these as the basic variables, we defined analogs of Fock operators and the vacuum directly in the polymer Hilbert space, and computed the expectation values of commutators and anti-commutators. The results give the desired low energy limit of usual quantum field theory, and predict a significant departure at high energy. We also showed that at least for the vacuum state, the
observed fermionic behavior at high energies is time evolution invariant.
The expectation value of the number operator in the state $|0_\mathbf{k}\rangle$ tends asymptotically to $1/4$ as $\gamma_{\mathbf{k}}\rightarrow \infty$ and that of the Hamiltonian goes to $0$. Thus, unfortunately, there is no change in the usual vacuum energy problem; we note, however, that we have not diagonalized the Hamiltonian, which is what may be necessary to probe this issue. Indeed, we have merely postulated and justified a choice of vacuum due to its correct low energy limit. This state may not be the true ground state of the Hamiltonian for $\gamma_{\mathbf{k}} \gg 1$, since we expect that such a state should depend on the parameters in the Hamiltonian, which in this case is $\lambda$. An exact determination of the energy spectrum, however, requires a numerical approach due to the $\lambda$ dependence of the Hamiltonian. This is presently under investigation \cite{HHS-spec}.
What is the physical interpretation of the apparent fermionic behavior at large $\gamma_{\mathbf{k}}$? Let us recall that the parameter $\lambda$ is not a discretization scale associated with space (or time). It is the scale associated with field translations once we have fixed the field momentum operator (\ref{mom}). Thus, the fermionic behavior is not a result of discretizing space (no such behaviour is observed for standard field theory on a lattice) and it would be puzzling if this were the cause here. Furthermore, there is no indication that the methods we have developed are breaking down at large $\gamma_{\mathbf{k}}$.
It is interesting to note that an effective repulsion is evident when a similar quantization method is applied to the gravitational collapse of a scalar field with quantum gravity corrections \cite{VH-qgc, KZ-qgc}. In these works a feature of polymer quantization
is applied to gravitational variables, which effectively lead to a mass gap at the onset of black hole formation. Although this is not directly an effect of the vanishing of certain commutators at small length scales, it is tempting to speculate that its ultimate origin may be related.
In addition to seeking a better understanding of the high energy behavior, our results open up several other potentially fruitful directions for further investigation. Among these are (i) the computation of the propagator, which would require a solution of the Heisenberg equations of motion for the Fock operators or an explicit construction of the exponentiated Hamiltonian, (ii) the free and interacting field theory on a general background, (iii) the quantization of a fermionic field theory, (iv) the application to ``weak field'' loop quantum gravity in order to study gravitons, and, finally, (v) applications to Hawking radiation and cosmology. The latter would be particularly interesting because the evolution of the universe could itself provide a transition from fermionic to bosonic behavior during expansion if this method of quantization is the correct one.
\vskip 2 cm
\begin{acknowledgments}
This work was supported in part by the Natural Science and Engineering Research Council of Canada. We thank Jack Gegenberg, Golam Hossain, Gabor Kunstatter and Sanjeev Seahra for discussions.
\end{acknowledgments}
|
\section{introduction}
The algebraic question for functorial coalgebra decompositions of the tensor algebras was arising from homotopy theory described as follows. The relevant topological question is how to decompose the loop spaces. This is a classical problem in homotopy theory with applications to homotopy groups. For instance, the classical results of Cohen-Moore-Neisendorfer~\cite{CMN} on the exponents of the homotopy groups of the spheres and Moore spaces were obtained from the study of the decompositions of the loop spaces of Moore spaces. The decompositions of the loop space functor $\Omega$ from $p$-local simply connected co-$H$-spaces to spaces was then introduced in~\cite{SW1} with the subsequent development in~\cite{STW,STW2,SW2, Theriault}. Namely one gets the natural decompositions $\Omega X\simeq \bar A(X)\times \bar B(X)$ for some homotopy functors $\bar A$ and $\bar B$ on $p$-local simply-connected co-$H$-spaces~$X$. Such decompositions may lose some information for an individual space $X$ in the sense that the functor $\bar A$ may be indecomposable but the space $\bar A(X)$ may have further decompositions. However the functorial decompositions has a good property that one can
freely change the co-$H$-spaces $X$ in the decomposition formulas because they are functorial.
Also there are examples of spaces $X$ such as Hopf invariant on complexes given in~\cite{GSW} with the property that the space $\bar A(X)$ is indecomposable for certain functor $\bar A$. A fundamental question concerning the functorial decompositions is how to determine the homology of the factors $\bar A(X)$ and $\bar B(X)$, which can be reduced to be a pure algebraic question as follows:
Let $V$ be any module over a field $\ensuremath{{\mathbf{k}}}=\mathbb{Z}/p$ and let $T(V)$ be the tensor algebra on ~$V$. Then $T(V)$ becomes a Hopf algebra by saying $V$ to be primitive. By forgetting the algebra structure on $T(V)$, we have the functor $T$ from modules to coalgebras. Let
\begin{equation}\label{equation1.1}
T(V)\cong A(V)\otimes B(V)
\end{equation}
be any natural coalgebra decomposition of $T(V)$ for some functors $A$ and $B$ from (ungraded) modules to coalgebras. From~\cite[Theorem 1.3]{SW1}, the functors $A$ and $B$ can be canonically extended as functors from graded modules to graded coalgebras and the above decomposition formula holds for any graded module $V$. Then from~\cite[Theorem 1.1]{STW2}, the functors $A$ and $B$ induce functor $\bar A$ and $\bar B$ from co-$H$-spaces to spaces with the natural decomposition $\Omega X\simeq \bar A (X)\times \bar B(X)$ with the property that there exist filtrations on mod $p$ homology $H_*(\bar A(X))$ and $H_*(\bar B(X))$ such that its associated graded modules $E^0H_*(\bar A(X))\cong A(\Sigma^{-1}\bar H_*(X))$ and $E^0H_*(\bar B(X))\cong B(\Sigma^{-1}\bar H_*(X))$, where $\Sigma^{-1}$ is the desuspension of the graded modules. Briefly speaking, any coalgebra decomposition of the functor $T$ as in formula~(\ref{equation1.1}) induces a natural decomposition of the loops on $p$-local simply connected co-$H$-spaces in which the homology of its factors can be determined by their corresponding algebraic functors.
The functors $A$ and $B$ in decomposition~(\ref{equation1.1}) are complementary to each other and so it suffices to understand one of them as a coalgebra summand of the functor $T$. There exist some important coalgebra summands of $T$ in~\cite[Theorem 6.5]{SW1} that give the functorial version of the Poincar\'e-Birkhoff-Witt Theorem. One of such functors is the functor $A^{{\mathrm{min}}}$, which is the smallest natural coalgebra summand of $T(V)$ containing $V$. The coalgebra complement of the functor $A^{{\mathrm{min}}}$, denoted by $B^{\mathrm{max}}$, has the property that $B^{\mathrm{max}}(V)$ can be chosen as a sub Hopf algebra of $T(V)$. However the determination of $A^{{\mathrm{min}}}(V)$ and $B^{\mathrm{max}}(V)$ seems out of current technology. As a consequence, the homology of the geometric realization $\bar A^{{\mathrm{min}}}(X)$ and $\bar B^{\mathrm{max}}(X)$ remains unknown. It is then important to find coalgebra summands $B$ of $T$ with the explicit information on $B(V)$ because in such a case the homology of the geometric realization $\bar B(X)$ can be understood.
The purpose of this article is to provide some explicit coalgebra summands $B$ of $T$. We are interested in the special cases where $B$ can be chosen as a sub Hopf algebra of $T$. This will give a relatively large coalgebra summand of $T$ because any subalgebra of tensor algebra is a tensor algebra, and so the complement functor $A$ of $B$ as in decomposition~(\ref{equation1.1}) becomes relatively small. Let $L_n(V)$ be the $n\,$th free Lie power on $V$, namely $L_n(V)$ is the homogenous component of the free Lie algebra $L(V)$ on $V$ of tensor length $n$. Our main result is as follows:
\begin{thm}\label{theorem1.1}
Let the ground ring be a field of characteristic $p$. Let $\{m_i\}_{i\in I}$ be finite or infinite set of positive integers prime to $p$ with each $m_i>1$. Then the sub Hopf algebra of $T(V)$
generated by
$$
L_{m_ip^r}(V)\quad \textrm{ for } i\in I,\ r\geq 0
$$
is a natural coalgebra summand of $T(V)$. In particular, the sub Hopf algebra $B(V)$ of
$T(V)$ generated by
$$
L_n(V)\quad \textrm{ for } n \textrm{ NOT a power of } p
$$
is a natural coalgebra summand of $T(V)$.
\end{thm}
By the maximum property of the functor $B^{\mathrm{max}}$, the sub Hopf algebras in the theorem are all contained in $B^{\mathrm{max}}$. According to~\cite[Proposition 11.1]{SW1} as well as~\cite{Bryant-Stohr}, the indecomposable elements in $B^{\mathrm{max}}$ does not has tensor length $p$ for $p>2$ and so the sub Hopf algebra $B(V)$ coincides with $B^{\mathrm{max}}(V)$ up to tensor length $p^2-1$. For the case $p=2$, the sub Hopf algebra $B(V)$ coincides with $B^{\mathrm{max}}(V)$ up to tensor length $7$ according to the computations in~\cite{SW4}. Our sub Hopf algebra $B(V)$ is strictly smaller than $B^{\mathrm{max}}(V)$ for general module $V$. However if the functors $B$ and $B^{\mathrm{max}}$ are extended to the functors from graded modules to graded modules in the sense of~\cite{SW1}, the sub Hopf algebra $B(V)$ coincides with $B^{\mathrm{max}}(V)$ for graded modules $V$ of dimension $\leq p-1$ with $V_{\mathrm{even}}=0$ according to ~\cite[Theorem 1.1]{Wu5}.
There is a canonical connection between coalgebra decompositions of $T$ and the decompositions of the Lie powers $L_n(V)$ as modules over the general linear groups by restricting decomposition~(\ref{equation1.1}) to the primitives. The decompositions of Lie powers have been actively studied in the recent development of representation theory~\cite{Bryant-Johnson, Bryant-Schocker, Bryant-Stohr, Donkin-Erdmann, Erdmann-Schocker}.
The article is organized as follows. In section~\ref{section2}, we investigate the sub-quotient functors of the tensor power functors $T_n\colon V\mapsto T_n(V)=V^{\otimes n}$ from modules to modules. These special functors are of course closely related to the tensor representation of the symmetric groups and the finite dimensional polynomial representations of the general linear groups (by evaluating on a fixed module $V$). They are also related the modules over the Schur algebras and the modules over the Steenrod algebra~\cite{Kuhn1, Kuhn2,Kuhn3}. In this section, we introduce exact functors $\gamma_n(-)$ from the category of functors from modules to modules to the category of the modules over the symmetric groups in this section which are variations of the classical Schur functor~\cite{ABW, Schur, Green}. In geometry, the summands of the tensor power functors $T_n$ are closely related to decompositions of self-smash products~\cite{SW3, Wu}.
In section~\ref{section3}, we investigate the subfunctors of the Lie powers $L_n$ that occur as the summands of the functor $T_n$, which we call $T_n$-projective subfunctors of $L_n$. According to Theorem~\ref{theorem3.9}, these functors are closely related to the summands of the Lie powers $L_n(V)$ that occur as summands of $V^{\otimes n}$ studied in~\cite{Bryant-Schocker, Donkin-Erdmann, Erdmann-Schocker}.
We give a coalgebra decomposition of $T$ called the Block Decomposition in section~\ref{section4}. According to Theorem~\ref{theorem4.3}, this is a coalgebra decomposition of $T$ in the form
$$
T\cong \bigotimes_{i=1}^\infty C^{m_i},
$$
where $\{m_i\}$ is the set of all positive integers prime to $p$ and the primitives of $C^{m_i}$ are exactly given by the primitives of the tensor algebra $T$ with tensor length $m_ip^r$ for $r\geq 0$. In other words, the primitives of the tensor algebra $T$ with tensor length $mp^r$ for $r\geq 0$ for each $m$ prime to $p$ can be blocked into a coalgebra summand $C^m$ of $T$.
The proof of Theorem~\ref{theorem1.1} is given in section~\ref{section5}. In section~\ref{section6}, we give some applications of our decomposition theorem to Lie powers by restricting to the primitives.
\section{The structure on the Tensor Powers}\label{section2}
\subsection{Tensor Algebras}\label{subsection2.1}
Let $V$ be any module and let
$$T(V)=\bigoplus_{n=0}^{\infty} V^{\otimes n}$$ be the \textit{tensor algebra} generated by $V$, where $V^{\otimes n}=\ensuremath{{\mathbf{k}}}$ and the multiplication on $T(V)$ is given by the formal tensor product of monomials. The tensor algebra admits the universal property that, for any associated algebra $A$ and any linear map $f\colon V\to A$, there exists a unique algebra map $\tilde f\colon T(V)\to A$ such that $\tilde f|_V=f$. In particular, the linear map
$$
V\longrightarrow T(V)\otimes T(V) \quad x\mapsto x\otimes 1+1\otimes x
$$
extends uniquely to an algebra map $\psi\colon T(V)\to T(V)\otimes T(V)$ and so $T(V)$ has the canonical Hopf algebra structure with the multiplication given by the formal tensor product of monomials and the comultiplication given by $\psi$. We refer to~\cite{MM} as a classical reference for Hopf algebras and quasi-Hopf algebras. The comultiplication $\psi$ is coassociative and cocommutative. The module $T(V)$ with the comultiplication $\psi\colon T(V)\to T(V)\otimes T(V)$ is called \textit{shuffle coalgebra} as its graded dual
$$
T^*(V)=\bigoplus_{n=0}^\infty \left(V^{\otimes n}\right)^*\cong \bigoplus_{n=0}^\infty (V^*)^{\otimes n}
$$
is the usual shuffle algebra under the multiplication $\psi^*\colon T^*(V)\otimes T^*(V)\to T^*(V)$.
We are interested in the functor $T\colon V\mapsto T(V)$. There are three terminologies on this functor:
\begin{enumerate}
\item The functor $T^H\colon V\mapsto T(V)$ from modules to Hopf algebras;
\item The functor $T^C\colon V\mapsto T(V)$ from modules to coalgebras by forgetting the multiplication;
\item The functor $T^M\colon V\mapsto T(V)$ from modules to modules by forgetting both multiplication and comultiplication.
\end{enumerate}
The notation of the functor $T$ refers one of $T^H$, $T^C$ or $T^M$ if the working category is clear.
By taking the tensor length from the definition of $T(V)$, the functor $T^M$ admits a natural decomposition that
\begin{equation}\label{equation2.1}
T^M\cong \bigoplus\limits_{n=0}^\infty T_n,
\end{equation}
where $T_n(V)=V^{\otimes n}$ with $T_0(V)=\ensuremath{{\mathbf{k}}}$. Thus the functors $T^H$, $T^C$ and $T^M$ are graded functors. From the well-known property (see for instance~\cite[Lemma 3.8]{GW}) that
\begin{equation}\label{equation2.2}
\mathrm{H o m}(T_n,T_m)=\left\{
\begin{array}{lcl}
0&\textrm{ if }& n\not=m\\
\ensuremath{{\mathbf{k}}}(\Sigma_n)&\textrm{ if }&n=m,\\
\end{array}\right.
\end{equation}
the decomposition of $T^M$ is in fact an orthogonal decomposition. A direct consequence is that the comultiplication $\psi$ (as a natural transformation) is uniquely determined by the multiplication on $T(V)$ for having the Hopf structure.
\begin{prop}
Let $\Delta_V\colon T^M(V)\to T^M(V)\otimes T^M(V)$ be a natural transformation such that $T^M(V)$ with the usual multiplication together with the comultiplication given by $\Delta_V$ is a quasi-Hopf algebra for every $V$. Then $\Delta_V=\psi_V$ for all $V$.
