1: \documentclass[letterpaper]{epl2}
2:
3: % \newif\ifpdf
4: % \ifx\pdfoutput\undefined
5: % \pdffalse % we are not running PDFLaTeX
6: % \else
7: % \pdfoutput=1 % we are running PDFLaTeX
8: % \pdfcompresslevel=9
9: % \pdftrue
10: % \fi
11:
12: %\usepackage{axodraw}
13: \usepackage{amsmath}
14: \usepackage{graphicx}
15: \usepackage{bm}
16: \newcommand{\<}{\langle}
17: \renewcommand{\>}{\rangle}
18: \newcommand{\rr}{{\bf r}}
19: \newcommand{\uu}{{\bf u}}
20: \newcommand{\kk}{{\bf k}}
21: \newcommand{\RR}{{\bf R}}
22: \newcommand{\qq}{{\bf q}}
23: \newcommand{\QQ}{{\bf Q}}
24: \newcommand{\GG}{{\bf G}}
25: \newcommand{\dd}{{\vec{\bf \delta}}}
26: \newcommand{\cc}{{\hat c}}
27: \newcommand{\ah}{{\hat a}}
28: \newcommand{\bb}{{\hat b}}
29: \newcommand{\znpk}{z_{n'k}}
30: \newcommand{\znk}{z_{nk}}
31: \newcommand{\ccd}{{\hat c^\dagger}}
32: \newcommand{\ahd}{{\hat a^\dagger}}
33: \newcommand{\bbd}{{\hat b^\dagger}}
34: %\def\KeyWord#1{$\backslash$\IfColor{$\!\!$\textRed{#1}\textBlack}{#1}$\!\!$}
35: \newcommand{\la}{\langle}
36: \newcommand{\ra}{\rangle}
37: \newcommand{\up}{\uparrow}
38: \newcommand{\dn}{\downarrow}
39: \newcommand{\rar}{\rightarrow}
40: \def \bsig {\mbox{\boldmath$\sigma $}}
41: \def \btau {\mbox{\boldmath$\tau $}}
42:
43:
44: \title{The Berry phase and the pump flux in stochastic chemical
45: kinetics}
46: \shorttitle{The Berry phase in chemical kinetics}
47:
48: \author{N.A. Sinitsyn \inst{1,2} \and Ilya Nemenman \inst{2}}
49: \shortauthor{N.A. Sinitsyn and Ilya Nemenman}
50:
51: \institute{
52: \inst{1}Center for Nonlinear Studies and \\
53: \inst{2}Computer, Computational and Statistical Sciences Division,
54: Los Alamos National Laboratory, Los Alamos, NM 87545 USA
55: }
56: \pacs{82.20.-w}{Chemical kinetics and dynamics}
57: \pacs{03.65.Vf}{Phases: geometric; dynamic or topological}
58: \pacs{05.10.Gg}{Stochastic analysis methods (Fokker-Planck, Langevin, etc.)}
59: \pacs{05.40.Ca}{Noise}
60: \pacs{87.15.Ya}{Biological and medical physics: Fluctuations}
61:
62: \abstract{ We study a classical two-state stochastic system in a sea
63: of substrates and products (absorbing states), which can be
64: interpreted as a single Michaelis-Menten catalyzing enzyme or as a
65: channel on a cell surface. We introduce a novel general method and
66: use it to derive the expression for the full counting statistics of
67: transitions among the absorbing states. For the evolution of the
68: system under a periodic perturbation of the kinetic rates, the
69: latter contains a term with a purely geometrical (the Berry phase)
70: interpretation. This term gives rise to a pump current between the
71: absorbing states, which is due entirely to the stochastic nature of
72: the system. We calculate the first two cumulants of this current,
73: and we argue that it is observable experimentally. }
74:
75:
76: \begin{document}
77:
78: \maketitle
79:
80:
81: %\section{Introduction}
82:
83: Single molecule experiments \cite{english-etal-06} have led to a
84: resurgence of work on the stochastic version of the classical
85: Michaelis-Menten (MM) enzymatic mechanism that describes the
86: enzyme-driven catalytic conversion of a substrate into a product
87: \cite{MM}. A lot of progress (analytical and numerical) has been made
88: under various assumptions and for various internal enzyme structures
89: \cite{qian-elson-02,rao-arkin-03,qian-xie-06,gopich-szabo-06,xue-etal-06}. Still,
90: many questions linger. Specifically, linear signal transduction
91: properties of biochemical networks, such as frequency dependent gains,
92: are of a high interest
93: \cite{detwiler-etal-00,samoilov-etal-02,samoilov-etal-05}. These are
94: studied by probing responses to periodic perturbations in the input
95: signals (in our case, kinetic rates) due to fluctuations in chemical
96: concentrations, in temperature, or due to other signals. However,
97: such approaches usually disregard a phenomena known from the theory of
98: stochastic ratchets \cite{reimann-02}, especially in the context of
99: biological transport
100: \cite{julicher-etal-02,parrondo-02,qian-98,astumian-bio98}. Namely, a
101: system's symmetry may break, and, under an influence of a periodic or
102: random zero-mean perturbation, the system may respond with a finite
103: flux in a preferred direction. This, indeed, happens for the MM
104: reaction \cite{astumian-03}, and we study the phenomenon in detail in
105: this work.
106:
107:
108: The ratchet or pump effect manifests itself during periodic driving of
109: simple classical systems, such as a channel in a cell membrane, which
110: is formally equivalent to the MM enzyme
111: \cite{westerhoff-86,chen-87,astumian-87,tsong-88,astumian-pra89,astumian-jchph89,robertson-90,robertson-91,markin-91,astumian-03,tsong-chang-02}. The
112: prevalence of the transport terminology over the chemical one in the
113: literature is grounded in experiments, where driving is achieved by
114: application of periodic electric field or nonequilibrium noise that
115: modulate barrier heights for different channel conformations. The
116: resulting cross-membrane flux has contributions that have no analog in
117: stationary conditions.