\end{prop}
\begin{proof}
For every $V$, from the property that $\mathrm{H o m}(T_n,T_m)=0$ for $n\not=m$, we have
$$
\Delta_V(T_1(V))\subseteq T_1(V)\otimes T_0(V)\oplus T_0(V)\otimes T_1(V)
$$
and the counit
$$
\epsilon_V \colon T(V)=\bigoplus_{n=0}^\infty T_n(V)\longrightarrow T_0(V)
$$
is the canonical projection by sending each $T_i(V)$ to $0$ for $i>0$ and $\epsilon|_{T_{0}(V)}=\mathrm{id}$. Since both $\Delta_V$ and $\psi_V$ has the counit uniquely given by $\epsilon_V$, we have
$$
\Delta_V|_{T_1(V)}=\psi_V|_{T_1(V)}.
$$
It follows that $\Delta_V=\psi_V$ because both of them are algebra map with respect to formal tensor product.
\end{proof}
\begin{rem}
Given a module $V$, of course one could have many comultiplication on $T(V)$ such that $T(V)$ is Hopf. The proposition states that $\psi$ is the only one possible comultiplication on $T(V)$ as a natural transformation.
\end{rem}
\subsection{Subfunctors of the Tensor Algebra Functor}\label{subsection2.2}
Let $\mathcal{C}$ and $\mathcal{D}$ be categories and let $A,B\colon \mathcal{C}\to \mathcal{D}$ be functors. We call $A$ is a \textit{subfunctor} (\textit{quotient functor}) of $B$ if there is a natural transformation $\phi\colon A\to B$ such that
$$
\phi_X\colon A(X)\to B(X)
$$
is injective (surjective) for every object $X\in\mathcal{C}$. A subfunctor (quotient functor) of $T$ refers to a subfunctor (quotient functor) of $T^H$, $T^C$ or $T^M$. A subfunctor (quotient functor) of $T^H$ is called a \textit{sub Hopf functor} (\textit{quotient Hopf functor}) of $T$. Similarly we have \textit{sub coalgebra functor} (\textit{quotient coalgebra functor}) of $T$ and \textit{submodule functor} (\textit{quotient module functor}) of $T$. A \textit{graded subfunctor} (\textit{graded quotient functor}) of $T$ refers to a subfunctor of $T^H$, $T^C$ or $T^M$ as functors from modules to graded Hopf algebras, graded coalgebras or graded modules, respectively.
\begin{prop}\label{proposition2.3}
Let $B$ be a subfunctor of $T$. Then $B$ is a graded subfunctor of ~$T$.
\end{prop}
\begin{proof}
Let $\phi_V\colon B(V)\to T(V)$ be the natural monomorphism and let $$B_n(V)=\phi_V^{-1}(T_n(V))$$ for any $V$. From the fact that $T=\bigoplus\limits_{n=0}^\infty T_n$ is an orthogonal decomposition, the functor $B$ admits a graded decomposition $$B=\bigoplus\limits_{n=0}^\infty B_n.$$
Thus if $B$ is a subfunctor of $T^M$, then $B$ is graded subfunctor of $T^M$. By the orthogonal property of $T$ as in equation~(\ref{equation2.2}),
$$
\mathrm{H o m}(B_n,B_m)=0
$$
for $n\not=m$. Thus if $B$ is a subfunctor of $T^H$ or $T^C$, then the multiplication and comultiplication on $B$ whence it is defined as a natural transformation must be graded and hence the result.
\end{proof}
\begin{cor}\label{corollary2.4}
Let $C$ be a quotient functor of $T$. Then $C$ is a graded quotient functor of $T$.
\end{cor}
\begin{proof}
Let $B$ be the kernel of $T^M\twoheadrightarrow C$ by forgetting the possible Hopf or coalgebra structure on $C$. By Proposition~\ref{proposition2.3}, $B$ is a graded subfunctor of $T^M$ and so $C$ is a graded quotient functor of $T^M$, where $C_n=T_n/B_n$ for each $n\geq 0$. By the orthogonal property of $T$ as in equation~(\ref{equation2.2}),
$$
\mathrm{H o m}(C_n,C_m)=0
$$
for $n\not=m$. Thus if $C$ is a quotient functor of $T^H$ or $T^C$, then the multiplication and comultiplication on $C$ whence it is defined as a natural transformation must be graded and hence the result.
\end{proof}
\subsection{The Associated Symmetric Group Modules of the Functors}\label{subsection2.3}
Let $\bar V_n$ be the $n$-dimensional $\ensuremath{{\mathbf{k}}}$-module with a fixed choice of basis $\{x_1,\ldots,x_n\}$. For each $1\leq i\leq n$, define the linear transformation
$$
d_i\colon \bar V_n\longrightarrow \bar V_{n-1}
$$
by setting
$$
d_i(x_j)=\left\{
\begin{array}{lcl}
x_j&\textrm{ if }& j<i,\\
0&\textrm{ if }&j=i,\\
x_{j-1}&\textrm{ if }& j>i.\\
\end{array}\right.
$$
The right $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-action on $\bar V_n$ is given by $$x_i\cdot \sigma=x_{\sigma(i)}$$ for $1\leq i\leq n$ and $\sigma\in \Sigma_n$.
Then, for each $1\leq i\leq n$ and any $\sigma\in\Sigma_n$, clearly there exists a unique permutation $d_i\sigma\in \Sigma_{n-1}$ such that the diagram
\begin{equation}\label{equation2.3}
\begin{diagram}
\bar V_n&\rTo^{\sigma}_{x_j\mapsto x_{\sigma(j)}}&\bar V_n\\
\dTo>{d_i}&&\dTo>{d_{\sigma(i)}}\\
\bar V_{n-1}&\rTo^{d_i\sigma}&\bar V_{n-1}\\
\end{diagram}
\end{equation}
commutes. Let $B$ be a functor from modules to modules. Then $B(\bar V_n)$ is a right $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module induced by the action of $\ensuremath{{\mathbf{k}}}(\Sigma_n)$ on $\bar V_n$. Define
\begin{equation}\label{equation2.4}
\gamma_n(B)=\bigcap_{i=1}^n\mathrm{K er} (B(d_i)\colon B(\bar V_n)\to B(\bar V_{n-1})).
\end{equation}
By applying the functor $B$ to diagram~(\ref{equation2.3}), $\gamma_n(B)$ is a $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-submodule of $B(\bar V_n)$. Let
$$
\phi\colon B\to C
$$
be a natural transformation of functors from modules to modules. Then clearly $\phi_{\bar V_n}$ induces a $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-map
$$
\gamma_n(\phi)\colon \gamma_n(B)\longrightarrow \gamma_n(C).
$$
\begin{prop}\label{proposition2.5}
Let
$$
A\rInto^{j} B\rOnto^{p} C
$$
be a short exact sequence of functors from modules to modules. Then there is a short exact sequence of $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-modules
$$
\gamma_n(A)\rInto^{\gamma_n(j)}\gamma_n(B)\rOnto^{\gamma_n(p)}\gamma_n(C).
$$
Thus $\gamma_n(-)$ is an exact functor from the category of functors from modules to modules to the category of $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-modules.
\end{prop}
\textbf{Note.} The exact functor $\gamma_n(-)$ is a variation of the Schur functor given in~\cite{Green} in the following sense: Let $B$ is a sub-quotient functor of $T_n$ and let $V$ be a module with $m=\mathrm{dim\,} V\geq n$. Then $B(V)$ is a sub-quotient $\ensuremath{{\mathbf{k}}}(GL_m(\ensuremath{{\mathbf{k}}}))$-module of $V^{\otimes n}$. Let $\bar V_n$ embed into $V$ in the canonical way such that $V=\bar V\oplus V'$. In our definition, $\gamma_n(B)\subseteq B(\bar V)\subseteq B(V)$. According to~\cite[Section 1.2, p.71]{Donkin-Erdmann}, $B(V)\mapsto \gamma_n(B)$ is the Schur functor.
\begin{proof}
Define coface operation $d^i\colon \bar V_{n-1}\to \bar V_n$ by setting $$
d^i(x_j)=\left\{
\begin{array}{lcl}
x_j&\textrm{ if } & j<i\\
x_{j+1}&\textrm{ if }& j\geq i\\
\end{array}\right.
$$
for $1\leq i\leq n$. Then the sequence of module $\{\bar V_{n+1}\}_{n\geq 0}$ with faces $$d_1,\ldots,d_n\colon \bar V_{n}\to \bar V_{n-1}$$ relabeled as $d_0,\ldots,d_{n-1}$ and cofaces $$d^1,\ldots,d^n\colon \bar V_{n-1}\to \bar V_n$$ relabeled as $d^0,\ldots,d^{n-1}$ by shifting indices down by $1$ forms a bi-$\Delta$-groups in the sense of~\cite[Section 1.2]{Wu2}. By applying the functors to the bi-$\Delta$-group $\{\bar V_{n+1}\}_{n\geq 0}$, one gets a short exact sequence of bi-$\Delta$-groups
$$
\{A(\bar V_{n+1})\}_{n\geq 0}\rInto \{B(\bar V_{n+1})\}_{n\geq 0}\rOnto \{C(\bar V_{n+1})\}_{n\geq 0}.
$$
The assertion then follows by~\cite[Proposition 1.2.10]{Wu2}.
\end{proof}
\begin{cor}\label{corollary2.6}
Let $\phi\colon A\to B$ be a natural transformation between functors from modules to modules. Suppose that
$$
\gamma_n(\phi)\colon \gamma_n(A)\to \gamma_n(B)
$$
is an isomorphism for each $n\geq1$. Then
$$
\phi_V\colon A(V)\to B(V)
$$
is an isomorphism for any finite dimensional module $V$. Thus if both $A$ and $B$ preserve colimits, then $\phi$ is a natural equivalence.
\end{cor}
\begin{proof}
Let $C$ be the cokernel of $\phi$. Suppose that $C(V)\not=0$ for some finite dimensional module $V$. Let
$$
n={\mathrm{min}}\{ k\ | \ C(V)\not=0\ \mathrm{dim}(V)=k\}.
$$
Then $\gamma_n(C)=C(\bar V_n)\not=0$. By Proposition~\ref{proposition2.5}, $\gamma_n(C)=0$ which is a contradiction. Thus $C(V)=0$ for any finite dimensional module $V$. Similarly, for $D$ the kernel of $\phi$, we have $D(V)=0$ for any finite dimensional module $V$, finishing the proof.
\end{proof}
\subsection{$T_n$-projective functors}\label{subsection2.4}
Consider the functor $T_n$. Let $\gamma_n=\gamma_n(T_n)$. Then $\gamma_n$ is the $\ensuremath{{\mathbf{k}}}$-submodule of $\bar V_n^{\otimes n}$ spanned by the monomials
$$
x_{\sigma(1)}\otimes\cdots \otimes x_{\sigma(n)}
$$
for $\sigma\in\Sigma_n$ with the right symmetric group action explicitly given by
\begin{equation}\label{equation2.5}
(x_{i_1}\otimes \cdots\otimes x_{i_n})\cdot \sigma=x_{\sigma(i_1)}\otimes \cdots\otimes x_{\sigma(i_n)}
\end{equation}
for $\sigma\in \Sigma_n$ and the monomials $x_{i_1}\cdots x_{i_n}\in \gamma_n$. Observe that
\begin{equation}\label{equation2.6}
\gamma_n\cong \ensuremath{{\mathbf{k}}}(\Sigma_n)
\end{equation}
as a $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module.
Let $V$ be any $\ensuremath{{\mathbf{k}}}$-module and let $a_1,\ldots,a_n\in V$. We write $a_1\cdots a_n$ for the tensor product $a_1\otimes\cdots\otimes a_n\in V^{\otimes n}$ if there are no confusions. Let the symmetric group $\Sigma_n$ act on $V^{\otimes n}$ by permuting positions. More precisely the left $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-action on $V^{\otimes n}$ is given
\begin{equation}\label{equation2.7}
\sigma\cdot (a_1\cdots a_n)=a_{\sigma(1)}\cdots a_{\sigma(n)}
\end{equation}
for $\sigma\in \Sigma_n$ and the monomials $a_1\cdots a_n\in V^{\otimes n}$. Let $B$ be any functor from modules to modules. Define the functor $\gamma_n^B(-)$ by setting
\begin{equation}\label{equation2.8}
\gamma_n^B(V)=\gamma_n(B)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}
\end{equation}
for any module $V$. Clearly
$$
\gamma_n^{T_n}(V)=\gamma_n\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\cong T_n(V).
$$
\begin{prop}\label{proposition2.7}
Let
$$B\rInto^{j} T_n\rOnto^p C $$
be a short exact sequence of functors from modules to modules. Then the natural isomorphism $\gamma_n\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\cong T_n(V)$ induces a natural commutative diagram of exact sequences
\begin{diagram}
\gamma_n(B)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)} V^{\otimes n} &\rTo^{\gamma_n(j)\otimes\mathrm{id}} &\gamma_n\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)} V^{\otimes n}&\rOnto^{\gamma_n(p)\otimes\mathrm{id}} &\gamma_n(C)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\\
\dTo&&\dTo>{\cong}&&\dTo\\
B(V)&\rInto& T_n(V)&\rOnto& C(V)\\
\end{diagram}
with a natural exact sequence
$$
\mathrm{Tor}^{\ensuremath{{\mathbf{k}}}(\Sigma_n)}_1(\gamma_n(C), V^{\otimes n})\hookrightarrow \gamma_n(B)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n} \rightarrow B(V)\rightarrow \gamma_n(C)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\twoheadrightarrow C(V).
$$
for any module $V$.
\end{prop}
\begin{proof}
By taking the image of $j_V$, we may consider $B(V)$ is a submodule of $T_n(V)$ for any module $V$. Let
$$
\theta\colon \gamma_n\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)} V^{\otimes n}\to T_n(V)
$$
be the isomorphism and let $\Phi^B_V=\theta\circ (\gamma_n(j)\otimes \mathrm{id}).$ Observe that the isomorphism $\theta$ is given by
$$
\theta(x_1\cdots x_n\otimes a_1\cdots a_n)=a_1\cdots a_n
$$
for $a_1,\ldots,a_n\in V$. Let $a=a_1\cdots a_n\in V^{\otimes n}$ be any monomial with $a_j\in V$ for $1\leq j\leq n$ and let
$$
\alpha=\sum_{\sigma\in\Sigma_n}k_{\sigma}x_{\sigma(1)}\cdots x_{\sigma(n)}\in \gamma_n(B).
$$
Then
$$
\begin{array}{rcl}
\Phi^B_V(\alpha\otimes a_1\cdots a_n)&=&\sum\limits_{\sigma\in\Sigma_n} k_{\sigma} x_{\sigma(1)}\cdots x_{\sigma(n)}\otimes a_1\cdots a_n\\
&=&\sum\limits_{\sigma\in\Sigma_n} k_{\sigma} (x_{1}\cdots x_{n})\cdot\sigma\otimes a_1\cdots a_n\\
&=&\sum\limits_{\sigma\in\Sigma_n} k_{\sigma} x_{1}\cdots x_{n}\otimes \sigma\cdot(a_1\cdots a_n)\\
&=&\sum\limits_{\sigma\in\Sigma_n}k_{\sigma} x_{1}\cdots x_{n}\otimes a_{\sigma(1)}\cdots a_{\sigma(n)} \\
&=&\sum\limits_{\sigma\in\Sigma_n} k_{\sigma} a_{\sigma(1)}\cdots a_{\sigma(n)}\in V^{\otimes n}.\\
\end{array}
$$
Define a linear transformation $f_a\colon \bar V_n\to V$ by setting
$$
f_a(x_i)=a_i
$$
for $1\leq i\leq n$. Consider $T_n(f_a)=f_a^{\otimes n}\colon T_n(\bar V_n)\to T_n(V)$. Then
$$
T_n(f_a)(\alpha)=f_a^{\otimes n}\left(\sum_{\sigma\in\Sigma_n}k_{\sigma}x_{\sigma(1)}\cdots x_{\sigma(n)}\right)=\Phi^B_V(\alpha\otimes a_1\cdots a_n).
$$
From the commutative diagram
\begin{diagram}
B(\bar V_n)&\rInto^{j_{\bar V_n}}& T_n(\bar V_n)\\
\dTo>{B(f_a)}&&\dTo>{T_n(f_a)}\\
B(V)&\rInto^{j_V}& T_n(V),\\
\end{diagram}
since
$$
\alpha\in \gamma_n(B)\subseteq B(\bar V_n),
$$
we have
\begin{equation}\label{equation2.9}
\Phi^B_V(\alpha\otimes a_1\cdots a_n)=T_n(f_a)(\alpha)\in B(V).