118:
119: In a related system, a quantum pump
120: \cite{brouwer-98,makhlin-mirlin-01,moskalets-buttiker-02}, pump fluxes
121: admit a purely geometrical Berry phase interpretation
122: \cite{pump_berry}. A similar interpretation has been introduced for
123: some special classical cases \cite{shi,hannay-85}. However, the Berry
124: phase has not yet been derived for classical stochastic Markov chains
125: that model chemical kinetics, even though recent developments strongly
126: hint at the possibility
127: \cite{hill-book,qian1-98,qian-00,astumian-03,ao-04,kwon-05}. For
128: example, such systems admit introduction of a vector potential for the
129: fluxes to characterize circular motion \cite{qian1-98}, allow
130: reformulation of the Langevin dynamics in the form that strongly
131: resembles wave packet equations in quantum mecanics with a nontrivial
132: Berry curvature of Bloch bands \cite{ao-04,kwon-05,niu}, and result in
133: pump currents that depend on contour integrals over the the evolution
134: of the system's parameters \cite{astumian-03}. Although currents in
135: most analyzed models were produced by the lack of the detailed balance
136: rather than by external time-dependent perturbations, a dual
137: description in terms of an external noise induced ratchet effect is
138: usually possible. Conversely, time-dependent perturbations can break
139: the detailed balance and induce the catalytic cycle \cite{chen-87}.
140:
141: In the present work we demonstrate that a theory based on the Berry
142: phase can be constructed for a purely classical adiabatically slowly
143: driven stochastic dynamics. The theory leads to an elegant
144: interpretation of prior results, and it makes new predictions even for
145: the minimal model of the MM reaction. Additionally, the theory fills
146: in an important gap in the current state of the field by providing a
147: simple, yet general recipe for calculation of mean particle fluxes and
148: their fluctuations.
149:
150: In a standard minimal model, a single MM enzyme catalytic reaction
151: consists of an enzyme-substrate complex formation and its decay into
152: the enzyme and the product. Both reactions may be reversible
153: \cite{qian-elson-02}. We assume no multiple internal enzyme states
154: (see \cite{english-etal-06,qian-xie-06} for the complementing
155: discussion). The reaction can be viewed as a single bin (the enzyme)
156: that can contain either zero or one substrate particles in it. The
157: particle can escape from the bin by jumping into one of the two
158: absorbing states, either returning to the substrate -- the Left, or
159: converting into the product -- the Right, cf.~Fig.~\ref{system}. In
160: turn, if the bin is empty, either of the absorbing states can emit a
161: new particle into it. All transition times are exponentially
162: distributed with the rates defined as in Fig.~\ref{system}. We
163: investigate the mean particle current $J$ from $L$ into $R$ on time
164: scales much larger than the typical transition times between any two
165: states. Our main goal is to understand the effect of periodic driving,
166: i.\ e., periodic rate changes.
167:
168:
169: We emphasize again that this system is equivalent to a cross-membrane
170: transport problem, where the L/R states correspond to compartments on
171: different sides of the membrane, and the bin is a membrane channel
172: complex. Additionally, this system is relevant to transport through a
173: quantum dot in the Coulomb blockade regime \cite{nazarovFCS}.
174:
175: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
176:
177: \begin{figure}[t]
178: \centerline{\includegraphics[width=6 cm]{system}}
179: \caption{\label{system}Transition rates into and out of the absorbing
180: states $L$ (substrate) and $R$ (product). Note that the rates $k_1$
181: and $k_{-2}$ denote total rates (rather than per mole of
182: substrate/product), and they can be modulated, for example, by
183: changing concentrations of the substrate and the product,
184: respectively.}
185: \end{figure}
186:
187: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
188:
189:
190:
191: Let $P_{e}$ and $P_{f}$ be probabilities of an empty/filled bin
192: (unbound/bound enzyme). Neglecting particle discreteness and assuming
193: that all rates change slowly, so that $P_{e/f}$ are always in
194: equilibrium, one derives for the instantaneous $L\to R$ current
195: \begin{equation}
196: j_{\rm cl} (t)=\frac{\kappa_+(t)-\kappa_-(t)}{K(t)},\,\; K\equiv \sum_{m}
197: k_m,\,\; \kappa_{\pm}=k_{\pm1}k_{\pm2},
198: \label{jcl}
199: \end{equation}
200: and the summation is over $m=-2,-1,1,2$. This is the classical (MM)
201: current. Notice that, even if $\<k_1k_2\>=\<k_{-1}k_{-2}\>$ (where
202: $\<\dots\>$ denotes time averaging), $J_{\rm cl}\equiv\<j_{\rm cl}\>\neq0$
203: due to nonlinearities of the MM reaction.
204:
205: How is this result affected by the particle discreteness? The master
206: equation for the system's evolution is:
207: \begin{equation}
208: \frac{d}{dt} \left[ \begin{array}{l}
209: P_{e}\\
210: P_{f}
211: \end{array} \right] = - \left[
212: \begin{array}{ll}
213: \,\,\,\,k_1+k_{-2} & -k_{-1}-k_2\\
214: -k_1-k_{-2} &\,\,\,\, k_{-1}+k_2
215: \end{array} \right]
216: \left[ \begin{array}{l}
217: P_{e}\\
218: P_{f}
219: \end{array} \right],
220: \label{ev1}
221: \end{equation}
222: and we assume the adiabatic approximation, i.\ e., the rates $k_m$
223: oscillate with the frequency $\omega\ll \min_m k_m$. Since
224: $P_e=1-P_f$, (\ref{ev1}) is equivalent to a formally solvable first
225: order linear differential equation. However, expressing the solution
226: in a simple form requires approximations (e.g., small fluctuations)
227: even for a harmonic dynamics of the parameters
228: \cite{astumian-03}. Additionally, the solution provides little
229: information about the fluxes beyond their mean values, and it is not
230: easy to generalize to the case of more complicated, multi-state
231: stochastic systems. Thus here we pursue a different analysis
232: technique.
233:
234: The formal solution of (\ref{ev1}) is
235: \begin{equation}
236: {\bf p}(t)=\hat{T}\left(e^{-\int_{t_0}^{t}\hat{H}(t)dt}\right){\bf p}(t_0).
237: \label{ev12}
238: \end{equation}
239: where $\hat{T}$ stands for the time-ordering operator, ${\bf
240: p}(t)\equiv [ P_{e}(t),P_{f}(t) ]^T$, and
241: \begin{equation}
242: \hat{H}= \left[
243: \begin{array}{ll}
244: \,\,\,\, k_1 +k_{-2} & -k_{-1}-k_2 \\
245: -k_1-k_{-2} & \,\,\,\, k_{-1}+k_2
246: \end{array} \right].
247: \label{h0}
248: \end{equation}
249:
250: It is useful to separate the parts of $\hat{H}$ that are directly
251: responsible for the transition into or out of the $R$-state, namely
252: $\hat{H}\equiv\hat{H}_0 - \hat{V}_{+}-\hat{V}_{-}$, where
253: \begin{eqnarray}
254: &&\hat{H}_0= \left[
255: \begin{array}{cc}
256: k_1+k_{-2} & -k_{-1} \\
257: -k_1 & k_{-1}+k_2
258: \end{array} \right],
259: \label{h00}\\
260: &&\hat{V}_+= \left[
261: \begin{array}{cc}
262: 0 & k_2 \\
263: 0 & 0
264: \end{array} \right], \,\,\,\,\,\,\,\,
265: \hat{V}_-= \left[
266: \begin{array}{cc}
267: 0 & 0 \\
268: k_{-2} & 0
269: \end{array} \right].
270: \label{VV}
271: \end{eqnarray}
272: With this, we can write a formal expression for the probability $P_n$ to
273: have $n$ net transitions from the bin into $R$ during time $t$. This
274: is similar to \cite{gopich-szabo-06}, but we also allow $k_{-2}\neq0$,
275: counting transitions from $R$ into the bin with the negative sign. For
276: example,
277: \begin{align}
278: P_{n=1}&= {\bf 1}^+\left[ \int_{t_0}^{ t} dt_1 U_0(t,t_1)
279: \hat {V}_{+}(t_1) U_0(t_1,t_0) \right.+ \nonumber\\
280: +&\int_{t_0}^{t} dt_3 U_0(t,t_3)
281: \hat {V}_{+}(t_3)
282: \int_{t_0}^{t3} dt_2 U_0(t_3,t_2)
283: \hat {V}_{-}(t_2)\times \nonumber\\
284: \times&\left.\int_{t_0}^{t2} dt_1 U_0(t_2,t_1)\hat {V}_{+}(t_1)
285: U_0(t_1,t_0) +\cdots\right] {\bf p}(t_0). \label{p1}
286: \end{align}
287: Here ${\bf 1}$ is the vector with all unit entries and
288: \begin{equation}
289: U_0(t,t') = \hat{T}\left(e^{-\int_{t'}^t\hat{H}_0(t)dt}\right).
290: \label{u0}
291: \end{equation}
292: We introduce the probability generating function
293: (pgf) for the number of transitions from the bin into R,
294: \begin{equation}
295: Z(\chi)=e^{S(\chi)}=
296: \sum_{s=-\infty}^{\infty} P_{n=s}e^{is\chi}.
297: \label{pgf1}
298: \end{equation}
299: $\chi$ is called {\em the counting field}, $S(\chi)$ is {\em the full
300: counting statistics}, and its derivatives give cumulants (or
301: connected correlation functions) of $P_n$, e.g. $\<n\>=-i \partial
302: S(\chi)/\partial \chi |_{\chi=0} $, $\<\delta^2
303: n\>\equiv\<n^2\>-\<n\>^2 = (-i)^2 \partial^2 S(\chi)/\partial \chi^2
304: |_{\chi=0}$, etc.
305:
306: We derive the full counting statistics similarly to
307: \cite{gopich-szabo-06,nazarovFCS}, but we extend the method by
308: allowing for an explicit time-dependence of $\hat{H}$. First notice
309: that (\ref{pgf1}) with the expansion of probabilities as in (\ref{p1})
310: is equivalent to an expansion of an evolution operator with a
311: $\chi$-dependent Hamiltonian:
312: \begin{equation}
313: Z= {\bf 1}^+\hat{T}\left(e^{-\int_{t_0}^{t}\hat{H}(\chi,t) dt}\right) {\bf p}(t_0),
314: \label{pdf2}
315: \end{equation}
316: where
317: \begin{align}
318: \hat{H}(\chi,t)&=\hat{H}_0(t)-\hat{V}_+(t)e^{i\chi}-
319: \hat{V}_{-}(t)e^{-i\chi}\nonumber\\
320: =&\left[
321: \begin{array}{cc}
322: k_1 + k_{-2} & -k_{-1}-k_2 e^{i\chi} \\
323: -k_1-k_{-2}e^{-i\chi} & k_{-1}+k_2
324: \end{array} \right].
325: \label{hchi}
326: \end{align}
327: We define the two instantaneous eigenstates of $\hat {H}(\chi,t)$ as
328: $|u_0\rangle$ and $|u_1\rangle$. There are also the left eigenstates $
329: \langle u_0|$ and $\langle u_1|$ with the same eigenvalues; we
330: normalize them so that $ \langle u_{n}(t)|u_m (t)\rangle=\delta_{nm}$.
331:
332:
333:
334: Let's introduce an intermediate time scale $\delta t$, $ 1/\omega \gg
335: \delta t \gg \max_m (1 /k_m)$. During this time, typically, many
336: transitions happen but $k_m$'s are approximately constant. Consider
337: time-points $t_j=t_0+j\delta t$, $j=1\dots N$, and $t_N\equiv t$. In
338: the adiabatic limit, one can approximately rewrite the expression for
339: the time ordered exponent in (\ref{pdf2}) as a product of evolution
340: operators over time intervals of the size $\delta t$ with parameters
341: set to constants at times $t_j$, i.\ e.,
342: \begin{equation}
343: Z\approx {\bf 1}^+ e^{-\hat{H}(\chi,t_N)\delta t}
344: e^{-\hat{H}(\chi,t_{N-1})\delta t} \dots e^{-\hat{H}(\chi,t_0)\delta t} {\bf p}(t_0).
345: \label{ZZ}
346: \end{equation}
347: Now we insert the resolution of the identity
348: \begin{equation}
349: \hat{1}= |u_0 (t_i)\rangle \langle u_0(t_i)|
350: + |u_1 (t_i)\rangle\langle u_1(t_i)|
351: \label{one}
352: \end{equation}
353: after every exponent at $t_i$. We define $|u_0(t_i) \rangle$ as the
354: eigenstate of $\hat{H}(\chi,t_i)$ corresponding to
355: the eigenvalue $\lambda_0$ with the smaller real
356: part. Since $\delta t\gg1/k_i$, $\vert \exp\left[-\lambda_0(\chi,t_i)\delta
357: t\right]\vert \gg \vert \exp\left[-\lambda_1(\chi,t_i)\delta t\right]\vert$, and
358: terms containing $|u_1(t)\rangle$ can be neglected. Moreover, after
359: long evolution, any information about the initial and the final states
360: is lost. Thus we rewrite the pgf as
361: \begin{equation}
362: Z\approx e^{-\lambda_{min}(\chi,t_0)\delta t} \prod_{i=1}^Ne^{-\lambda_0(\chi,t_i)\delta t}
363: \langle u_0(t_i)|u_0(t_{i-1})\rangle.