\end{equation}
It follows that
$$
\mathrm{Im}(\Phi^B_V)\subseteq B(V)
$$
for any module $V$. Thus we have the left square commutative diagram in the statement.
By Proposition~\ref{proposition2.5}, there is a short exact sequence of $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-modules
$$
\gamma_n(B)\rInto^{\gamma_n(j)} \gamma_n\rOnto^{\gamma_n(p)} \gamma_n(C).
$$
Since $\gamma_n$ is a free $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module, there is a exact sequence
$$
\mathrm{Tor}_1^{\ensuremath{{\mathbf{k}}}(\Sigma_n)}(\gamma_n(C),V^{\otimes n})\hookrightarrow
\gamma_n(B)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n} \to \gamma_n\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n} \twoheadrightarrow \gamma_n(C)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)} V^{\otimes n}.
$$
Hence we have the right square of the commutative diagram as well as the exact sequence in the statement.
\end{proof}
\begin{example}\label{example2.8}
{\rm We give an example that the natural transformation
$$
\gamma_n(B)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n} \longrightarrow B(V)
$$
could be neither epimorphism nor monomorphism for subfunctors $B$ of $T_n$. Let $\ensuremath{{\mathbf{k}}}$ be a field of characteristic $2$. The Lie power
$L_2(V)$ is the submodule of $V^{\otimes 2}$ spanned by $[a,b]=ab-ba$ for $a,b\in V$. The restricted Lie power $L^{\mathrm{res}}_2(V)$ is the submodule of $V^{\otimes 2}$ spanned by $a^2, [a,b]$ for $a,b\in V$. Then
$$
\gamma_2(L_2)=\gamma_2(L^{\mathrm{res}}_2)
$$
is the $1$-dimensional submodule of $\ensuremath{{\mathbf{k}}}(\Sigma_2)$ generated by $1-\tau$. Since $\ensuremath{{\mathbf{k}}}$ is of characteristic $2$, $\gamma_2(L_2)=\gamma_2(L^{\mathrm{res}}_2)$ is the trivial $\ensuremath{{\mathbf{k}}}(\Sigma_2)$-module and so
$$
\gamma_2(L^{\mathrm{res}}_2)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_2)}V^{\otimes 2}=S_2(V),
$$
the $2$-fold symmetric product of $V$. The natural transformation
$$
\gamma_2(L^{\mathrm{res}}_2)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_2)}V^{\otimes 2}\longrightarrow L^{\mathrm{res}}_2(V)
$$
is not epimorphism for $V$ with $\mathrm{dim\,} V\geq 1$ because its image is given by $L_2(V)$. The kernel of this natural transformation is measured by
$$
\mathrm{Tor}_1^{\ensuremath{{\mathbf{k}}}(\Sigma_2)}(\gamma_2/\mathrm{Lie}(2), V^{\otimes 2})\not=0
$$
for $V$ with $\mathrm{dim\,} V\geq 1$.
\hfill $\Box$
}\end{example}
Let $B$ be a functor from modules to modules. The \textit{dual functor $B^*$} is defined as follows. For any finite dimensional module $V$, define
$$
B^*(V)=B(V^*)^*,
$$
where $V^*=\mathrm{H o m}_{\ensuremath{{\mathbf{k}}}}(V,\ensuremath{{\mathbf{k}}})$ is the dual $\ensuremath{{\mathbf{k}}}$-module of $V$, and, for a general module $V$, let
$$
B^*(V)=\operatorname{co l i m}_{V_{\alpha}}B^*(V_{\alpha})
$$
be the direct limit of the module $B^*(V_{\alpha})$ subject to the direct system given by the diagram of all finite dimensional submodules of $V$ with inclusions. Clearly $T_n^*=T_n$.
\begin{prop}\label{proposition2.9}
Let $B$ be a functor from modules to modules. Then $\gamma_n(B^*)$ is the dual $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module of $\gamma_n(B)$ for each $n\geq 1$.
\end{prop}
\begin{proof}
For the basis $\{x_1,\ldots,x_n\}$ for $\bar V_n$, let $\{x_1^*,\ldots,x_n^*\}$ be the standard dual basis of $\bar V_n^*$. Let
$$
\theta_n\colon \bar V_n\to \bar V^*_n
$$
be the linear transformation such that $\theta_n(x_j)=x_j^*$ for $1\leq j\leq n$. Then it is routine to check that the composite
$$
\gamma_n(B^*)\subseteq B^*(\bar V_n)=B(\bar V_n^*)^*\rTo^{B(\theta_n)^*} B(\bar V_n)^*\rOnto \gamma_n(B)^*
$$
is an isomorphism of $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-modules.
\end{proof}
We call a direct sum of copies of $T_n$ a \textit{free $T_n$-functor}. A functor $B$ from modules to modules is called \textit{$T_n$-projective} if there exists a free $T_n$-functor $F$ together with natural transformations $s\colon B\to F$ and $r\colon F\to B$ such that $r\circ s\colon B\to B$ is a natural equivalence. In other words, a $T_n$-projective functor means a summand (or retract) of a free $T_n$-functor.
\begin{prop}\label{proposition2.10}
Let $B$ be a functor from modules to modules.
\begin{enumerate}
\item[1)] If $B$ is a $T_n$-projective functor, then $\gamma_n(B)$ is a projective $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module and there is a natural isomorphism
$$
\gamma_n(B)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\cong B(V)
$$
for any module $V$.
\item[2)] If $B$ is a subfunctor of a direct sum of finite copies of $T_n$ with the property that $\gamma_n(B)$ is a projective $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module, then $B$ is a $T_n$-projective functor. Moreover there is a natural equivalence $B\cong B^*$.
\item[3)] If $B$ is a quotient functor of a direct sum of finite copies of $T_n$ with the property that $\gamma_n(B)$ is a projective $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module, then $B$ is a $T_n$-projective functor. Moreover there is a natural equivalence $B\cong B^*$.
\item[4)] Let $B$ be a sub-quotient functor of a direct sum of finite copies of $T_n$. Suppose that $B$ is a $T_n$-projective functor. Then $B$ is both projective and injective in the category of sub-quotient functors of direct sums of finite copies of $T_n$.
\end{enumerate}
\end{prop}
\begin{proof}
The proof of assertion~(1) is straightforward. Assertion (3) follows from (2) by considering the dual functor.
(2). Let $B$ be a subfunctor of $F$, where $F$ is a finite direct sum of copies of $T_n$. Let $C=F/B$. Then there is a short exact sequence
$$
\gamma_n(B)\rInto \gamma_n(F)\rOnto \gamma_n(C).
$$
Since $\gamma_n(B)$ is a finitely generated projective $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module, it is an injective $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module and so the above short exact sequence splits off as $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-modules. It follows that $\gamma_n(C)$ is a projective $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module because $\gamma_n(F)$ is a free $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module. $$
\mathrm{Tor}^{\ensuremath{{\mathbf{k}}}(\Sigma_n)}_1(\gamma_n(C),V^{\otimes n})=0.
$$
Since $F$ is a direct sum of copies of the functor $T_n$, we can apply the exact sequence in Proposition~\ref{proposition2.7}. In particular, the natural transformation
\begin{equation}\label{equation2.10}
\Phi^B_V\colon \gamma_n(B)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\longrightarrow B(V)
\end{equation}
is a natural monomorphism. The natural inclusion $B\hookrightarrow F$ induces an natural epimorphism $F=F^*\twoheadrightarrow B^*$. By Proposition~\ref{proposition2.7}, there is a natural epimorphism
\begin{equation}\label{equation2.11}
\Phi^{B^*}_V\colon \gamma_n(B^*)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n} \longrightarrow B^*(V).
\end{equation}
By Proposition~\ref{proposition2.9}, $$\gamma_n(B^*)\cong\gamma_n(B)^*\cong \gamma_n(B)$$ as $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-modules because $\gamma_n(B)$ is a finitely generated $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-projective modules. Let $V$ be any finite dimensional $\ensuremath{{\mathbf{k}}}$-module. From equations~(\ref{equation2.10}) and~(\ref{equation2.11}, we have
$$
\mathrm{dim\,} B(V)=\mathrm{dim\,} B(V^*)^*=\mathrm{dim\,} B^*(V)\leq \mathrm{dim\,} (\gamma_n(B)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n})\leq \mathrm{dim\,} B(V).
$$
Thus $\Phi^B_V$ and $\Phi^{B^*}_V$ are isomorphisms for any finite dimensional module $V$. Since the functors $\gamma_n(B)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}(-)^{\otimes n}$, $B$ and $B^*$ preserve colimits, the natural transformations $\Phi^B$ and $\Phi^{B^*}$ are natural equivalences. The assertion now follows from the fact that $\gamma_n(B)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}$ is a natural summand of
$
\gamma_n(F)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\cong F(V).
$
(4). Let $\mathcal{C}$ be the category of sub-quotient functors of direct sums of copies of $T_n$. It suffices to show that $T_n$ is projective and injective in the $\mathcal{C}$. Let $B$ be an object in $\mathcal{C}$ with a natural epimorphism
$q\colon B \twoheadrightarrow T_n$. It induces an epimorphism $\gamma_n(q)\colon \gamma_n(B)\to \gamma_n(T_n)=\gamma_n$. Since $\gamma_n$ is $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-projective, there is a $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-cross-section $s\colon \gamma_n(T_n)\to \gamma_n(B)$. Now the natural transformation
$$
T_n(V)\cong \gamma_n(T_n)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\rTo^{s\otimes\mathrm{id}} \gamma_n(B)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\rTo^{\Phi^B_V} B(V)
$$
is a cross-section to $q$ and so $T_n$ is projective in $\mathcal{C}$. Since $T_n\cong T_n^*$ is self-dual, $T_n$ is also injective in $\mathcal{C}$. The proof is finished.
\end{proof}
We remark that if $B$ is a sub-quotient functor of $T_n$ with the property that $\gamma_n(B)$ is $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-projective, it is possible that $B$ is not $T_n$-projective. For instance, for the ground field $\ensuremath{{\mathbf{k}}}$ being of characteristic $2$, the functor $B=L^{\mathrm{res}}_2/L_2$ has the property that $\gamma_2(B)=0$ with $B$ not $T_2$-projective.
\section{The Structure on Lie Power Functors}\label{section3}
\subsection{The Lie Power Functors and the Symmetric Group Modules $\mathrm{Lie}(n)$}\label{subsection3.1}
In this section, the ground ring is a field $\ensuremath{{\mathbf{k}}}$. Let $V$ be a module. The \textit{free Lie algebra $L(V)$} generated by $V$ is the smallest sub Lie algebra of the tensor algebra $T(V)$ containing $V$, where the Lie structure on $T(V)$ is given by $[a,b]=ab-ba$. The functor $L$ admits a graded structure that
$$
L(V)=\bigoplus_{n=1}^\infty L_n(V),
$$
where $L_n(V)=L(V)\cap T_n(V)$ which is called the \textit{$n$th Lie power} of $V$. By applying equation~(\ref{equation2.4}) to the functor $L_n$, we have the symmetric group module
$$
\mathrm{Lie}(n)=\gamma_n(L_n).
$$
Let $\bar V$ be the $n$-dimensional module with a basis $\{x_1,\ldots,x_n\}$ as in subsection~\ref{subsection2.3}. From the definition,
$$
\mathrm{Lie}(n)=L_n(\bar V_n)\cap\gamma_n
$$
spanned by the homogenous Lie elements of length $n$ in which each $x_i$ occurs exactly once. By
the Witt formula, $\mathrm{Lie}(n)$ is of dimension $(n-1)!$. Following from the antisymmetry and the Jacobi identity, $\mathrm{Lie}(n)$
has a basis given by the elements
$$
[[x_1,x_{\sigma(2)}],x_{\sigma(3)}],\ldots,x_{\sigma(n)}]
$$
for $\sigma\in \Sigma_{n-1}$. (See ~\cite{Cohen2}.)
\begin{prop}\label{proposition3.1}
There is a natural short exact sequence
$$
\mathrm{Tor}^{\ensuremath{{\mathbf{k}}}(\Sigma_n)}_1(\gamma_n/\mathrm{Lie}(n), V^{\otimes n}) \rInto \mathrm{Lie}(n)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\rOnto L_n(V)
$$
for any module $V$.
\end{prop}
\begin{proof}
By Proposition~\ref{proposition2.7}, it suffices to show that the natural transformation
$$
\Phi^{L_n}_V\colon \mathrm{Lie}(n)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n} \longrightarrow L_n(V)
$$
is an epimorphism. Let $[[a_1,a_2],\ldots a_n]\in L_n(V)$ with $a_1,\ldots,a_n\in V$. Let
$$
\alpha=[[x_1,x_2],\ldots,x_n]\in \mathrm{Lie}(n).
$$
Then exists unique $k_{\sigma}\in\ensuremath{{\mathbf{k}}}$ such that
$$
\alpha=[[x_1,x_2],\ldots,x_n]=\sum_{\sigma\in \Sigma_n}k_{\sigma}x_{\sigma(1)}\cdots x_{\sigma(n)}.
$$
Following the lines in the proof of Proposition~\ref{proposition2.7}, we have
\begin{equation}\label{equation3.1}
\Phi^{L_n}_V(\alpha\otimes a_1\cdots a_n)=\sum_{\sigma\in \Sigma_n}k_{\sigma}a_{\sigma(1)}\cdots a_{\sigma(n)}=[[a_1,a_2],\ldots,a_n].
\end{equation}
The assertion follows from the fact that $L_n(V)$ is the $\ensuremath{{\mathbf{k}}}$-module spanned by the Lie elements $[[a_1,a_2],\ldots,a_n]$ with $a_j\in V$.
\end{proof}
For any natural transformation $\phi\colon L_n\to L_n$, we have the $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-linear map
$$
\gamma_n(\phi)\colon \gamma_n(L_n)=\mathrm{Lie}(n)\longrightarrow \gamma_n(L_n)=\mathrm{Lie}(n).
$$
This defines a ring homomorphism
$
\gamma\colon \mathrm{End}(L_n)\longrightarrow \mathrm{End}_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}(\mathrm{Lie}(n)).
$
\begin{prop}\label{proposition3.2}
If $n\not=m$, then $\mathrm{H o m}(L_n,L_m)=0$. Moreover the ring homomorphism $$
\gamma\colon \mathrm{End}(L_n)\rTo \mathrm{End}_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}(\mathrm{Lie}(n))
$$
is an isomorphism with a natural commutative diagram of functors
\begin{diagram}
\mathrm{Lie}(n)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)} V^{\otimes n}&\rTo^{\gamma_n(\delta)\otimes\mathrm{id}}& \mathrm{Lie}(n)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\\
\dOnto>{\Phi^{L_n}_V}&&\dOnto>{\Phi^{L_n}_V}\\
L_n(V)&\rTo^{\delta_V}& L_n(V)\\
\end{diagram}
for any natural transformation $\delta\colon L_n\to L_n$.
\end{prop}
\begin{proof}
Let $\phi\colon L_n\to L_m$ be a natural transformation. Let
$\bar\beta_n\colon T_n\to L_n$ be the natural epimorphism defined by
$
\bar\beta_n(a_1\cdots a_n)=[[a_1,a_2],\ldots,a_n]
$
for any module $W$ and any monomial $a_1\cdots a_n\in T_n(W)=W^{\otimes n}$. Then the
composite $$T_n\rOnto^{\bar \beta} L_n\rTo^\phi L_m\rInto T_m$$ is a natural
transformation, which is zero as $\mathrm{H o m}(T_n,T_m)=0$ for $n\not=m$. Thus $\phi=0$.
For the second statement, let $\delta\colon L_n\to L_n$ be a natural transformation. Let $V$ be any module. Consider
$
[[a_1,a_2],\ldots,a_n]\in L_n(V)
$
with $a_j\in V$. Let $f_a\colon \bar V_n\to V$ be the linear map with $f_a(x_j)=a_j$ for $1\leq j\leq n$. Then there is a commutative diagram
\begin{equation}\label{equation3.2}
\begin{diagram}
\gamma_n(L_n)&\subseteq& L_n(\bar V_n)&\rTo^{L_n(f_a)}& L_n(V)\\
\dTo>{\gamma_n(\delta)}&&\dTo>{\delta_{\bar V_n}}&&\dTo>{\delta_V}\\
\gamma_n(L_n)&\subseteq& L_n(\bar V_n)&\rTo^{L_n(f_a)}& L_n(V).\\
\end{diagram}
\end{equation}
Thus
$$
\begin{array}{rlr}
&\delta_V\circ\Phi_V^{L_n}([[x_1,x_2],\ldots,x_n]\otimes a_1\cdots a_n) & \\
=&\delta_V([[a_1,a_2],\ldots,a_n]) &\textrm{ by equation~(\ref{equation3.1})}\\
=&L_n(f_a)(\gamma_n(\delta)([[x_1,x_2],\ldots,x_n]))&\textrm{ by equation~(\ref{equation3.2})}\\
=&\Phi_V^{L_n}(\gamma_n(\delta)([[x_1,x_2],\ldots,x_n])\otimes a_1\cdots a_n)&\textrm{ by equation~(\ref{equation2.9})}\\
\end{array}
$$
and so the diagram in the statement commutes. It follows that the map $$\gamma\colon \mathrm{End}(L_n)\to \mathrm{End}_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}(\mathrm{Lie}(n))$$ is a monomorphism.
For showing that $\gamma$ is an epimorphism, let $\theta\colon \mathrm{Lie}(n)\to \mathrm{Lie}(n)$ be any $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-linear map. Since $\ensuremath{{\mathbf{k}}}(\Sigma_n)$ is a Fr\"{o}benius algebra, the free $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module $\gamma_n$ is injective and so there is a commutative diagram of exact sequences of $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-modules
\begin{diagram}
\mathrm{Lie}(n)&\rInto& \gamma_n&\rOnto&\gamma_n/\mathrm{Lie}(n)\\
\dTo>{\theta}&&\dTo>{\tilde\theta}&&\dTo>{\bar\theta}\\
\mathrm{Lie}(n)&\rInto& \gamma_n&\rOnto&\gamma_n/\mathrm{Lie}(n).\\
\end{diagram}
It follows that there is a commutative diagram of short exact sequence of functors
\begin{diagram}
\mathrm{T or}^{\ensuremath{{\mathbf{k}}}(\Sigma_n)}_1(\gamma_n/\mathrm{Lie}(n),V^{\otimes n})&\rInto& \mathrm{Lie}(n)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}&\rOnto^{\Phi^{L_n}_V}& L_n(V)\\
\dTo>{\mathrm{T or}(\bar\theta,\mathrm{id})}&&\dTo>{\theta\otimes \mathrm{id}}&&\dTo>{\delta}\\
\mathrm{T or}^{\ensuremath{{\mathbf{k}}}(\Sigma_n)}_1(\gamma_n/\mathrm{Lie}(n),V^{\otimes n})&\rInto& \mathrm{Lie}(n)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}&\rOnto^{\Phi^{L_n}_V}& L_n(V)\\
\end{diagram}
for some natural transformation $\delta\colon L_n\to L_n$. By taking $V=\bar V_n$ and restricting to the submodule $\gamma_n\subseteq \bar V^{\otimes n}$, we have the commutative diagram
\begin{diagram}
\mathrm{Lie}(n)&\rTo^{g}_\cong&\mathrm{Lie}(n)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}\gamma_n&\rTo^{\Phi^{L_n}_{\bar V_n}|}_{\cong}&\mathrm{Lie}(n)\\
\dTo>{\theta}&&\dTo>{\theta\otimes\mathrm{id}}&&\dTo>{\gamma_n(\delta)}\\
\mathrm{Lie}(n)&\rTo^{g}_\cong&\mathrm{Lie}(n)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}\gamma_n&\rTo^{\Phi^{L_n}_{\bar V_n}|}_{\cong}&\mathrm{Lie}(n),\\
\end{diagram}
where $g(\alpha)=\alpha\otimes x_1\cdots x_n$. Thus $\theta=\gamma_n(\delta)$
because
$$
\begin{array}{rcl}
\Phi^{L_n}_{\bar V}\circ g([[x_{\sigma(1)},x_{\sigma(2)}],\ldots,x_{\sigma(n)}])&=& \Phi^{L_n}_{\bar V}([[x_{1},x_{2}],\ldots,x_{n}]\cdot\sigma\otimes x_1\cdots x_n)\\
&=& \Phi^{L_n}_{\bar V}([[x_1,x_2],\ldots,x_n]\otimes \sigma\cdot(x_1\cdots x_n))\\
&=& \Phi^{L_n}_{\bar V}([[x_1,x_2],\ldots,x_n]\otimes x_{\sigma(1)}\cdots x_{\sigma(n)})\\
&=&[[x_{\sigma(1)},x_{\sigma(2)}],\ldots,x_{\sigma(n)}].\\
\end{array}
$$
The proof is finished.
\end{proof}
\begin{cor}\label{corollary3.3}
There is a one-to-one correspondence, multiplicity preserving, between the decompositions of
the functor $L_n$ and the decompositions of $\mathrm{Lie}(n)$ over
$\ensuremath{{\mathbf{k}}}(\Sigma_n)$.\hfill $\Box$
\end{cor}
\subsection{The $T_n$-projective Subfunctors of $L_n$}\label{subsection3.2}
Let $Q$ be a subfunctor of $L_n$. Then $Q$ is a subfunctor of $T_n$ because $L_n$ is a subfunctor of $T_n$. By Proposition~\ref{proposition2.10}, the functor $Q$ is $T_n$-projective if and only if $\gamma_n(Q)$ is a $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-projective module. From Corollary~\ref{corollary3.3}, we have the following:
\begin{prop}\label{proposition3.4}
There is a one-to-one correspondence, multiplicity preserving, between $T_n$-projective sub
functors of $L_n$ and $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-projective submodules of
$\mathrm{Lie}(n)$ given by $Q\mapsto \gamma_n(Q)$.\hfill $\Box$
\end{prop}
According to~\cite[Lemma 6.2 and Theorem 7.4]{SW1}, there exists a subfunctor $L^{\mathrm{max}}_n$ of $L_n$ with
$\mathrm{Lie}^{\mathrm{max}}(n)=\gamma_n(L_n^{\mathrm{max}})$ that has the following maximum property:
\begin{enumerate}
\item[1)] $\mathrm{Lie}^{\mathrm{max}}(n)$ is a $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-projective submodule of $\mathrm{Lie}(n)$ and
\item[2)] any $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-projective submodule of $\mathrm{Lie}(n)$ is isomorphic to a summand of $\mathrm{Lie}^{\mathrm{max}}(n)$ as a $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-module.
\end{enumerate}
From the above maximum property, $\mathrm{Lie}^{\mathrm{max}}(n)$ is unique up to isomorphisms of $\ensuremath{{\mathbf{k}}}(\Sigma_n)$-modules. By the above proposition, $L^{\mathrm{max}}_n$ is unique up to natural equivalences with the maximum property that
\begin{enumerate}
\item[1)] $L^{\mathrm{max}}_n$ is a $T_n$-projective subfunctor of $L_n$ and
\item[2)] any $T_n$-projective subfunctor of $L_n$ is isomorphic to a summand of $L^{\mathrm{max}}_n$.
\end{enumerate}
From Proposition~\ref{proposition2.10}, we have the following:
\begin{prop}\label{proposition3.5}
There is a natural isomorphism
$$
\Phi^{L^{\mathrm{max}}_n}_V\colon \mathrm{Lie}^{\mathrm{max}}(n)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\longrightarrow L^{\mathrm{max}}_n(V)
$$
for any module $V$.\hfill $\Box$
\end{prop}
\subsection{The $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-module $L_n(V)$}\label{subsection3.3}
In this subsection, the ground field $\ensuremath{{\mathbf{k}}}$ is an infinite field of characteristic $p>0$ and $V$ is a fixed $\ensuremath{{\mathbf{k}}}$-module with the action of the general linear group $\mathrm{G L}(V)=\mathrm{G L}_m(\ensuremath{{\mathbf{k}}})$ from the right, where $m=\mathrm{dim\,} V$. Let $\mathrm{G L}(V)$ act on
$T_n(V)=V^{\otimes n}$ through diagonal, that is,
$$
(a_1\cdots a_n)\cdot g=(a_1g)\cdots (a_ng)
$$
for $a_i\in V$ and $g\in \mathrm{G L}(V)$. Recall that the \textit{Schur algebra} is defined by
$$
S(V,n)=\mathrm{End}_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}(V^{\otimes n}),
$$
where the left action of $\Sigma_n$ on $V^{\otimes n}$ is given by permuting factors. By the classical Schur-Weyl duality, the group $\mathrm{G L}(V,n)$ generates algebra $S(V,n)=\mathrm{End}_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}(V^{\otimes n})$ and so there is an epimorphism of rings
$$
\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))\longrightarrow S(V,n).
$$
Observe that if $M$ is a sub-quotient of a direct sum of copies of $V^{\otimes n}$, then the $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-action factors through its quotient algebra $S(V,n)$. Thus if $M$ and $N$ are sub-quotients of direct sums of copies of $V^{\otimes n}$, then
$$
\mathrm{H o m}_{\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))}(M,N)=\mathrm{H o m}_{S(V,n)}(M,N).
$$
Recall from~\cite{Green} that the category of $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-modules that are sub-quotients of direct sums of copies of $V^{\otimes n}$ is equivalent to the category of modules over the Schur algebra $S(V,n)$, which is denoted by $\mathrm{Mod}(S(V,n))$.
Let $B$ be a sub-quotient of a direct sum of copies of $T_n$. The action of $\mathrm{G L}(V)$ on $V$ induces an action on $B(V)$ via the functor $B$. Thus $B(V)$ is a module over $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$. Since $B(V)$ is a sub-quotient of a direct sum of copies of $V^{\otimes n}$, $B(V)$ is an object in $\mathrm{Mod}(S(V,n))$. Thus we have a functor:
$$
\begin{array}{rl}
\Theta\colon &B\mapsto B(V)\\
& \mathrm{H o m}(A, B)\longrightarrow \mathrm{H o m}_{\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))}(A(V),B(V))=\mathrm{H o m}_{S(V,n)}(A(V),B(V))\\
\end{array}
$$
from the category of sub-quotients of direct sums of copies of $T_n$ to $\mathrm{Mod}(S(V,n))$.
\begin{lem}\label{lemma3.6}
Let $B$ be a sub-quotient of a free $T_n$-functor and let $V$ be a module with $\mathrm{dim\,}(V)\geq n$. Then $B=0$ if and only if $B(V)=0$.
\end{lem}
\begin{proof}
If $B=0$, clearly $B(V)=0$. Assume that $B(V)=0$. Let $B=\tilde B/B'$ with $B'\hookrightarrow \tilde B \hookrightarrow F$, where $F$ is a direct sum of copies of $T_n$. It is routine to check that $\gamma_j(T_n)=0$ for $j>n$. Thus $\gamma_j(F)=0$ for $j>n$ and so
$$
\gamma_j(B')=\gamma_j(\tilde B)=\gamma_j(B)=0
$$
for $j>n$. Since $B(V)=0$, we have $B(\bar V_n)=0$ because $\mathrm{dim\,} V\geq \mathrm{dim\,}\bar V_n=n$. Thus $\gamma_j(B)=0$ for $j\leq n$. The assertion follows by Corollary~\ref{corollary2.6}.
\end{proof}
\begin{cor}\label{corollary3.7}
Let $A$ and $B$ sub-quotients of free $T_n$-functors and let $V$ be a module with $\mathrm{dim\,}(V)\geq n$. Then
$$
\Theta\colon \mathrm{H o m}(A,B)\longrightarrow \mathrm{H o m}_{\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))}(A(V),B(V))=\mathrm{H o m}_{S(V,n)}(A(V),B(V))
$$
is a monomorphism.
\end{cor}
\begin{proof}
Let $f\colon A\to B$ be a natural transformation such that $f_V\colon A(V)\to B(V)$ is $0$. Let $C=\mathrm{Im}(f\colon A\to B)$. Then $C(V)=0$. Thus $C=0$ and hence the result.
\end{proof}
A direct sum of finite copies of $T_n$ is called a \textit{finite free $T_n$-functor}.
\begin{prop}\label{proposition3.8}
Let $B$ be a sub-quotient of a finite free $T_n$-functor and let $A$ be a quotient functor of a finite free $T_n$-functor. Suppose that $\mathrm{dim\,} V\geq n$. Then
the homomorphism
$$
\Theta_{A,B}\colon \mathrm{H o m}(A,B)\longrightarrow \mathrm{H o m}_{\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))}(A(V),B(V))=\mathrm{H o m}_{S(V,n)}(A(V),B(V))
$$
is an isomorphism.
\end{prop}
\begin{proof}
By Schur-Weyl duality, the monomorphism
$$
\Theta_{T_n,T_n}\colon \mathrm{H o m}(T_n,T_n)\longrightarrow \mathrm{H o m}_{\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))}(T_n(V),T_n(V))
$$
is an epimorphism and so $\Theta_{A,B}$ is an isomorphism when $A$ and $B$ are free $T_n$-functors. According to~\cite[p.94]{Donkin}, $V^{\otimes n}$ is projective over $S(V,n)$. Let $A$ be a free $T_n$ functor. By tracking the exact sequence for
$$
\Theta_{A,-} \colon \mathrm{H o m}(A,-)\longrightarrow \mathrm{H o m}_{S(V,n)}(A(V), -)
$$
together with the fact that $\Theta_{A,B}$ is always a monomorphism, we have
$$
\Theta_{A,B} \colon \mathrm{H o m}(A,B)\longrightarrow \mathrm{H o m}_{S(V,n)}(A(V), B(V))
$$
for any sub-quotient $B$ of a free $T_n$-functor. Let $A$ be a quotient of a free functor $F$ with an epimorphism $\phi\colon F\to A$. Let $C=\mathrm{K er}(\phi)$. From the commutative diagram of exact sequences
\begin{diagram}
\mathrm{H o m}(A,B)&\rInto^{\phi^*}&\mathrm{H o m}(F, B)&\rTo&\mathrm{H o m}(C,B)\\
\dInto>{\Theta_{A,B}}&&\cong\dTo>{\Theta_{F,B}}&&\dInto>{\Theta_{C,B}}\\
\mathrm{H o m}_{S(V,n)}(A(V),B(V))&\rInto^{\phi^*}&\mathrm{H o m}_{S(V,n)}(F(V), B(V))&\rTo&\mathrm{H o m}_{S(V,n)}(C(V),B(V)),\\
\end{diagram}
the monomorphism $\Theta_{A,B}$ is an epimorphism and hence the result.