364: \label{ZZ1}
365: \end{equation}
366: Finally, we approximate $\langle u_0(t_{i})|u_0(t_{i-1})\rangle
367: \approx \exp[-\delta t \langle u_0(t_{i})
368: | \partial_t|u_0(t_{i-1})\rangle]$, which allows us to write
369: \begin{equation}
370: S(\chi) \approx S_{\rm geom}+S_{\rm cl}=-\frac{T}{T_0}\int_{0}^{T_0}dt\, [
371: \langle u_0|\partial_t|u_0\rangle+\lambda_0(\chi,t) ],
372: \label{fcs2}
373: \end{equation}
374: where $T_0=2\pi / \omega$ is the period of the rate oscillations and
375: $T\gg T_0$ is the total measurement time. The last term in
376: (\ref{fcs2}) would be the same even for a time-independent problem,
377: and it has been discussed before \cite{nazarovFCS}. Diagonalizing
378: $\hat{H}$, and denoting $e_{\pm\chi}\equiv e^{\pm i\chi}-1$, we get for this term
379: \begin{equation}
380: S_{\rm cl}=\frac{-T}{2T_0} \int_0^{T_0} dt \left[ K-
381: \sqrt{K^2+4(\kappa_+e_{+\chi}+\kappa_-e_{-\chi})} \right].
382: \label{Scl}
383: \end{equation}
384:
385: The first term in (\ref{fcs2}) is more interesting. Since it depends
386: only on the choice of the contour in the ${\bf k}$ space, but not on
387: the rate of motion along the contour (at least in the adiabatic
388: approximation), it has a geometric interpretation. We write
389: \begin{equation}
390: S_{\rm geom}=-\frac{T}{T_0}\oint_{{\bf c}} {\bf A} \cdot d{\bf
391: k},\;\,
392: A_m = \langle u_0({\bf k})|\partial_{k_m}|u_0({\bf k})\rangle,
393: \label{Sgeom}
394: \end{equation}
395: where the integral is over the contour in the 4-dimentional
396: parameter space (the ${\bf k}$-space) drawn
397: during one cycle.
398:
399: To estimate the error in our results, we note that (\ref{pdf2}) is
400: exact, and then assumptions followed. First, we neglected the initial
401: nonequilibrium relaxation on a time scale of $\sim k_m^{-1}$. Since
402: the $S(\chi)\sim T$, this introduces an error in $S$ of
403: $\sim1/(k_{m}T)$, which vanishes for long observation times. Second,
404: we projected the evolution only to the subspace of states with the
405: smallest real part of the eigenvalue. The resulting error is
406: exponentially suppressed in the adiabatic limit by
407: $\exp(-(\lambda_{1}-\lambda_0)/\omega)$. Third, there is the coarse
408: graining in (\ref{ZZ}). To the lowest order, it introduces errors in
409: $S$ in the form of commutators, such as $[H(\chi,t_i),H(\chi,t_j)]$, $|t_i-t_j|
410: < \delta t$. Since $\langle u_0(t+\delta
411: t)|[H(\chi,t_i),H(\chi,t_j)]|u_0(t)\rangle \sim O([\omega \delta t]^2)$ for $t
412: <t_i,t_j<t+\delta t$, this error is less significant than the
413: $O(\omega)$ contribution from the geometric term in $S$. Finally, the
414: error due to the approximation of $\<u_0(t_i)|u_0(t_{i-1})\>$ in
415: (\ref{ZZ1},~\ref{fcs2}) is of the same order.
416:
417: In a chemical system, the rates $k_m$ can be changed by many means,
418: e.g., by varying the system's temperature. However, the simplest
419: scenario is to couple the substrate and the product to particle baths
420: and to vary the corresponding chemical potentials for both
421: species. Since the rates $k_1$ and $k_{-2}$ are proportional to the
422: particle numbers, they will oscillate as well. Thus in what follows we
423: assume that $k_1$ and $k_{-2}$ are time dependent, while $k_2$ and
424: $k_{-1}$ are constants. Then, using Stokes theorem, we write
425: \begin{equation}
426: \oint {\bf A} \cdot d{\bf k} = \iint_{{\bf s_c}} dk_{1}dk_{-2} F_{k_1,k_{-2}},
427: \label{stokes}
428: \end{equation}
429: where the integration is over the surface ${\bf s_c}$ enclosed by the
430: contour ${\bf c}$, and
431: \begin{equation}
432: F_{k_1,k_{-2}}=\frac{\partial A_{{-2}}}{\partial k_{1}}-
433: \frac{\partial A_{1}}{\partial k_{-2}}.
434: \label{berry}
435: \end{equation}
436: We will call $F_{k_1,k_{-2}}$ the {\em Berry curvature} by analogy
437: with similar definitions in quantum mechanics. The advantage of
438: working with $F$ rather than with the {\em potentials} $A_m$ is that
439: the Berry curvature is gauge invariant, i.e., it does not depend on an
440: arbitrary ${\bf k}$-dependent normalization of $|u_0({\bf
441: k})\rangle$. It is a truly measurable quantity. In our case,
442: \begin{equation}
443: F_{k_1,k_{-2}}=\frac{e_{-\chi}(e^{i\chi}k_2+k_{-1})}{[4\kappa_+e_{+\chi}+
444: 4\kappa_-e_{-\chi}+K^2]^{3/2}}.
445: \label{berry2}
446: \end{equation}
447: Note that, with $\chi=0$, the Berry curvature is zero, so it is the
448: counting field $\chi$ that introduces a nontrivial topology in the
449: phase space of the eigenstates of $\hat{H}(\chi,t)$. More generally, a
450: normal Markovian evolution corresponds to $\chi=0$, where the
451: normalization of $P$ ensures that $F|_{\chi=0}=0$. This may be the
452: reason why nobody discussed the Berry phases in the context of Markov
453: chains.