\end{proof}
A $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-submodule $M$ of $L_n(V)$ is called \textit{$T_n$-projective} if $M$ is isomorphic to a summand of a direct summation of $T_n(V)$ as modules over $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$. Let
$\bar\beta_n\colon T_n\to L_n$ be the natural epimorphism defined by
$
\bar\beta_n(a_1\cdots a_n)=[[a_1,a_2],\ldots,a_n]
$
for any module $W$ and any monomial $a_1\cdots a_n\in T_n(W)=W^{\otimes n}$. Let $\beta_n$ be the composite
$
T_n\rOnto^{\bar\beta_n} L_n\rInto^{i}T_n.
$
\begin{thm}\label{theorem3.9}
Suppose that $\mathrm{dim\,} V\geq n$. Then
\begin{enumerate}
\item The ring homomorphism
$$
\Theta_{L_n,L_n}\colon \mathrm{H o m}(L_n,L_n)\longrightarrow \mathrm{H o m}_{\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))}(L_n(V),L_n(V))
$$
is an isomorphism.
\item There is a one-to-one correspondence, multiplicity preserving between summands of the functor $L_n$ and $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-summands of $L_n(V)$.
\item There is a one-to-one correspondence, multiplicity preserving between $T_n$-projective subfunctors of $L_n$ and $T_n$-projective $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-submodules of $L_n(V)$.
\item For the functor $L^{\mathrm{max}}_n$, the module $L_n^{\mathrm{max}}(V)$ is the maximum $T_n$-projective submodule of $L_n(V)$ in the sense that any $T_n$-projective $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-submodule of $L_n(V)$ is isomorphic to a summand of $L_n^{\mathrm{max}}(V)$.
\item For any choice of the functor $L_n^{\mathrm{max}}$, the socle $\mathrm{Soc}(L_n^{\mathrm{max}}(V))$ is uniquely determined by
$$
\bar\beta_n(\mathrm{Soc}(V^{\otimes n}))=\beta_n(\mathrm{Soc}(V^{\otimes n})).
$$
as the submodule of $L_n(V)\subseteq T_n(V)$.
\item For any choice of the functor $L_n^{\mathrm{max}}$, the head $\mathrm{Hd}(L_n^{\mathrm{max}}(V))$ is uniquely determined by
$$
\beta_n(\mathrm{Hd}(V^{\otimes n})).
$$
as the quotient module of $T_n(V)$.
\end{enumerate}
\end{thm}
\textbf{Note.} By assertion (5), the functor $L_n^{\mathrm{max}}$ and the module $L_n^{\mathrm{max}}(V)$ are determined by evaluating the map $\bar\beta_n$ or $\beta_n$ on simple $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-submodules of $V^{\otimes n}$.
\begin{proof}
Since $L_n$ is a quotient functor of $T_n$, assertion (1) is direct consequence of Proposition~\ref{proposition3.8}. Assertion (2) follows from (1) immediately. Assertion (4) is a direct consequence of 3. The proof of assertion (6) is similar to that of assertion (5).
For proving assertion (3), let $M$ be a $T_n$-projective $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-submodule of $L_n(V)$. According to~\cite[p.94]{Donkin}, $T_n(V)$ is an injective module over $S(V,n)$ and so is $M$ because $M$ is a summand of a direct sum of finite copies of $T_n$. Thus the inclusion
$$
j\colon M \hookrightarrow L_n(V) \hookrightarrow T_n(V)
$$
admits a $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-retraction $r\colon T_n(V)\to M$. The composite $$e=j\circ r\colon T_n(V)\to T_n(V)$$ is an idempotent in
$$
\mathrm{End}_{\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))}(T_n(V)).
$$
The natural transformation $\alpha=\Theta^{e}\colon T_n\to T_n$ is an idempotent. Let $B=\mathrm{Im}(\alpha)$. Then $B$ is a $T_n$-projective subfunctor of $L_n$ with $B(V)=M$ and hence assertion (3).
(5). Let
$$
V^{\otimes n}=\bigoplus_{i\in I} P_i
$$
be a decomposition over $S(V,n)$ such that each $P_i$ is indecomposable. The map $\bar\beta_n\colon V^{\otimes n}\to L_n(V)$ induces a map
$$
\bar\beta_n\colon \mathrm{Soc}(V^{\otimes n}=\bigoplus_{i\in I} \mathrm{Soc}(P_i)\longrightarrow \mathrm{Soc}(L_n(V)).
$$
Note that each indecomposable $S(V,n)$-summand of $V^{\otimes n}$ has a unique socle. (See for instance~\cite[(6.4b)]{Green}.
Thus there exists $I'\subseteq I$ such that
$$
P=\bigoplus_{i\in I'}P_i
$$
has the property that
$$
\bar\beta_n| \colon \mathrm{Soc}(P)\longrightarrow \bar\beta_n(\mathrm{Soc}(V^{\otimes n}))
$$
is an isomorphism. It follows that
$$
\bar\beta_n|\colon P\longrightarrow L_n(V)
$$
is a monomorphism because it restricts to the socle is a monomorphism. Since $P$ is an injective $S(V,n)$-module, the map $\bar \beta_n|_P$ has a retraction. Thus $T_n$-projective $S(V,n)$-module $P$ is isomorphic to a $S(V,n)$-summand of $L_n(V)$. From the maximum property of $L^{\mathrm{max}}(n)$, $P$ is isomorphic to a $S(V,n)$-summand of $L^{\mathrm{max}}_n(V)$. In particular,
$$
\bar\beta_n(\mathrm{Soc}(V^{\otimes n}))=\bar\beta_n|(\mathrm{Soc}(P))\subseteq \mathrm{Soc}(L^{\mathrm{max}}_n(V)).
$$
On the other hand, since $L^{\mathrm{max}}_n(V)$ is $S(V,n)$-projective, the inclusion $$j\colon L^{\mathrm{max}}_n(V)\hookrightarrow L_n(V)$$ admits a $S(V,n)$-lifting $\tilde j\colon L^{\mathrm{max}}_n(V)\to V^{\otimes n}$ such that $j=\bar\beta\circ\tilde j$. Thus
$$
\mathrm{Soc}(L^{\mathrm{max}}_n(V))\subseteq \bar\beta_n(\mathrm{Soc}(V^{\otimes n})).
$$
Note that the inclusion $i\colon L_n(V)\hookrightarrow T_n(V)$ induces a monomorphism $i|\colon \mathrm{Soc}(L_n(V))\hookrightarrow \mathrm{Soc}(T_n(V))$. Thus
$$
\bar\beta_n(\mathrm{Soc}(V^{\otimes n}))=\beta_n(\mathrm{Soc}(V^{\otimes n}))
$$
and hence the result.
\end{proof}
We give a remark that if $\mathrm{dim\,} V<n$, then assertions (3) and (4) are not true by the following example.
\begin{example}\label{example3.10}
Let $\ensuremath{{\mathbf{k}}}$ be of characteristic $3$ and let $V$ be a $2$-dimensional module with a basis $\{u,v\}$. Then the canonical map
$$
f\colon L_2(V)\otimes V \longrightarrow L_3(V)\ [a_1,a_2]\otimes a_3\mapsto [[a_1,a_2],a_3]
$$
is an isomorphism of modules over $\ensuremath{{\mathbf{k}}}(\mathrm{G L}_2(\ensuremath{{\mathbf{k}}}))$. Since $L_2(V)$ is a $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-summand of $V^{\otimes 2}$, $L_2(V)\otimes V$ is $T_3$-projective. Thus $L_3(V)$ is $T_3$-projective. On the other hand, it is easy to see that the functor $L^{\mathrm{max}}_3=0$ and so $L^{\mathrm{max}}_3(V)=0$.
In this case, $L_3(V)$ is not an injective $S(V,3)$-module. In fact, the inclusion $L_3(V)\hookrightarrow V^{\otimes 3}$ does not have $S(V,3)$-retraction by expecting the Steenrod module structure on $V^{\otimes 3}$. Also it is easy to check that $L_3(V)$ is not a projective $S(V,3)$-module.\hfill $\Box$
\end{example}
We call $M\subseteq L_n(V)$ \textit{functorial $T_n$-projective} if there exists a $T_n$-projective subfunctor $Q$ of $L_n$ such that $M=Q(V)$. (\textbf{Note.} Here we require that $Q(V)$ is strictly equal to $M$ rather than isomorphic to $M$.)
\begin{prop}\label{proposition3.12}
Assume that the ground field $\ensuremath{{\mathbf{k}}}$ has infinite elements. Let $V$ be any $\ensuremath{{\mathbf{k}}}$-module and let $M$ be a $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-submodule of $L_n(V)$. Then $M$ is functorial $T_n$-projective if and only if $M$ satisfies the following two conditions:
\begin{enumerate}
\item[1)] There exists a $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-linear map $r\colon V^{\otimes n}\to M$ such that $r|_M$ is the identity.
\item[2)] The inclusion $M\hookrightarrow L_n(V)$ admits the following lifting
\begin{diagram}
& &V^{\otimes n}\\
&\ruDashto &\dOnto>{\bar\beta_n}\\
M&\rInto& L_n(V).\\
\end{diagram}
as modules over $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$.
\end{enumerate}
\end{prop}
\begin{proof}
Suppose that $M$ is functorial $T_n$-projective. Let $Q$ be a subfunctor of $L_n$ with $Q(V)=M$. The inclusion
$$
Q\rInto L_n\rInto T_n
$$
admits a natural retraction because $\gamma_n(Q)$ is injective. By evaluating at $V$, condition 1 is satisfied. Since $\gamma_n(Q)$ is projective, there a natural lifting $\tilde j\colon Q\to T_n$ such that $\beta_n\circ \tilde j$ is the inclusion of $Q$ in $L_n$ and so condition 2 is satisfied by evaluating at $V$.
Conversely suppose that $M$ satisfies conditions 1 and 2. Let $j\colon M \hookrightarrow L_n(V)$ and $ L_n(V)\hookrightarrow V^{\otimes n}$ be the inclusions. Let $\tilde j\colon M\to V^{\otimes n}$ be a $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-map such that $\beta_n\circ \tilde j=j$. By the Schur-Weyl duality, the map
$$
\ensuremath{{\mathbf{k}}}(\Sigma_n)=\mathrm{H o m}(T_n,T_n)\longrightarrow \mathrm{End}_{\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))}(V^{\otimes n})
$$
is an epimorphism. There exists a natural transformation $\alpha\colon T_n\to T_n$ such that $\alpha_V=\tilde j\circ r$.
Let $\theta=\beta_n\circ \alpha$. Consider the colimit of the sequence
$$
T_n\rTo^{\theta} T_n\rTo^{\theta} T_n\rTo\cdots,
$$
there exists $k>>0$ such that
$$
Q=\mathrm{Im}(\theta^k)\longrightarrow \operatorname{co l i m}_{\theta} T_n
$$
is an isomorphism because, by taking $\gamma_n(-)$ to the above sequence, the submodules $\mathrm{Im}(\gamma_n(\theta^t))$ of $\gamma_n(T_n)$ has the monotone decreasing in dimensions:
$$
\mathrm{dim\,} \mathrm{Im}(\gamma_n(\theta))\geq \mathrm{dim\,} \mathrm{Im}(\gamma_n(\theta^2))\geq \mathrm{dim\,} \mathrm{Im}(\gamma_n(\theta^3))\geq\cdots.
$$
Since $Q=\beta_n(\alpha\theta^{k-1}(T_n))$, $Q$ is a $T_n$-projective subfunctor of $L_n$. By evaluating at $V$, we check that $Q(V)=M$. Since
$$
\begin{array}{rcl}
\theta_V\circ\theta_V&=&\beta_n\circ\alpha_V\circ \beta_n\circ\alpha_V\\
&=&i\circ\bar\beta_n \circ \tilde j\circ r\circ i\circ\bar\beta_n \circ \tilde j\circ r\\
&=&i\circ j\circ r\circ i\circ j\circ r\\
&=& i\circ j\circ r\\
&=& \theta_V,\\
\end{array}
$$
the map $\theta_V$ is an idempotent. It follows that
$$
Q(V)=\mathrm{Im}(\theta_V)=i\circ j\circ r(V^{\otimes n})=M
$$
and hence the result.
\end{proof}
\section{Coalgebra Structure on Tensor Algebras}\label{section4}
In this section, the tensor algebra $T(V)$ admits the comultiplication
$$
\psi\colon T(V)\longrightarrow T(V)\otimes T(V)
$$
described in subsection~\ref{subsection2.1}.
\subsection{Changing Ground-rings}
Some results in the representation theory help us to change the
ground ring. Let $\mathbb{Z}_{(p)}$ be the $p$-local integers. By modular
representation theory of symmetric groups (see, for example
~\cite[Exercise 6.16, p.142]{CR}), any idempotent in
$(\mathbb{Z}/p)(\Sigma_n)$ lifts to an idempotent in $\mathbb{Z}_{(p)}(\Sigma_n)$.
It is well-known ~\cite{James} that any irreducible modules $M$ over
$\mathbb{Z}/p(\Sigma_n)$ is absolutely irreducible, that is, for any
extension field $\ensuremath{{\mathbf{k}}}$, $M\otimes\ensuremath{{\mathbf{k}}}$ is irreducible over
$\ensuremath{{\mathbf{k}}}(\Sigma_n)$. Thus there is a one-to-one correspondence between
idempotents in $\mathbb{Z}/p(\Sigma_n)$ and the idempotents in
$\ensuremath{{\mathbf{k}}}(\Sigma_n)$.
Let $R$ be any commutative ring with identity. Consider $T\colon
V\mapsto T(V)$ as the functor from projective $R$-modules to
coalgebras over $R$. Denote by $\mathrm{co a l g}^R(T,T)$ the set of self
natural coalgebra transformation of $T$. Let $\ensuremath{{\mathbf{k}}}$ be any field of
characteristic $p$. We have the canonical functions
$$
R\colon \mathrm{co a l g}^{\mathbb{Z}_{(p)}}(T,T)\rTo\mathrm{co a l g}^{\mathbb{Z}/p}(T,T)
$$
by modulo $p$ and
$$
K\colon \mathrm{co a l g}^{\mathbb{Z}/{p}}(T,T)\rTo \mathrm{co a l g}^{\ensuremath{{\mathbf{k}}}}(T,T)
$$
by tensoring with $\ensuremath{{\mathbf{k}}}$ over $\mathbb{Z}/p$. By~\cite[Corollary 6.9]{SW1},
there is a one-to-one correspondence between natural indecomposable
retract of $T$ over $\ensuremath{{\mathbf{k}}}$ and the indecomposable
$\ensuremath{{\mathbf{k}}}(\Sigma_n)$-projective submodule of $\mathrm{Lie}(n)$ for $n\geq
1$. Thus we have the following:
\begin{prop}\label{proposition4.1}
The functions $R$ and $K$ have the following properties:
\begin{enumerate}
\item The map $
R\colon \mathrm{co a l g}^{\mathbb{Z}_{(p)}}(T,T)\rTo\mathrm{co a l g}^{\mathbb{Z}/p}(T,T)$ induces a
one-to-one correspondence betweens idempotents. Thus every natural
coalgebra decompositions of $T$ over $\mathbb{Z}/p$ lifts to a natural
coalgebra decomposition over $\mathbb{Z}_{(p)}$.