454:
455: Knowing the full counting statistics, one can study both the particle
456: flux and its fluctuations. Importantly, from (\ref{fcs2},
457: \ref{Sgeom}), all cumulants will have the geometric and the classical
458: terms. In particular, for the mean $L\to R$ flux per unit time, we get
459: \begin{equation}
460: J=J_{\rm pump}+J_{\rm cl}=\iint_{{\bf s_c}} d^2k
461: \frac{k_2+k_{-1}}{T_0K^3}+
462: \int_0^{T_0} dt\, \frac{j_{\rm cl}(t)}{T_0},
463: \label{JJ}
464: \end{equation}
465: where $j_{\rm cl}$ is the classical current defined in (\ref{jcl}).
466: Generally, the ratio of geometric and classical terms is $\sim\omega
467: /k_m \ll 1$.
468:
469: For the large observation time $T$, the total expected $L\to R$
470: particle flux is $\<n(T)\>=JT$. Similarly, its fluctuations are given
471: by $\<\delta^2n(T)\>=J^{(2)}T$, where
472: \begin{align}
473: &J^{(2)}=-\frac{1}{T}\left.\frac{\partial^2 \left( S_{\rm geom} + S_{\rm cl}\right)}{\partial \chi^2}\right|_{\chi=0}=J_{\rm pump}^{(2)}+J_{\rm cl}^{(2)}
474: ,\label{noise}\\
475: &J^{(2)}_{\rm pump}=\iint_{{\bf s_c}} d^2k
476: \left[ \frac{k_2-k_{-1}}{T_0K^3}-
477: \frac{12(k_2+k_{-1})(\kappa_+-\kappa_-)}{T_0K^5} \right],
478: \label{pump_noise}\\
479: &J^{(2)}_{\rm cl}=\frac{1}{T_0}\int_0^{T_0}dt
480: \left[
481: \frac{\kappa_++\kappa_-}{K}-\frac{(\kappa_+-\kappa_-)^2}{K^3} \right].
482: \label{cl-noise}
483: \end{align}
484: The ratio of the pump and the classical noises is again $\sim \omega/
485: k_m$. Importantly, we see that, in general, expectations of the total,
486: the classical, and the pump currents are not equal to their variances.
487: Thus none of the currents in the MM problem is Poissonian. This is a
488: general property of complex, multi-step reactions, which is often
489: neglected in computational studies of biochemical reaction
490: networks. Also note that $J^{(2)}_{\rm pump}$ in (\ref{pump_noise}) is
491: not necessarily positive and can, in fact, {\em decrease} the total
492: noise. However, the total noise variance $J^{(2)}$ is strictly positive.
493:
494: While the pump current in this system has been analyzed,
495: cf.~\cite{astumian-03}, to our knowledge, an expression for $J$
496: parameterized by the rates, rather than by internal enzyme parameters
497: has not been available. Even more importantly, expressions for the
498: noise (either classical or pump) for variable kinetic rates have not
499: existed either.
500:
501: The smallness of the pump effect compared to the classical current
502: complicates its observation. However, several opportunities exist.
503: One is the dependence of $J_{\rm pump}$ on the frequency of the
504: perturbation, while $J_{\rm cl}\neq J_{\rm cl}(\omega)$. The second
505: possibility is to vary the rates along a contour with $J_{\rm
506: cl}=0$. However, even in this case the noise may still be dominated
507: by the classical contribution.
508:
509: To test our predictions, we choose one particular such contour with
510: $J_{\rm cl}=0$, $J_{\rm pump}\neq 0$ for a numerical analysis:
511: \begin{equation}
512: k_{1}=A+R\cos(\omega t),\; k_{-2} =A+R\sin(\omega t),\; k_{-1}=k_2=1.
513: \label{timeev}
514: \end{equation}
515: To estimate $J^{(2)}$ numerically, we implemented a Gillespie-like
516: scheme \cite{gillespie-77} that admits an explicit time dependence of
517: the rates. However, since $J_{\rm pump}\ll \sqrt{J^{(2)}_{\rm cl}}$,
518: such Monte-Carlo estimation of $J$ would take too long. Instead, we
519: numerically solved the master equation (\ref{ev1}) for $P_f(t)$ with a
520: time discretization $dt\ll \min_m (1/k_m)$. This gave a better
521: precision than an analytical result of \cite{astumian-03}, which
522: assumes $R\to0$. Knowing $P_f$,
523: $j(t)=k_2P_f-k_{-2}(1-P_f)$. Fig.~\ref{Jpump} shows an excellent
524: agreement of our theory with the numerical results for both $J$ and
525: $J^{(2)}$.
526:
527:
528:
529: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
530:
531: \begin{figure}[t]
532: \centerline{{\bf(a)}}
533: %\ifpdf
534: \includegraphics[width=8cm]{mean_data}
535: %\else
536: % \includegraphics[height=8cm, angle=270]{mean_data} \fi
537: \centerline{{\bf(b)}}
538: % \ifpdf
539: \includegraphics[width=8cm]{var_data}
540: %\else \includegraphics[height=8cm, angle=270]{var_data} \fi
541: \centering
542: \caption{Comparison of analytical predictions and numerical results
543: for a system defined by (\ref{timeev}) with $A=1.5$. Here the theory
544: gives: $J_{\rm cl}=0$, $J_{\rm pump}= \omega R^2[2(2-R^2+4A+2A^2)
545: ]^{-3/2}$, $J_{\rm cl}^{(2)}=1-(50-2R^2)(25-2R^2)^{-3/2}$, and
546: $J^{(2)}_{\rm pump}=0$. Points with error bars mark numerical
547: results, while lines correspond to theoretical predictions. {\bf
548: (a)} Mean $L\to R$ particle flux $J$ obtained by integrating
549: (\ref{ev1}) with $dt=0.001$ and averaged over $\sim 100$ rate
550: oscillations. Error bars are negligible, and so is the computational
551: time to generate the data. {\bf (b)} Mean $L\to R$ flux variance
552: $J^{(2)}$ for $\omega=0.03$ Hz. Each point is a result of simulating
553: $3\cdot 10^6$ elementary particle hops, and covers about 6000 rate
554: oscillations. Error bars denote one standard deviation, as estimated
555: from the posterior variance of $J^{(2)}$. These simulations took
556: less than a day on a modern PC using Octave. }
557: \label{Jpump}
558: \end{figure}
559:
560: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
561:
562: Intuitively, the flux cannot keep increasing linearly as $\omega$
563: grows and the adiabatic approximation fails. To understand the
564: behavior at large frequencies, we consider the case of $\max k_m \ll
565: \omega$. We look for the probabilities in the form
566: $P_{e/f}=\bar{P}_{e/f} +\delta P_{e/f}$, where $\bar{P}_{e/f}$ are
567: calculated for the time-averaged parameter values, namely, $\bar{P}_e
568: = (\<k_2\>+\<k_{-1}\>)/\<K\>$, $\bar{P}_f=1-\bar{P}_e$, and $\delta
569: P_{e/f}$ are small and oscillate fast. The latter can be found from
570: (\ref{ev1}). For example, when $k_1=\<k_1\> +\delta k_1 \cos(\omega
571: t)$, $k_{-2}=\<k_{-2}\>+\delta k_{-2} \sin(\omega t)$, and $k_2$,
572: $k_{-1}$ are constants, we find $\delta P_{e} = -(\delta k_1 \bar{P}_e
573: /\omega) \sin(\omega t) +(\delta k_{-2} \bar{P}_f/\omega )\cos(\omega
574: t)$. Then the current averaged over the oscillation period is
575: $J=J_{\rm cl}(\{\<k_m\>\})+J_{\rm pump}$, where $J_{\rm pump} =
576: \delta k_1 \delta k_{-2}\bar{P}_e/(2\omega)$ is the correction due to
577: the fast rate oscillations. For the parameters as in (\ref{timeev}),
578: we again get $J_{\rm cl}=0$, but $J_{\rm pump}\neq0$. However, the
579: dependency of $J_{\rm pump}$ on $\omega$ has changed from $\sim\omega$
580: in the adiabatic case to $\sim 1/\omega$ for fast oscillations. This
581: agrees with \cite{astumian-03}, which used a different approximation
582: and showed that there is a single maximum in the pump current at
583: $\omega \sim k_m$.
584:
585: The phenomenon of $J_{\rm pump}\neq 0$ can be explained in simple
586: terms. Since a particle spends a finite time bound to the enzyme, the
587: values of $k_1$ and $k_{-2}$ cannot influence the system during the
588: mean unbinding time following a binding event. If, during an
589: oscillation, the left binding rate is higher than the right one during
590: the upramp of the cycle, then $k_1$ ``shields'' growing values of
591: $k_{-2}$ from having an effect, while $k_{-2}$ shields decreasing
592: values of $k_1$ during the downramp. This leads to a phase-dependent
593: asymmetry that is the source of the pump flux. Larger frequencies lead
594: to more shielding, hence the linear $\omega$ dependence. However, for
595: $\omega\to\infty$, the information about the phase of the oscillations
596: is lost while a particle is bound, decreasing the asymmetry and the
597: flux.
598:
599:
600:
601: Is the pump flux observable experimentally in biochemical (rather than
602: channel transport) experiments? Exact zeroing of $J_{\rm cl}$, as in
603: the numerical example above, which would make $J_{\rm pump}$ the
604: leading effect, may be difficult to achieve. However, the classical
605: current is also small near the classical steady state,
606: $\<k\>_1k_2=k_{-1}\<k_{-2}\>$. In this case, if
607: $k_{1,-2}=\<k_{1,-2}\>+\delta k_{1,-2}\cos(\omega t-\phi_{1,-2})$, and
608: if the oscillations are small, $\delta k_{1,-2}/k_{1,-2}\ll 1$, then
609: we can disregard variation of the Berry curvature inside the contour, and
610: \begin{align}
611: J_{\rm pump}&\approx \omega \frac{k_2+k_{-1}}{2\<K\>^3 }\delta k_1 \delta
612: k_{-2}\sin\phi,\\
613: J_{\rm pump}^{(2)}&
614: \approx \omega \frac{k_2-k_{-1}}{2\<K\>^3 }\delta k_1 \delta
615: k_{-2}\sin\phi,
616: \end{align}
617: where $\phi=\phi_{-2}-\phi_1$. This should be compared to the
618: classical contributions in the same limit
619: \begin{align}
620: &J_{\rm cl} \approx \frac{k_2 \delta k_1^2-
621: k_{-1} \delta k_{-2}^2 + (k_2-k_{-1})
622: \delta k_1\delta k_{-2} \cos \phi}{2\<K\>^2},\\
623: &J_{\rm cl}^{(2)} \approx \frac{2\<k_1\>k_2}{\<K\>}.
624: \end{align}
625: We see that, while $J_{\rm pump}/J_{\rm cl} \sim \omega /\<K\>$, the
626: ratio of the variances is further suppressed, $J^{(2)}_{\rm
627: pump}/J_{\rm cl}^{(2)}\sim (\omega/\<K\>)\delta k_1\delta
628: k_{-2}/(k_1k_{-2})$. Thus overcoming the classical variance is the
629: biggest concern for a successful experiment.
630:
631:
632: Our model has a single substrate and a single product. Thus the enzyme
633: we are discussing is an EC 5 enzyme. Properties of such enzymes vary
634: dramatically depending on a reaction, a biological species, and
635: mutations in protein sequences \cite{brenda}. A typical range of $k_2$
636: {\em in vitro} is $10^{-2}\dots10^4$ s$^{-1}$. Similarly, the
637: Michaelis constant, $K_M=(k_{-1}+k_2)[S]/k_1$ in our notation, varies
638: between $0.01$ and $10$ mM (here $[S]$ is the substrate
639: concentration). For our analysis, we take $k_2\sim 10$ s$^{-1}$, and
640: $K_M=1$ mM. While little is known about $k_{-1}$ and $k_2$ separately,
641: it's reasonable to assume $k_{-1}\sim k_2$. Similarly, we take
642: $k_{-2}\sim k_{1}[P]/[S]$ since both rates are often dominated by the
643: mean particle-enzyme collision time \cite{qian-elson-02} (for example,
644: for triose phosphate isomerase reaction, $k_{-2}/k_1\approx 2 [P]/[S]$
645: \cite{knowles-albery-77}). Many enzymes with similar parameters have
646: been characterized in the BRENDA database \cite{brenda}. Then, with
647: $[S]\sim [P]\sim 1$ mM and with $\delta k_{1,-2}/k_{1,-2}\sim10$\%, we get
648: $J_{\rm pump}\sim J^{(2)}_{\rm pump}\sim 10^{-2}/T_0$, $J_{\rm cl}\sim
649: 10^{-1}$, and $J_{\rm cl}^{(2)}\sim 10$ particles per
650: second. Oscillation periods of $\sim 1$ s are attainable (and still
651: satisfy $\omega\ll k_m$), which puts the flux ratio at $\sim
652: 10^{-1}$. Different dependence on $\phi$ can further improve
653: detectability of $J_{\rm pump}$ compared to $J_{\rm cl}$. At these
654: parameters, the pump flux becomes equal to the total flux standard
655: deviation for an experiment lasting a few days. Alternatively, working
656: with, say, $\sim10^6$ enzymes, $J_{\rm pump}$ and $J^{(2)}_{\rm pump}$
657: become observable in less than a minute. While real experiments will
658: certainly have additional complications, it is clear that our
659: predictions should be experimentally testable, at least in principle.