\item The map $
K\colon \mathrm{co a l g}^{\mathbb{Z}_{p}}(T,T)\rTo\mathrm{co a l g}^{\ensuremath{{\mathbf{k}}}}(T,T)$ induces a
one-to-one correspondence betweens idempotents. Thus natural
coalgebra decompositions of $T$ only depends on the characteristic
of the ground field.\hfill $\Box$
\end{enumerate}
\end{prop}
By this proposition, we can freely change the ground fields $\ensuremath{{\mathbf{k}}}$
with the same characteristic and lift natural coalgebra
decompositions over $p$-local integers if it is necessary.
\subsection{Block Decompositions}
From now on in this section, the ground field $\ensuremath{{\mathbf{k}}}$ is algebraically closed with
$\mathrm{char}(\ensuremath{{\mathbf{k}}})=p$. For any coalgebra $C$, let $PC$ be the set of the primitives of $C$. If $C$ is a functor from modules to coalgebras, then $PC$ is a functor from modules to modules. Recall from Proposition~\ref{proposition2.3} that any subfunctor of $T$ is a graded subfunctor. Thus if $C$ is sub-quotient coalgebra functor of $T^C$, then $C$ is graded and so we have the homogenous functors $C_n$ and $P_nC=PC\cap C_n$ for each $n$. For the case $C=T$, $PT(V)=L^{\mathrm{res}}(V)$ is the free restricted Lie algebra generated by $V$ and $P_nT=L^{\mathrm{res}}_n$ for each $n$.
For natural transformations $f,g\colon T\to T$, the \textit{convolution product $f\ast g$} is defined by the composite
$$
T(V)\rTo^{\psi} T(V)\otimes T(V)\rTo^{f\otimes g} T(V)\otimes T(V)\rTo^{\mathrm{multi.}} T(V).
$$
If $f$ and $g$ are natural coalgebra transformation, clearly $f\ast g$ is also a natural coalgebra transformation. For any element $\zeta\in\ensuremath{{\mathbf{k}}}$, define $\lambda_\zeta\colon T(V)\to T(V)$ by setting
\begin{equation}\label{equation4.1}
\lambda_{\zeta}(a_1\cdots a_n)=\zeta^n a_1\cdots a_n
\end{equation}
for $a_1,\ldots,a_n\in V$. In other words, $\lambda_{\zeta}\colon T(V)\to T(V)$ is the (unique) Hopf map such that $\lambda_{\zeta}(a)=\zeta a$ for $a\in V$. Let $\chi\colon T(V)\to T(V)$ be the conjugation of the Hopf algebra $T(V)$, namely $\chi$ is the anti-homomorphism such that $\chi(a)=-a$ for $a\in V$. More precisely,
\begin{equation}\label{equation4.2}
\chi(a_1\cdots a_n)=(-1)^{n}a_na_{n-1}\cdots a_1
\end{equation}
for $a_1,\ldots,a_n\in V$. For any element $\zeta\in\ensuremath{{\mathbf{k}}}$, we have the natural coalgebra transformation
\begin{equation}\label{equation4.3}
\theta_{\zeta}=\lambda_\zeta\ast\chi\colon T(V)\to T(V).
\end{equation}
If $\alpha\in P_nT(V)$, then
\begin{equation}\label{equation4.4}
\theta_{\zeta}(\alpha)=(\zeta^n-1)\alpha
\end{equation}
by the definition of convolution product because $\psi(\alpha)=\alpha\otimes 1+1\otimes\alpha$. For general monomials in $T_n(V)$, it is straightforward to have the formula
\begin{equation}\label{equation4.5}
\theta_{\zeta}(a_1\cdots a_n)=\sum
_{
\begin{array}{c}
\sigma(1)<\cdots<\sigma(k)\\
\sigma(k+1)<\cdots<\sigma(n)\\
\sigma\in\Sigma_n\\
0\leq k\leq n\\
\end{array}}
\left(\zeta^k+(-1)^{n-k}\right) a_{\sigma(1)}\cdots a_{\sigma(k)}a_{\sigma(k+1)}\cdots a_{\sigma(n)}.
\end{equation}
The maps $\theta_\zeta$ are useful for obtaining natural coalgebra decompositions of $T(V)$.
\begin{thm}\label{theorem4.2}
Let the ground ring $\ensuremath{{\mathbf{k}}}$ be a field of characteristic $p$. Then there exists a natural coalgebra decomposition
$$
T(V)\cong C(V)\otimes D(V)
$$
for any module $V$ with the property that
$$
PC_n=\left\{
\begin{array}{lcl}
0&\textrm{ if }& n \textrm{ is not a power of } p,\\
P_nT &\textrm{ if } & n=p^r \textrm{ for some } r.\\
\end{array}\right.
$$
\end{thm}
\textbf{Note.} From the decomposition, we have $P_nD=0$ if $n$ is a power of $p$ and $P_nD=P_nT$ if $n$ is not a power of $p$. The theorem gives a decomposition that one can put all primitives of tensor length of powers of $p$ in a one coalgebra factor and put the rest primitives in another coalgebra factor.
\begin{proof}
Let $\{m_1<m_2<m_3<\cdots\}$ be the set of all positive integers prime
to $p$ excluding $1$ and let $\zeta_{m_i}$ be a primitive $m_i$th
root of $1$. We are going to construct by induction a sequence of sub coalgebra functor $C(k)$ of $T$, with the inclusion denoted by $j_k\colon C(k) \hookrightarrow T$, and a sequence of quotient coalgebra functor $q_k\colon T\to E(k)$ with the following properties:
\begin{enumerate}
\item[1)] $C(k+1)$ is a subfunctor $C(k)$ for each $k\geq 0$.
\item[2)] There exists a coalgebra natural transformation $q'_k\colon E(k)\to E(k+1)$ such that $q_{k+1}=q'_k\circ q_k$ for each $k\geq 0$.
\item[3)] The composite $q_k\circ j_k\colon C(k)\to E(k)$ is a natural isomorphism.
\item[4)] The primitives $P_nC(k)=0$ if $n$ is divisible by one of $m_1,m_2,\ldots,m_k$.
\item[5)] The primitive $P_nC(k)=P_nT$ if $n$ is not divisible by any of $m_1,m_2,\ldots,m_k$.
\end{enumerate}
Let $C(0)=E(0)=T$ and let $i_0=q_0=\mathrm{id}$. The construction of $C(1)$ and $E(1)$ is as follows. Let $E(1)=\operatorname{co l i m}_{\theta_{\zeta_{m_1}}}T$ be the colimit of the sequence of coalgebra natural transformation
$$
T\rTo^{\theta_{\zeta_{m_1}}} T\rTo^{\theta_{\zeta_{m_1}}} T\rTo\cdots.
$$
Let $q_1\colon T\to E(1)$ be the map to its colimit. By~\cite[Theorem 4.5]{SW1}, there exists a sub coalgebra functor $C(1)$ of $T$, with the inclusion denoted by $j_1\colon C(1)\to T$, such that $q_1\circ i_1$ is a natural isomorphism. From Equation~\ref{equation4.4},
$$
\theta_{\zeta_{m_1}}\colon P_nT\to P_nT
$$
is zero if $m_1|n$ and an isomorphism if $m_1\nmid n$. Thus $P_nE(1)=\operatorname{co l i m}_{\theta_{\zeta_{m_1}}}P_nT=0$ if $m_1|n$ and
$$
q_1\colon P_nT\longrightarrow P_nE(1)
$$
is an isomorphism if $m_1 \nmid n$. Since $C(1)\cong E(1)$, conditions (4) and (5) holds. Now suppose that we have construct $C(j)$ and $E(j)$ satisfying conditions (1)-(5) for $j\leq k$. Let $f\colon T\to T$ be the composite
$$
T\rOnto^{q_k}E_n\rTo^{(q_k\circ j_k)^{-1}}_{\cong} C(k)\rInto^{j_k}T \rTo^{\theta_{\zeta_{m_{k+1}}}}T
$$
and let $E(k+1)=\operatorname{co l i m}_{f}T$. Let $q_{k+1}\colon T\to E(k+1)$ be the canonical map to its colimit. Notice that
$$
q_{k+1}\circ f=q_{k+1}\colon T\longrightarrow E(k+1).
$$
Let $q'_k=q_{k+1}\circ j_k\circ (q_k\circ j_k)^{-1}$. Then $q_{k+1}=q'_k\circ q_k$ and so condition (2) satisfies. Since $f$ factors through the subfunctor $C(k)$, there exists a subfunctor $C(k+1)$ of $C(k)$, with the inclusion into $T$ denoted by $j_{k+1}$, such that $q_{k+1}\circ j_{k+1}$ is a natural isomorphism. Hence we have conditions (1) and (3). Let $\alpha\in P_nT(V)$. Then
$$
f(\alpha)=\theta_{\zeta_{m_{k+1}}}((j_k\circ(q_k\circ j_k)^{-1}\circ q_k)(\alpha))=(\zeta_{m_{k+1}}^n-1)((j_k\circ(q_k\circ j_k)^{-1}\circ q_k)(\alpha).
$$
Thus $f(\alpha)=0$ if $n$ is divisible by one of $m_1,\ldots,m_{k+1}$ and
$$
f\colon P_nT\longrightarrow P_nT
$$
is an isomorphism if $n$ is not divisible by any of $m_1,\ldots,m_{k+1}$. It follows that $P_nE(k+1)=0$ if $n$ is divisible by one of $m_1,\ldots,m_{k+1}$ and
$$
q_{k+1}\colon P_nT\longrightarrow P_nE(k+1)
$$
is an isomorphism if $n$ is not divisible by any of $m_1,\ldots,m_{k+1}$. Since $C(k+1)\cong E(k+1)$, we have conditions (4) and (5). The induction is finished.
Now let
$$
C=\bigcap_{k=0}^\infty C(k)
$$
be the intersection of the subfunctors $C(k)$ of $T$ and let $E(\infty)$ be the colimit of the sequence
$$
T\rOnto^{q_1}E(1)\rOnto^{q'_2}E(2)\rOnto^{q'_3}E(3)\rOnto\cdots.
$$
From condition (3), each $C(k)$ is coalgebra retract of $T$ and so each $C(k)$ is a functor from modules to coassociative and cocommutative quasi-Hopf algebras with the multiplication on $C(k)$ given by
$$
C(k)\otimes C(k)\rInto T\otimes T\rTo^{\mathrm{multi}} T\rOnto C(k),
$$
where we use the notation of quasi-Hopf algebra given in~\cite{MM}. By conditions (1)-(3), $C(k+1)$ is a coalgebra retract of $C(k)$ and so there is a natural coalgebra decomposition
$$
C(k)\cong C(k+1)\otimes C'(k)
$$
by~\cite[Lemma 5.3]{SW1}. From conditions (4) and (5), $P_nC(k+1)=P_nC(k)$ for $n<m_{k+1}$ and so $P_nC'(k)=0$ for $n<m_{k+1}$. It follows that
$$
C'(k)_n=0
$$
for $0<n<m_{k+1}$. Thus
$$
C(k+1)_n=C(k)_n
$$
for $n<m_{k+1}$ and from conditions (1)-(3),
$$
q_k\colon E(k)_n\longrightarrow E(k+1)_n
$$
is an isomorphism for $n<m_{k+1}$. Notice that the integers $m_k\to\infty$ as $k\to\infty$. Let $n$ be a fixed positive integer. For the integers $k$ with $m_k>n$, we have $C_n=C(k)_n$ and
$$
E(k)_n\rTo^{q'_k}_{\cong} E(k+1)_n\rTo^{q'_{k+1}}_{\cong} E(k+2)_n\rTo.
$$
It follows that the composite
$$
C_n=C(k)_n \rInto^{j_k} T_n\rOnto^{q_k} E(k)_n\rTo E(\infty)_n
$$
is an isomorphism. Thus the composite
$$
C\rInto T\rTo E(\infty)
$$
is an isomorphism and so $C$ is a coalgebra retract of $T$. This gives a natural coalgebra decomposition
$$
T\cong C\otimes D
$$
for some coalgebra retract $D$ of $T$. From conditions (4) and (5), we have $P_nC=0$ if $n$ is not a power of $p$ and $PC_{p^r}=PT_{p^r}$ for $r\geq 0$. The proof is finished.
\end{proof}
\begin{thm}[Block Decomposition Theorem]\label{theorem4.3}
Let $\ensuremath{{\mathbf{k}}}$ be a field of characteristic $p$.
Let $\{m_i\}_{i\geq0}$ be the set of all positive integers prime to $p$
with the order that $m_0=1<m_1<m_2<\cdots$. Then there exist natural coalgebra retracts $C^{m_i}$ of $T$ with a natural coalgebra decomposition
$$
T(V)\cong\bigotimes_{i=0}^\infty C^{m_i}(V)
$$
such that
$$
P_nC^{m_i}=\left\{
\begin{array}{ccc}
P_nT &\textrm{ if } & n=m_ip^r \textrm{ for some } r\geq0,\\
0&&\textrm{ otherwise }.\\
\end{array}\right.
$$
\end{thm}
\begin{proof}
We are going to show by induction that there exists natural coalgebra retracts $C^{m_i}$ of $T$, for $0\leq i\leq k$, with a natural coalgebra decomposition
\begin{equation}\label{equation4.6}
T(V)\cong \left( \bigotimes_{i=0}^k C^{m_i}(V)\right)\otimes D^k(V)
\end{equation}
for some natural coalgebra retract $D^k$ of $T$ such that
$$
P_nC^{m_i}=\left\{
\begin{array}{ccc}
P_nT &\textrm{ if } & n=m_ip^r \textrm{ for some } r\geq0,\\
0&&\textrm{ otherwise }.\\
\end{array}\right.
$$
for $0\leq i\leq k$. The statement holds for $k=0$ by Theorem~\ref{theorem4.2}, where $C^{m_0}$ is the natural coalgebra retract $C$ of $T$ given in Theorem~\ref{theorem4.2}. Suppose that the statement holds for $k$. Following the lines in the proof of Theorem~\ref{theorem4.2} with using $\{\theta_{\zeta_{m_i}}\}$ for $i\geq k+2$, there is a natural coalgebra decomposition
$$
D^k(V)\cong C^{m_{k+1}}(V)\otimes D^{k+1}(V).
$$
In brief, the first construction of $E^{m_{k+1}}(1)=\operatorname{co l i m}_gT$ is the colimit of the map $g$ given by the composite
$$
T\rOnto D^k \rInto T\rTo^{\theta_{\zeta_{m_{k+2}}}}T
$$
and the inductive construction follows the lines in the proof of Theorem~\ref{theorem4.2} with pre-composing $T\twoheadrightarrow D^k\hookrightarrow T$. This gives a monotone decreasing sequence of natural coalgebra retracts $C^{m_{k+1}}(i)$ of $D^k$ for $i=1,2,\ldots$ and the resulting natural coalgebra retract $C^{m_{k+1}}=\bigcap\limits_{i=1}^\infty C^{m_{k+1}}(i)$ of $D^k$ has the property in primitives that $P_nC^{m_{k+1}}=0$ if $n$ is divisible by one of $m_{k+2},m_{k+3},\ldots$ and $$P_nC^{m_{k+1}}=P_nD^k$$ if $n$ is not divisible by any of $m_{i}$ with $i\geq k+2$. Together with the fact that $P_nD=0$ if $n=m_ip^r$ for some $0\leq i\leq k$ and $r\geq0$ and $P_nD=P_nT$ otherwise, we have $P_nC^{m_{k+1}}=P_nT$ if $n=m_{k+1}p^r$ for some $r\geq 0$ and $P_nC^{m_{k+1}}=0$ otherwise. The induction is finished.