660:
661:
662:
663: In conclusion, we constructed the Berry phase theory of a purely
664: classical adiabatic stochastic pump in a simple Michaelis-Menten
665: enzymatic mechanism. Our approach allowed calculation of the particle
666: flux and its variance, including the classical and the pump effect
667: contributions. We believe that these predictions can be checked
668: experimentally, and it should be interesting to consider their
669: importance in the context of enzymatic signal processing.
670:
671: While we analyzed only one specific model system, the Berry phase
672: approach is general and can be employed for many processes that can be
673: reduced to slowly driven Markov dynamics. Examples range from charge
674: transport in quantum dots at strong decoherence, to particle transport
675: through cell membrane channels, and to various biochemical reaction
676: systems. For these problems, our method has several advantages
677: compared to already known techniques. First, it provides a formal
678: recipe for an analysis of an arbitrary Markov chain: one should
679: construct a Hamiltonian with counting fields, find its eigenvalue with
680: the lowest real part and the Berry curvature from the corresponding
681: eigenstate, and then the counting statistics is given by (\ref{fcs2},
682: \ref{Sgeom}). Second, the method provides a solution not only for
683: average fluxes, but also for the full counting statistics and thus for
684: all flux cumulants. This is important in the context of elimination of
685: fast degrees of freedom for construction of coarse-grained biochemical
686: reaction models (e.~g., MM or Hill phenomenological laws), where a
687: correct treatment of noise in the remaining degrees of freedom has
688: always been a point of contention. Third, the method allows to
689: transfer many results of the well-developed Berry phase theory for
690: dissipative quantum dynamics \cite{garrison-88,vertex-1,vertex-2} to
691: the field of chemical kinetics. In particular, techniques exist to
692: extend our work and to calculate flux cumulants to all orders in
693: $\omega/ k_{min}$ \cite{Garanin}.
694:
695:
696: \begin{acknowledgments}
697: We thank W.\ Hlavacek, F.\ Mu, M.\ Wall, C.\ Unkefer and F.\
698: Alexander for useful discussions and critical reading of the
699: manuscript. We are grateful to our two anonymous referees for
700: insightful comments, which substantially improved this letter. Our
701: work was funded in part by DOE under Contract No.\
702: DE-AC52-06NA25396. IN was further supported by NSF Grant No.\
703: ECS-0425850.
704: \end{acknowledgments}
705:
706: \begin{thebibliography}{0}
707:
708: \bibitem{english-etal-06}
709: \Name{English B.~P., et al.}
710: \REVIEW{Nat.\ Chem.\ Biol.}{2}{2006}{87}.
711:
712: \bibitem{MM}
713: \Name{Michaelis L. \and Menten M.~L.},
714: \REVIEW{Biochem. Z.}{49}{1913}{333}.
715:
716:
717:
718: \bibitem{qian-elson-02}
719: \Name{Qian H. \and Elson E.~L.},
720: \REVIEW{Biophys.\ Chem.}{101–102}{2002}{565}.
721:
722: \bibitem{rao-arkin-03}
723: \Name{Rao C.~V. \and Arkin A.~P.}
724: \REVIEW{J. Chem. Phys.}{118}{2003}{4999}.
725:
726: \bibitem{qian-xie-06}
727: \Name{Qian H. \and Xie X.~S.},
728: \REVIEW{Phys. Rev. E.}{74}{2006}{010902(R)}.
729:
730: \bibitem{gopich-szabo-06}
731: \Name{Gopich I.V. \and Szabo A.}
732: \REVIEW{J. Chem. Phys.}{124}{2006}{154712}.
733:
734: \bibitem{xue-etal-06}
735: \Name{Xue X., Liu F. \and Ou-Yang Z.-c.}
736: \REVIEW{Phys. Rev. E}{74}{2006}{030902(R)}.
737:
738: \bibitem{detwiler-etal-00}
739: \Name{Detwiler P., et al.}
740: \REVIEW{Biophys.\ J.}{79}{2000}{2801}.
741:
742: \bibitem{samoilov-etal-02}
743: \Name{Samoilov M., Arkin A.~P. \and Ross J.}
744: \REVIEW{J. Phys. Chem. A}{106}{2002}{10205}.
745:
746: \bibitem{samoilov-etal-05}
747: \Name{Samoilov M., Plyasunov S. \and Arkin A.~P.}
748: \REVIEW{Proc. Natl. Acad. Sci. (USA)}{102}{2005}{2310}.
749:
750: \bibitem{reimann-02}
751: \Name{Reimann P.}
752: \REVIEW{Phys. Rep.}{361}{2002}{57}.
753:
754: \bibitem{julicher-etal-02}
755: \Name{J\"{u}licher F., Ajdari A. \and Prost J.}
756: \REVIEW{Rev. Mod. Phys.}{69}{1997}{1269}.
757:
758: \bibitem{parrondo-02}
759: \Name{Parrondo J.~M.~R. \and de~Cisneros B.~J.}
760: \REVIEW{Appl. Phys. A}{75}{2002}{179}.
761:
762: \bibitem{qian-98}
763: \Name{Qian H.}
764: \REVIEW{Phys. Rev. Lett.}{81}{1998}{3063}.