Now decomposition~\ref{equation4.6} induces a commutative diagram
\begin{diagram}
T\cong \left(\bigotimes_{i=0}^k C^{m_i}\right)\otimes D^k\cong \left(\bigotimes_{i=0}^{k+1} C^{m_i}\right)\otimes D^{k+1} &\rOnto^{q_{k+1}}_{\mathrm{proj.}}& \bigotimes_{i=0}^{k+1} C^{m_i}\\
&\rdOnto^{q_k}&\dOnto>{\mathrm{proj.}}\\
&& \bigotimes_{i=0}^k C^{m_i}\\
\end{diagram}
that induces a coalgebra natural transformation
$$
T\rTo^q \bigotimes_{i=0}^\infty C^{m_i}
$$
which is an isomorphism because, for each $n$, $D^k_n=0$ for sufficiently large $k>>0$. We finish the proof.
\end{proof}
\section{Proof of Theorem~\ref{theorem1.1}}\label{section5}
In this section, the ground ring is a field $\ensuremath{{\mathbf{k}}}$ of characteristic $p$.
\begin{lem}\label{lemma5.1}
Let $Q$ be a $T_n$-projective sub functor of $L^{\mathrm{res}}_n$. Then $Q$ is a subfunctor of $L_n$ and the sub
Hopf algebra $T(Q(V))$ of $T(V)$ generated by $Q(V)$ is a natural
coalgebra retract of $T(V)$.
\end{lem}
\begin{proof}
It is easy to see that $\gamma_n(L^{\mathrm{res}}_n)=\gamma_n(L_n)=\mathrm{Lie}(n)$. By Proposition~\ref{proposition3.1}, the image of the natural transformation
$$
\Phi^{L^{\mathrm{res}}_n}_V\colon \gamma_n(L^{\mathrm{res}}_n)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\longrightarrow L^{\mathrm{res}}_n(V)
$$
is $L_n(V)$. Since $Q$ is $T_n$-projective,
$$
\Phi^Q_V\colon \gamma_n(Q)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\longrightarrow Q(V)
$$
is an isomorphism by Proposition~\ref{proposition2.10} (1). From the commutative diagram
\begin{diagram}
\gamma_n(Q)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}&\rTo& \gamma_n(L^{\mathrm{res}}_n)\otimes_{\ensuremath{{\mathbf{k}}}(\Sigma_n)}V^{\otimes n}\\
\cong\dTo>{\Phi^Q_V}&&\dTo>{\Phi^{L^{\mathrm{res}}_n}_V}\\
Q(V)&\rInto& L^{\mathrm{res}}_n(V),\\
\end{diagram}
we have $Q\subseteq L_n$. By Proposition~\ref{proposition2.10}(4), there is a natural linear transformation
$$
r_V\colon T_n(V)=V^{\otimes
n}\rTo Q(V)
$$
with $r_V|_{Q(V)}=\mathrm{id}_{Q(V)}$. Let
$$
H_n\colon T(V)\rTo T(V^{\otimes n})
$$
be the algebraic James-Hopf map induced by the geometric James-Hopf
map by taking homology. Then there is a commutative diagram
\begin{diagram}
T(Q(V))&\rInto& T(L_n(V))&\rInto& T(V)\\
&\rdEq&&\rdTo>{T(j_V)}&\dTo>{H_n}\\
&&T(Q(V))&\lTo^{T(r_V)} &T(V^{\otimes n}),\\
\end{diagram}
where the maps in the top row are the inclusions of sub Hopf
algebras, $j_V$ is the canonical inclusion and the right triangle
commutes by the geometric theorem in~\cite[Theorem 1.1]{Wu1}. Thus
the sub Hopf algebra $T(Q(V))$ of $T(V)$ admits a natural coalgebra
retraction and hence the result.
\end{proof}
A natural sub Hopf algebra $B(V)$ of $T(V)$ is called
\textit{coalgebra-split} if the inclusion $B(V)\to T(V)$ admits a
natural coalgebra retraction. For a Hopf algebra $A$, denote by $QA$
the set of indecomposable elements of $A$. Let $IA$ be the
augmentation ideal of $A$. If $B(V)$ is a natural sub Hopf algebra
of $T(V)$, then there is a natural epimorphism $IB(V)\to QB(V)$. Let
$Q_nB(V)$ be the quotient of $B_n(V)=IB(V)\cap T_n(V)$ in $QB(V)$.
\begin{thm}\label{theorem5.2}
Let $B(V)$ be a natural sub Hopf algebra of $T(V)$. Then the
following statements are equivalent to each other:
\begin{enumerate}
\item[1)] $B(V)$ is a natural coalgebra-split sub Hopf algebra of $T(V)$.
\item[2)] There is a natural linear transformation $r\colon T(V)\to
B(V)$ such that $r|_{B(V)}$ is the identity.
\item[3)] Each $Q_nB$ is naturally equivalent to a $T_n$-projective sub functor of $L_n$.
\item[4)] Each $Q_nB$ is a $T_n$-projective functor.
\end{enumerate}
\end{thm}
\begin{proof}
$(1)\Longrightarrow (2)$ and $(3)\Longrightarrow (4)$ are obvious. By~\cite[Theorem 8.6]{SW1},
$(2)\Longrightarrow (1)$. Thus $(1)\Longleftrightarrow (2)$. From
the proof of~\cite[Theorem 8.8]{SW1}, $(2)\Longrightarrow (3)$.
$(4)\Longrightarrow (2)$. Since $B(V)$ is a sub Hopf algebra of
primitively generated Hopf algebra $T(V)$, $B(V)$ is primitively
generated and so
$$
r_n\colon P_nB(V)=B(V)\cap L^{\mathrm{res}}_n(V)\rTo Q_nB(V)
$$
is a natural epimorphism, where $L^{\mathrm{res}}(V)=PT(V)$ is the free
restricted Lie algebra generated by $V$. Since $Q_nB$ is $T_n$-projective, the map
$r_n$ admits a natural cross-section $s_n\colon Q_nB(V)\rInto
P_nB(V)$ by Proposition~\ref{proposition2.10} (4).
Now we show that the inclusion $B(V)\to T(V)$ admits a natural
linear retraction. By identifying $Q_nB(V)$ with $s_n(Q_nB(V))$, we
have
$$
B(V)=T\left(\bigoplus_{k=1}^\infty Q_kB(V)\right)\subseteq T(V).
$$
Since each $Q_kB$ is a retract of the functor $T_k$,
$$
Q_{i_1}B\otimes\cdots\otimes Q_{i_t}B
$$
is a retract of $T_{i_1+i_2+\cdots+i_t}$ for any sequence
$(i_1,\ldots,i_t)$.
Note that $\{Q_iB(V) \ | \ i\geq 1 \}$ are algebraically
independent. Thus the summation
$$
\sum_{i_1+i_2+\cdots+i_t=q} Q_{i_1}B(V)\otimes
Q_{i_2}B(V)\otimes\cdots \otimes Q_{i_t}B(V)\subseteq T_q(V)= V^{\otimes q}.
$$
is a direct sum. From the fact that
$$\bigoplus_{i_1+i_2+\cdots+i_t=q} Q_{i_1}B\otimes
Q_{i_2}B\otimes\cdots \otimes Q_{i_t}B$$
is $T_n$-projective, there is
natural linear retraction
$$
V^{\otimes q}\rTo \bigoplus_{i_1+i_2+\cdots+i_t=q}
Q_{i_1}B(V)\otimes Q_{i_2}B(V)\otimes\cdots \otimes Q_{i_t}B(V)
$$
for any $q\geq 1$. Hence the inclusion $B(V)\to T(V)$ admits a
natural linear retraction.
\end{proof}
\begin{proof}[Proof of Theorem~\ref{theorem1.1}]
Let $B(V)$ be the sub Hopf algebra of $T(V)$
generated by
$$
L_{m_ip^r}(V)\quad \textrm{ for } i\in I,\ r\geq 0.
$$
Let $\{n_j\}_{j\geq 1}=\{m_ip^r\ | \ i\geq 1,\ r\geq0\}$ with $$n_1=m_1<n_2<\cdots.$$ Namely we rewrite the integers $m_ip^r$ in order. Let $B[k](V)$ be the sub Hopf algebra of $V$ generated by $L_{n_j}(V)$ for $1\leq j\leq k$. By Theorem~\ref{theorem5.2}, it suffices to show that $Q_nB$ is $T_n$-projective for $n\geq 1$. Let $n$ be a fixed positive integer. Choose $k$ such that $n_k\geq n$. Then the inclusion $B[k]\hookrightarrow B$ induces an isomorphism
$$
Q_nB[k]\cong Q_nB
$$
because $B[k]$ and $B$ has the same set of generators in tensor length $\leq n$. Thus it suffices to show the following statement:
\begin{enumerate}
\item[] \textit{For each $k\geq 1$, $B[k]$ is coalgebra-split.}
\end{enumerate}
The proof of this statement is given by induction on $k$. The statement holds for $k=1$ by Lemma~\ref{lemma5.1} because $L_{m_1}$ is $T_{m_1}$-projective by~\cite[Corollary 6.7]{SW1} from the assumption that $m_i$ is prime to~$p$. Suppose that $B[k-1]$ is coalgebra-split. Thus there is a coalgebra natural transformation $r\colon T\to B[k-1]$ such that $r|_{B[k-1]}$ is the identity map. Let $n_k=m_ip^r$ for some $i$ and $r$. Let $C^{m_i}$ be the natural coalgebra retract in Theorem~\ref{theorem4.3} with a natural coalgebra retraction $r_C\colon T\to C^{m_i}$. Define $f\colon T\to T$ to be the composite
\begin{equation}\label{equation5.1}
T\rTo^{r_C} C^{m_i}\rInto T\rTo^r B[k-1]\rInto T.
\end{equation}
Let $\tilde E=\operatorname{co l i m}_f T$ be the colimit with the canonical map
$$
q\colon T\longrightarrow \tilde E.
$$
As in lines of the proof of Theorem~\ref{theorem4.2}, there exists a coalgebra subfunctor $\tilde C$ of $C^{m_i}$ such that
$$
q|_{\tilde C}\colon \tilde C\longrightarrow \tilde E
$$
is an isomorphism by~\cite[Theorem 4.5]{SW1} with a natural coalgebra decomposition
\begin{equation}\label{equation5.2}
C^{m_i}\cong \tilde C\otimes \tilde D.
\end{equation}
According to~\cite[Lemma 5.3]{SW1}, the subfunctor $\tilde D$ of $C^{m_i}$ can be chosen as the cotensor product
$\ensuremath{{\mathbf{k}}}\Box_{\tilde E}C^{m_i}$ under the coalgebra map
$$
q|_{C^{m_i}}\colon C^{m_i}\longrightarrow \tilde E
$$
and so there is a left exact sequence
\begin{equation}\label{equation5.3}
P_n\tilde D\rInto P_nC^{m_i} \rTo^{P_nq|_{C^{m_i}}} \tilde E.
\end{equation}
for any $n$.
By restricting the map $f$ as the composite in equation~(\ref{equation5.1}) to the primitives, we have the map
$$
P_nf\colon P_nT\rTo^{P_nr_C} P_nC^{m_i}\rInto P_nT\rTo^{P_nr} P_nB[k-1]\rInto P_nT.
$$
If $n\not=m_ip^t$ for $t\geq 0$, then $P_nf=0$ because $P_nC^{m_i}=0$. If $n=m_ip^t$ for some $t\geq 0$ with $n<n_k$, then $P_nf$ is the identity map because $P_nC^{m_i}=P_nT$ and $P_nB[k-1]=P_nT=L^{\mathrm{res}}_n$ as the sub Hopf algebra $B[k-1]$ contains $L_{m_ip^s}$ for $s\geq0$. Thus
$$
P_n\tilde C=P_nC^{m_i}
$$
for $n<n_k$. From decomposition~(\ref{equation5.2}), we have
\begin{equation}\label{equation5.4}
P_nC^{m_i}=P_n\tilde C\oplus P_n\tilde D
\end{equation}
for all $n$ and so $P_n\tilde D=0$ for $n<n_k$. It follows that $\tilde D_n=0$ for $0<n<n_k$ and
\begin{equation}\label{equation5.5}
\tilde D_{n_k}=P_{n_k}\tilde D.
\end{equation}
Now consider the case $P_nf$ for $n=n_k=m_ip^r$. Since $P_nC^{m_i}=P_nT$, $P_nr_C=\mathrm{id}$ and so $P_nf\circ P_nf=P_nf$ with
$$
P_nf(\alpha)=P_nr(\alpha)
$$
for $\alpha\in P_nT$. Thus the composite
$$
\mathrm{Im}(P_nf)= P_nB[k-1]\rTo^{P_nq|_{B[k-1]}} P_n\tilde E=\operatorname{co l i m}_{P_nf}P_nT
$$
is an isomorphism. From exact sequence~(\ref{equation5.3}), we have
$$
P_n\tilde D=\mathrm{K er}(P_nr \colon P_nC^{m_i}=P_nT\to P_nB[k-1]).
$$
Let $j\colon P_nB[k-1]\hookrightarrow P_nC^{m_i}=P_nT$ be the inclusion. From the commutative diagram
\begin{diagram}
& & P_n\tilde D=\mathrm{K er}(P_nr)& &\\
& &\dInto&\rdTo^{\cong}&\\
P_n\tilde C&\rInto^i& P_nC^{m_i}&\rOnto&\mathrm{C o k er}(i)\\
&\rdTo^{\cong}&\dOnto>{P_nr}&&\\
&&P_nB[k-1],&&\\
\end{diagram}
the summation
$P_nB[k-1]+ P_n\tilde D $
in $P_nT=P_nC^{m_i}$ is a direct sum with the decomposition
$$
P_nC^{m_i}=P_nT=P_nB[k-1]\oplus P_n\tilde D.
$$
From definition of $B[k]$, $PB[k](V)$ is the restricted sub Lie algebra of $L^{\mathrm{res}}(V)=PT(V)$ generated by $L_{n_i}(V)$ for $1\leq i\leq k$. It follows that $P_nB[k]=L_n^{\mathrm{res}}=P_nT=P_nC^{m_i}$ and
$$
Q_nB[k]\cong P_nB[k]/P_nB[k-1]=P_nC^{m_i}/P_nB[k-1]\cong P_n\tilde D.
$$
From decomposition~(\ref{equation5.2}), $\tilde D_n$ is a natural summand of $C^{min}_n$. Since $C^{min}$ is a coalgebra retract of $T$, $C^{{\mathrm{min}}}_n$ is a natural summand of $T_n$. Thus $\tilde D_n$ is $T_n$-projective. By identity~(\ref{equation5.5}), $P_n\tilde D$ is $T_n$-projective. Thus $Q_nB[k]$ is $T_n$-projective. By Theorem~\ref{theorem5.2}, $B[k]$ is coalgebra-split. The induction is finished and hence the result.
\end{proof}
By inspecting the proof, we have the following slightly stronger statement.
\begin{thm}\label{theorem5.3}
Let $\mathcal{M}=\{m_i\}_{i\in I}$ be finite or infinite set of positive integers prime to $p$ with each $m_i>1$. Let $f\colon I\to \{0,1,2,\ldots\}\cup\{\infty\}$ be a function. Then the sub Hopf algebra $B^{\mathcal{M},f}(V)$ of $T(V)$
generated by
$$
L_{m_ip^r}(V)\quad \textrm{ for } i\in I,\ 0\leq r< f(i)
$$
is natural coalgebra-split. \hfill $\Box$
\end{thm}
\section{Decompositions of Lie Powers}\label{section6}
Let $m=kp^r$ with $k\not\equiv 0\mod{p}$ and $k>1$. According to~\cite[Theorem 10.7]{SW1}, the functor $L_{kp^r}$ admits the following functorial decomposition:
$$
L_{kp^r}=L'_{kp^r}\oplus L_p(L'_{kp^{r-1}})\oplus\cdots\oplus L_{p^r}(L'_k)
$$
for each $r\geq 0$ starting with $L'_k=L_k$, where each $L'_{kp^r}$ is a summand of $T_{kp^r}$ which is called $T_{kp^r}$-projective in our terminology. By evaluating on $V$, one gets a decomposition of the $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-module $L_{kp^r}(V)$ given ~\cite[Theorem 4.4]{Bryant-Schocker} using a different approach from representation theory, where $L'_{kp^r}(V)$ was denoted by $B_{kp^r}$ in~\cite{Bryant-Schocker}. From Theorem~\ref{theorem5.3}, we can obtain various new decompositions of $L_{kp^r}$ and so the decompositions of the $\ensuremath{{\mathbf{k}}}(\mathrm{G L}(V))$-module $L_{kp^r}(V)$ by evaluating on $V$.
Let $\mathcal{M}=\{m_i\}_{i\in I}$ be finite or infinite set of positive integers prime to $p$ with each $m_i>1$. Let $f\colon I\to \{0,1,2,\ldots\}\cup\{\infty\}$ be a function. Let $B^{\mathcal{M},f}(V)$ be the sub Hopf algebra of $T(V)$
generated by
$$
L_{m_ip^r}(V)\quad \textrm{ for } i\in I,\ 0\leq r< f(i).
$$
According to Theorem~\ref{theorem5.3}, $B^{\mathcal{M},f}$ is coalgebra-split and so
$Q_nB^{\mathcal{M},f}$ is $T_n$-projective by Theorem~\ref{theorem5.2}. Since $B^{\mathcal{M},f}(V)$ is sub Hopf algebra of the primitively generated Hopf algebra $T(V)$, it is primitively generated by~\cite[Proposition 6.13]{MM} and so there is a natural epimorphism
$$
\phi_n\colon P_nB^{\mathcal{M},f}\twoheadrightarrow Q_nB^{\mathcal{M},f}.
$$
From Proposition~\ref{proposition2.10}(4), the map $\phi_n$ admits a natural cross-section because $Q_nB^{\mathcal{M},f}$ is $T_n$-projective. Thus there is a subfunctor $D^{\mathcal{M},f}_{n}$ of $P_nB^{\mathcal{M},f}$ such that
$$\phi_n|\colon D^{\mathcal{M},f}_n\to Q_nB^{\mathcal{M},f}$$ is a natural isomorphism. Since $D^{\mathcal{M},f}_n$ is $T_n$-projective,
$$
D^{\mathcal{M},f}_n\subseteq P_nB^{\mathcal{M},f}\cap L_n
$$
from the lines in the proof of Lemma~\ref{lemma5.1}. Thus $D^{\mathcal{M},f}_n$ is a $T_n$-projective subfunctor of $L_n$. From the fact that $D^{\mathcal{M},f}\cong Q_nB^{\mathcal{M},f}$ and $B^{\mathcal{M},f}$ is isomorphic to the tensor algebra generated by $Q_nB^{\mathcal{M},f}$ with $n\geq 1$, the inclusion
$$
\bigoplus_{n=1}^\infty D_n\rInto B^{\mathcal{M},f}
$$
induces a natural isomorphism
\begin{equation}\label{equation6.1}
T\left(\bigoplus_{n=1}^\infty D_n\right)\cong B^{\mathcal{M},f}.
\end{equation}
Since the algebra $B^{\mathcal{M},f}$ is generated by $L_{m_ip^r}(V)$ for $m_i\in\mathcal{M}$ and $0\leq r< f(i)$, we have
\begin{equation}\label{equation6.2}
D^{\mathcal{M},f}_n=0\textrm{ if } n\not=m_ip^r \textrm{ for some } m_i\in \mathcal{M}\textrm{ and some } 0\leq r< f(i).
\end{equation}
Let $\{m_ip^r \ | \ m_i\in\mathcal{M}, 0\leq r< f(i)\}=\{n_1,n_2,\ldots\}$ with $n_1<n_2<\cdots$ and let $\alpha$ be the cardinality of the set $\{m_ip^r \ | \ m_i\in\mathcal{M}, 0\leq r< f(i)\}$. Then decomposition~(\ref{equation6.1}) becomes
\begin{equation}\label{equation6.3}
T\left(\bigoplus_{i=1}^\alpha D_{n_i}\right)\cong B^{\mathcal{M},f}.
\end{equation}
and so a natural isomorphism
\begin{equation}\label{equation6.4}
PT\left(\bigoplus_{i=1}^\alpha D_{n_i}\right)=L^{\mathrm{res}}\left(\bigoplus_{i=1}^\alpha D_{n_i}\right)\cong PB^{\mathcal{M},f}=B^{\mathcal{M},f}\cap L^{\mathrm{res}}.
\end{equation}
According to Proposition~\ref{proposition4.1}, the sub Hopf algebra $B^{\mathcal{M},f}$ is also a natural coalgebra retract of $T$ if we change the ground ring $R$ to be $p$-local integers. Notice that $PT=L$ and $PB^{\mathcal{M},f}=B^{\mathcal{M},f}\cap L$ when $R=\mathbb{Z}_{(p)}$. By changing the ground ring back to $\mathbb{Z}/p$ and then extending it to $\ensuremath{{\mathbf{k}}}$, we have the following decomposition:
\begin{equation}\label{equation6.5}
L\left(\bigoplus_{i=1}^\alpha D_{n_i}\right)\cong B^{\mathcal{M},f}\cap L.
\end{equation}
For applying the Hilton-Milnor Theorem to determine $B^{\mathcal{M},f}\cap L_n$, let us recall the terminology of \textit{basic product} from~\cite[p.512]{Whitehead}. Let $x_1,\ldots,x_k$ be letters. A monomial means a formal product $w=x_{i_1}x_{i_2}\cdots x_{i_n}$ with $1\leq i_1,\ldots,i_t\leq k$, where the word length $n$ is called the \textit{weight} of $w$. We define the \textit{basic products} of weight $n$, by induction on $n$; and for each such product, a non-negative integer $r(w)$, called its \textit{rank}. These are to be linearly ordered, in such a way that $w_1<w_2$ if the weight of $w_1$ is less than the weight of $w_2$. The \textit{serial number $s(w)$} is the number of basic products $\leq w$ in term of this ordering. The basic products of weight $1$ are the letters $x_1,\ldots,x_k$ with the order $x_1<x_2<\cdots<x_k$. The rank $r(x_i)=0$ and the serial number $s(x_i)=i$. Suppose that the basic products of weight less than $n$ have been defined and linearly ordered in such a way that $w_1<w_2$ if the weight of $w_1$ is less than that of $w_2$; and suppose that the rank $r(w)$ of such a product has been defined. Then the basic products of weight $n$ are all monomials of the $w_1w_2$ of weight $n$, for which $w_1$ and $w_2$ are basic products, $w_2<w_1$ and $r(w_1)\leq s(w_2)$. Give these an arbitrary linear order, and define $r(w_1w_2)=s(w_2)$.
Let $\mathcal{W}_k$ be the set of all basic products on the letters $x_1,\ldots,x_k$ by forgetting the ordering. Then
$$
\mathcal{W}_k\subseteq\mathcal{W}_{k+1}
$$
for each $k$. Let
$$
\mathcal{W}_{\infty}=\bigcup_{k=1}^\infty \mathcal{W}_k.
$$
The elements in $\mathcal{W}$ are called basic products on the sequence of the letters $x_i$ for $i\geq 1$. For each basic product $w=x_{i_1}\cdots x_{i_t}\in \mathcal{W}_{\alpha}$, define
\begin{equation}\label{equation6.6}
w(D^{\mathcal{M},f})=D^{\mathcal{M},f}_{n_{i_1}}\otimes\cdots\otimes D^{\mathcal{M},f}_{n_{i_t}}
\end{equation}
with the \textit{tensor length with respect to $D^{\mathcal{M},f}$}
$$
d(w)=n_{i_1}+n_{i_2}+\cdots+n_{i_t}
$$
and the natural transformation
$$
\phi_w\colon w(D^{\mathcal{M},f})(V)=\longrightarrow T\left(\bigoplus_{i=1}^\alpha D_{n_i}(V)\right)\cong B^{\mathcal{M},f}(V)
$$
given by
$$
\phi_w(z_1\otimes z_2\otimes\cdots\otimes z_t)=[[z_1,z_2,\ldots,z_t]
$$
for $z_j\in D^{\mathcal{M},f}_{n_{i_j}}(V)$. Then the map $\phi_w$ extends uniquely to a natural transformation of Hopf algebras
$$
T\phi_w\colon T(w(D^{\mathcal{M},f}))\longrightarrow T\left(\bigoplus_{i=1}^\alpha D_{n_i}(V)\right)\cong B^{\mathcal{M},f}(V)
$$
by the universal property of tensor algebras. Now by taking the homology from the Hilton-Milnor Theorem~\cite[Theorem 6.7]{Whitehead}, we have the natural isomorphism of coalgebras
\begin{equation}\label{equation6.7}
\theta\colon \bigotimes_{w}T(w(D^{\mathcal{M},f}))\rTo^{\cong} T\left(\bigoplus_{i=1}^\alpha D_{n_i}\right)\cong B^{\mathcal{M},f},
\end{equation}
where $w$ runs over all basic products in $\mathcal{W}_{\alpha}$, the tensor product is linearly ordered and the natural transformation $\theta$ is given by the ordered product of $T\phi_w$ which is well-defined because the tensor length $d(w)$ tends to $\infty$ as the weight of $w$ tends to $\infty$. By restricting to Lie powers, we have the decomposition
\begin{equation}\label{equation6.8}
\theta|\colon \bigotimes_{w}L(w(D^{\mathcal{M},f}))\rTo^{\cong} L\left(\bigoplus_{i=1}^\alpha D_{n_i}\right)\cong B^{\mathcal{M},f}\cap L.
\end{equation}
By taking tensor length, we obtain the following decomposition theorem.
\begin{thm}\label{theorem6.1}
Let $\mathcal{M}=\{m_i\}_{i\in I}$ be finite or infinite set of positive integers prime to $p$ with each $m_i>1$ and let $f\colon I\to \{0,1,2,\ldots\}\cup\{\infty\}$ be a function. Then there exists $T_{m_ip^r}$-projective subfunctor $D^{\mathcal{M},f}_{m_ip^r}$ of $L_{m_ip^r}$ for each $m_i\in\mathcal{M}$ and $0\leq r< f(i)$ such that
$$
L_{m_ip^r}=\bigoplus_{d(w)|m_ip^r}L_{m_ip^r/d(w)}(w(D^{\mathcal{M},f}))
$$
for $m_i\in\mathcal{M}$ and $0\leq r< f(i)$, where $w$ runs over basic products with $d(w)|m_ip^r$.\hfill $\Box$.
\end{thm}
\noindent\textbf{Note.} Since each $D_{n_i}$ is $T_{n_i}$-projective, the tensor product $w(D^{\mathcal{M},f})$ is $T_{d(w)}$-projective. If $m_ip^r/d(w)$ is prime to $p$, then the Lie power $L_{m_ip^r/d(w)}(w(D^{\mathcal{M},f}))$ is $T_{m_ip^r}$-projective. Thus the non $T_{m_ip^r}$-projective summands of $L_{m_ip^r}$ occurs in the those factors $L_{m_ip^r/d(w)}(w(D^{\mathcal{M},f}))$ with $m_ip^r/d(w)\equiv0\mod{p}$.
Another remark is that the multiplicity of each factor $L_{m_ip^r/d(w)}(w(D^{\mathcal{M},f}))$ can be determined as follows: Let $w$ be a basic product involving the letters $x_{j_1},\ldots,x_{j_k}$ such that $x_{j_i}$ occurs $l_i$ times and $d(w)|m_ip^r$. According to~\cite[(6.4), p.514]{Whitehead}, the multiplicity of the factor $L_{m_ip^r/d(w)}(w(D^{\mathcal{M},f}))$ is given by the formula
$$
\frac{1}{l}\sum_{d|l_0}\mu(d)\frac{\left(\frac{l}{d}\right)!}{\left(\frac{l_1}{d}\right)!\cdots \left(\frac{l_k}{d}\right)!},
$$
where $\mu$ is the M\"obius function, $l_0$ is the greatest common divisor of $l_1,\ldots,l_k$ and $l=l_1+\cdots +l_k$.
\begin{example}{\rm
Let $\ensuremath{{\mathbf{k}}}$ be of characteristic $2$. Let $\mathcal{M}=\{m_1=3\}$ and let $f(1)=3$. Then we have natural coalgebra-split sub Hopf
algebra
$$
B^{\mathcal{M},f}(V)=\la L_3(V), L_6(V),L_{12}(V),\ra
$$
of $T(V)$ with $D^{\mathcal{M},f}_3=L_3$, $D^{\mathcal{M},f}_6\cong L'_6=L_6/L_2(L_3)$ and
$$
D^{\mathcal{M},f}_{12}\cong L_{12}/\left([L'_6,L'_6]\oplus [[L'_6,L_3],L_3]\oplus
L_4(L_3)\right)
$$
From Theorem~\ref{theorem6.1}, we have the decomposition
$$
\begin{array}{rcl}
L_{12}&=&D_{12}\oplus L(D_3\oplus D_6)\cap L_{12}\\
&=&D_{12}\oplus L_4(D_3)\oplus L_2(D_6)\oplus [[D_6, D_3],D_3]\\
&\cong& D_{12}\oplus L_4(L_3)\oplus L_2(L'_6)\oplus [[L'_6,L_3],L_3].\\
\end{array}
$$
By comparing with~\cite[Theorem 10.7]{SW1} or~\cite[Theorem 4.4]{Bryant-Schocker}, the $T_{12}$-projective summand
$$[[L'_6,L_3],L_3]\cong L'_6\otimes L_3\otimes L_3$$ can be recognized in our decomposition for $L_{12}$. \hfill $\Box$ }
\end{example}
Let $\mathcal{M}$ be the set of all positive integers $m_i$ with $m_i$ prime to $p$ and $m_i>1$ and let $f(i)=\infty$ for all $i$. Then the set
$$
\{m_ip^r\ | \ m_i\in \mathcal{M}\ r\geq 0\}=\mathbb{N}\smallsetminus \{1,p,p^2,p^3,\ldots\}.
$$
Let
$$
\bar D_n=D^{\mathcal{M},f}_n
$$
for $n$ not a power of $p$. As a special case of Theorem~\ref{theorem6.1}, we have the following:
\begin{cor}\label{corollary6.3}
There exists $T_n$-projective subfunctor $\bar D_n$ of $L_n$ for each $n$ not a power of $p$ such that
$$
L_{m}=\bigoplus_{d(w)|m}L_{m/d(w)}(w(\bar D))
$$
for any $m$ not a power of $p$, where $w$ runs over basic products with $d(w)|m$.\hfill $\Box$.
\end{cor}
|
Subsets and Splits