765:
766: \bibitem{astumian-bio98}
767: \Name{Astumian A.~D. \and Derenyl I.}
768: \REVIEW{Eur. Biophys. J.}{27}{1998}{474}.
769:
770: \bibitem{astumian-03}
771: \Name{Astumian A.~D.}
772: \REVIEW{Phys. Rev. Lett.}{91}{2003}{118102}.
773:
774:
775: \bibitem{westerhoff-86}
776: \Name{Westerhoff H.~V., et al.}
777: \REVIEW{Proc. Natl. Acad. Sci. U.S.A.}{83}{1986}{4734}.
778:
779: \bibitem{chen-87}
780: \Name{Chen Y-D}
781: \REVIEW{Proc. Natl. Acad. Sci. U.S.A.}{84}{1987}{729}.
782:
783: \bibitem{astumian-87}
784: \Name{Astumian A.~D., et al.}
785: \REVIEW{Proc. Natl. Acad. Sci. U.S.A.}{84}{1987}{434}.
786:
787: \bibitem{tsong-88}
788: \Name{Tsong T.~Y. \and Astumian R.~D.}
789: \REVIEW{Ann. Rev. Physiol.}{50}{1988}{273}.
790:
791: \bibitem{astumian-pra89}
792: \Name{Astumian A.~D. \and Chock P.~B.}
793: \REVIEW{Phys. Rev. A.}{39}{1989}{6416}.
794:
795: \bibitem{astumian-jchph89}
796: \Name{Astumian A.~D. \and Robertson B.}
797: \REVIEW{J. Chem. Phys.}{91}{1989}{4891}.
798:
799: \bibitem{robertson-90}
800: \Name{Robertson B. \and Astumian A.~D.}
801: \REVIEW{Biophys. J.}{58}{1990}{969}.
802:
803: \bibitem{robertson-91}
804: \Name{Robertson B. \and Astumian A.~D.}
805: \REVIEW{J. Chem. Phys.}{94}{1991}{7414}.
806:
807: \bibitem{markin-91}
808: \Name{Markin V.~S. \and Astumian A.~D.}
809: \REVIEW{Biophys. J.}{59}{1991}{1317}.
810:
811: \bibitem{tsong-chang-02}
812: \Name{Tsong T.~Y. \and Chang C.~H.},
813: \REVIEW{AAAPS Bulletin}{13}{2003}{12}.
814:
815: \bibitem{brouwer-98}
816: \Name{Brouwer P.~W.}
817: \REVIEW{Phys. Rev. B}{58}{1998}{R10135}.
818:
819: \bibitem{makhlin-mirlin-01}
820: \Name{Makhlin Y. \and Mirlin A.~D.}
821: \REVIEW{Phys. Rev. Lett.}{87}{2001}{276803}.
822:
823:
824: \bibitem{moskalets-buttiker-02}
825: \Name{Moskalets M. \and B\"{u}ttiker M.}
826: \REVIEW{Phys. Rev. B}{66}{2002}{035306}.
827:
828: \bibitem{pump_berry}
829: \Name{Avron J.~E., et al.}
830: \REVIEW{Phys. Rev. B}{62}{2000}{R10618}.
831:
832: \bibitem{shi}
833: \Name{Shi Y. \and Niu Q.}.
834: \REVIEW{Europhys. Lett.}{59}{2002}{324}.
835:
836: \bibitem{hannay-85}
837: \Name{Hannay J.~H.}
838: \REVIEW{J. Phys. A.}{18}{1985}{221}.
839:
840: \bibitem{hill-book}
841: \Name{Hill T.~L.}
842: \Book{Free Energy Transduction and Biochemical Cycle Kinetics}
843: \Publ{Springer}
844: \Year{1995}.
845:
846: \bibitem{qian1-98}
847: \Name{Qian H.}
848: \REVIEW{Phys. Rev. Lett}{81}{1998}{3063}.
849:
850:
851: \bibitem{qian-00}
852: \Name{Qian H., Qian M.}
853: \REVIEW{Phys. Rev. Lett}{84}{2000}{2271}.
854:
855: \bibitem{ao-04}
856: \Name{P. Ao}
857: \REVIEW{J. Phys. A: Math. Gen.}{37}{2004}{L25}.
858:
859: \bibitem{kwon-05}
860: \Name{Kwon C., Ao P., Thouless D.~J.}
861: \REVIEW{Proc. Natl. Acad. Sci. U.S.A.}{102}{2005}{13029}.
862:
863: \bibitem{niu}
864: \Name{Niu Q., Sundaram G.}
865: \REVIEW{Phys. Rev. B}{59}{1999}{14915}.
866:
867: \bibitem{nazarovFCS}
868: \Name{Bagrets D.~A. \and Nazarov Y.~V.}
869: \REVIEW{Phys. Rev. B}{67}{2003}{085316}.
870:
871: \bibitem{gillespie-77}
872: \Name{Gillespie G.~T.}
873: \REVIEW{J. Phys. Chem}{81}{1977}{2340}.
874:
875: \bibitem{brenda}
876: \Name{Pharkya P., Nikolaev E.~V. \and Maranas C.~D.}
877: \REVIEW{Metab. Eng.}{5}{2003}{71}. Database release 2006.2.
878:
879: \bibitem{knowles-albery-77}
880: \Name{Knowles J.~R. \and Albery W.~J.}
881: \REVIEW{Accounts Chem. Res.}{10}{1977}{105}.
882:
883:
884: \bibitem{garrison-88}
885: \Name{Garrison J.~C. \and Wright E.~M.}
886: \REVIEW{J. Phys. A: Math. Gen.}{39}{1988}{13727}.
887:
888: \bibitem{vertex-1}
889: \Name{Ao P. \and Thouless D.~J.}
890: \REVIEW{Phys. Rev. Lett.}{76}{1993}{2158}.
891:
892: \bibitem{vertex-2}
893: \Name{Thouless D.~J., Ao P. \and Niu Q.}
894: \REVIEW{Phys. Rev. Lett.}{76}{1996}{3758}.
895:
896: \bibitem{Garanin}
897: \Name{Schilling R., Vogelsberger M. \and Garanin D.~A.}
898: \REVIEW{J. Phys. A: Math. Gen.}{39}{2006}{13727}.
899:
900: \end{thebibliography}
901:
902:
903: \end{document}
904:
905:
906:
907:
908:
909:
910:
911: