1:
2: %\documentclass[prb,twocolumn,floats,aps,showpacs]{revtex4}
3:
4: \documentclass[pre,twocolumn,showpacs]{revtex4}
5: \usepackage{epsfig}
6:
7:
8: \def\kx{\left( {p k_x \over 2} \right)}
9: \def\ky{\left( {p k_y \over 2} \right)}
10: \def\8{\infty}
11: \def\oh{\frac{1}{2}}
12: \def\ot{\frac{1}{3}}
13: \def\oq{\frac{1}{4}}
14: \def\tt{\frac{2}{3}}
15: \def\ft{\frac{4}{3}}
16: \def\tq{\frac{3}{4}}
17: \def\d{\partial}
18: \def\i{\imath\,}
19: \def\ih{\frac{\imath}{2}\,}
20: \def\undertext#1{\vtop{\hbox{#1}\kern 1pt \hrule}}
21: \def\ra{\rightarrow}
22: \def\lfa{\leftarrow}
23: \def\Ra{\Rightarrow}
24: \def\lra{\longrightarrow}
25: \def\ler{\leftrightarrow}
26: \def\lrb#1{\left(#1\right)}
27: \def\O#1{O\left(#1\right)}
28: \def\VEV#1{\left\langle\,#1\,\right\rangle}
29: \def\tr{\hbox{tr}\,}
30: \def\trb#1{\tr\lrb{#1}}
31: \def\dd#1{\frac{d}{d#1}}
32: \def\dbyd#1#2{\frac{d#1}{d#2}}
33: \def\pp#1{\frac{\partial}{\partial#1}}
34: \def\pbyp#1#2{\frac{\partial#1}{\partial#2}}
35: \def\ff#1{\frac{\delta}{\delta#1}}
36: \def\fbyf#1#2{\frac{\delta#1}{\delta#2}}
37: \def\pd#1{\partial_{#1}}
38:
39:
40:
41:
42:
43:
44: \def\br{\\ \nonumber & &}
45: %%\def\brr{\right. \\ \nonumber & &\left.}
46: %%\def\inv#1{\frac{1}{#1}}
47:
48: \def\be{\begin{equation}}
49: \def\ee{\end{equation}}
50: \def\bea{\begin{eqnarray} & &}
51: \def\eea{\end{eqnarray}}
52: \def\ct#1{\cite{#1}}
53: \def\rf#1{(\ref{#1})}
54: \def\EXP#1{\exp\left(#1\right)}
55: \def\TEXP#1{\hat{T}\exp\left(#1\right)}
56: \def\INT#1#2{\int_{#1}^{#2}}
57: \def\MAT{{\it Mathematica }}
58: \def\LHS{left-hand side }
59: \def\RHS{right-hand side }
60: \def\COM#1#2{\left\lbrack #1\,,\,#2\right\rbrack}
61: \def\AC#1#2{\left\lbrace #1\,,\,#2\right\rbrace}
62: %LOCAL
63: \def \PDF{probability distribution function }
64: \def\t{\tilde}
65: \def\ad{a^\dagger}
66: \def\cH{{\cal H}}
67: \def\Kinp{ \partial^2 \Pi\left(\phi_{i}-\phi_{i-1}\right)}
68: \def\E{{\cal E}}
69: \def\rfs#1{Eq.~(\ref{#1})}
70: \def\Urms {U_{\rm rms}}
71:
72:
73: \begin{document}
74:
75: \title{Burgers Equation Revisited}
76:
77: \author {V. Gurarie}
78: \affiliation{Department of Physics, Theoretical Physics, Oxford
79: University, 1 Keble Rd, Oxford OX1 1JP, United Kingdom}
80:
81: \date{14 March 2002}
82: %\date{\today}
83:
84:
85: \begin{abstract}
86: This paper studies the 1D pressureless turbulence (the Burgers
87: equation). It shows that reliable numerics in this problem is very
88: easy to produce if one properly discretizes the Burgers equation.
89: The numerics it presents confirms the $7/2$ power law proposed for
90: probability of observing large negative velocity gradients in this
91: problem. It also suggests that the entire probability function for
92: the velocity gradients could be universal, perhaps in some
93: approximate sense. In particular, the probability that the
94: velocity gradient is negative appears to be $p \approx 0.21 \pm
95: 0.01$ irrespective of the details of the random force. Finally, it
96: speculates that the theory initially proposed by Polyakov, with a
97: particular value of the ``anomaly" parameter, may indeed be exact,
98: at least as far as velocity gradients are concerned.
99: \end{abstract}
100:
101: \pacs{47.27.-i, 05.40.-a, 05.45.-a}
102:
103: \maketitle
104:
105: \section{Introduction}
106:
107: The problem of the randomly driven Burgers equation, or 1D
108: pressureless turbulence, attracted a lot of attention in the
109: literature in the last decade
110: \cite{Yakhot,Polyakov,BMP,GM,Sinai,Frisch,KG,Bec}. Indeed, it is a
111: tantalizing problem. On the one hand, it incorporates many
112: features one would expect from a real 3D turbulent fluid. On the
113: other hand, it is much simpler than the 3D turbulence of the
114: incompressible fluid, usually studied in the framework of the
115: Navier-Stokes equation. Its simplicity prompted some researchers
116: to suggest that it will become the ``Ising model'' of turbulence.
117: Yet many published papers later, we are still far from a complete
118: solution to this problem. In fact, many of the aspects of the
119: Burgers equation remain the subject of an ongoing debate. Several
120: competing methods exist in the literature, each with its own range
121: of applicability, which provide different, sometimes mutually
122: exclusive answers to various aspects of the 1D turbulence.
123:
124:
125: The situation could be improved if a way existed to check various
126: statements made about the Burgers equation. Such a check could be
127: provided by numerical simulations. But the numerical simulations
128: in turbulence are hard to conduct, they require large expenditure
129: of resources, and the answers they provide are often too ambiguous
130: to either confirm or disprove a particular theory. The situation
131: in Burgers turbulence seems to be no different. Despite several
132: influential papers on the numerical aspects of Burgers turbulence,
133: those studies failed to rule out completely or confirm any of the
134: competing theories currently on the market. The exception to this
135: seems to be the paper of J. Bec \cite{Bec}, where he confirmed the
136: asymptotics of a certain probability distribution function
137: (discussed below) predicted earlier in Ref. \cite{Sinai}.
138:
139:
140: In this paper, I would like to point out that there exists an
141: alternative way to study the Burgers equation numerically. The
142: method involves minimizing the Hopf-Cole functional instead of
143: trying to solve a complicated nonlinear differential equation.
144: That turned out to be extremely easy to do numerically. A ten line
145: program, which can be written in half an hour, produces, after
146: several minutes of computer time, clean graphs of probability
147: distribution functions. With the help of this program, I report
148: confirmation of some of the existing theories. I also hope that
149: the method, being so simple, can be used by other researchers to
150: do a quick check if their proposed theory indeed does not
151: contradict the numerical experiment.
152:
153: The rest of this paper is organized as follows. In section II I
154: formulate the problem of Burgers turbulence and summarize many of
155: the known facts about it. In section III I show how Burgers
156: equation must be discretized in order to facilitate doing
157: numerics. In section IV I discuss the statistics of the velocity
158: gradients of the Burgers equation using the methods of Lagrangian
159: trajectories. I also discuss the ``anomaly" hypothesis of Polyakov
160: and explain its meaning within the Lagrangian approach. In section
161: V I discuss the numerical simulations, show that they are
162: compatible with the asymptotic results of Refs.~\cite{Sinai,Bec}
163: and show that they are also compatible with the Polyakov
164: hypothesis if the ``anomaly" parameter $\beta$ is chosen to be
165: equal to $3/2$. Finally in the last section VI I also show that
166: the compatibility with the known results on the behavior of pinned
167: charge density waves \cite{AR,Fogler,GC,GC1} also forces the value
168: $\beta=3/2$. I speculate that since the problem of pinned charge
169: density waves and the asymptotics of Refs.~\cite{Sinai,Bec} relate
170: to different properties of Burgers turbulence, there is a chance
171: that the Polyakov's theory with $\beta=3/2$ is indeed exact, at
172: least as far as the calculation of probabilities of velocity
173: gradients is concerned, unless a remarkable and unlikely
174: coincidence is at play here.
175:
176: \section{Formulation of the Problem}
177: The Burgers turbulence problem can be formulated in just a few lines.
178: Take a pressureless fluid in
179: 1D, introduce its velocity field $u(x,t)$ where $x$ is space and $t$ is time and
180: write down the Navier-Stokes equation
181: \be
182: \label{burgers}
183: \partial_t u+u~\partial_x u - \nu~\partial^2_x u = f(x,t).
184: \ee Here $\nu$ is the viscosity of the fluid which is taken to be
185: very small. $f(x,t)$ is a force acting on the fluid, which is
186: taken to be random, white noise in time, and a smooth function in
187: space \be \VEV{f(x,t) f(x,t')}= \delta(t-t') F(x-x'), \ee where
188: $F(x)$ is a function which is positive at $x=0$ and smoothly goes
189: to zero when $x \gg L$, where $L$ sets the force length scale. As
190: far as boundary conditions are concerned, it is usually assumed
191: that $u$ is either periodic in space (in which case $F$ has to be
192: periodic as well), or the space is infinite and $u \rightarrow 0$
193: at infinity. Note that the center of mass motion decouples in both
194: cases, to give \be \dd{t} \int dx~u = \int dx~f, \ee which is but
195: the standard Brownian motion. For this reason one usually selects
196: the force $f$ to be a derivative of another function $h(x,t)$,
197: $f=\partial_x h$. With this restriction, the average velocity
198: vanishes $\VEV{u}=0$ and it makes sense to define the {\sl root
199: mean square velocity} as $U_{\rm rms} = \sqrt{\VEV{u^2}}$. It
200: measures the average amplitude of the fluctuation of the velocity
201: field. A basic assumption in the theory of turbulence is that this
202: quantity does not depend on the viscosity $\nu$ as it is taken to
203: zero. That allows to determine $\Urms$ via simple hydrodynamic
204: scaling to be $ \Urms \propto (L F(0))^{\frac{1}{3}}$.
205:
206:
207: Once the equation \rf{burgers} is written, one would like to calculate various averages and
208: probabilities. A particular popular question focuses on the probability distribution function
209: of the velocity at two different points. Define $P(\sigma)$ as the
210: probability of observing a velocity
211: gradient $\sigma=\partial_x u(x)$ in a particular point in space.
212: Similarly $P_d(v,r)$ is the probability of observing a difference in velocities
213: $v=u(x+r)-u(x)$
214: at two points separated by a distance $r$ from each other.
215: These quantities received a lot of attention
216: in the literature, possibly because they are related
217: to the turbulent advection of particles suspended in the fluid, which is a very important
218: topic in real 3D turbulence.
219:
220: The following is a highly subjective historical overview of the progress made in understanding
221: \rf{burgers}. As always, the choice of highlights is very personal one, and I apologize to the authors
222: whose contributions are not mentioned here.
223:
224: The first to address the behavior of $P_d(v)$ were V. Yakhot and A. Chekhlov \cite{Yakhot}, who
225: argued that it should not be symmetric under $v \rightarrow -v$. They were also the first
226: to study this function numerically.
227:
228: Soon thereafter, A. Polyakov \cite{Polyakov} suggested an approach
229: to calculation of these functions. He suggested an equation that
230: $P_d(v)$ should satisfy. The drawback of his approach was that it
231: was not just one equation but a family of equations parametrized
232: by a parameter (referred to as ``$b$-anomaly'') whose value
233: remained undetermined. At about the same time, Bouchaud, Mezard
234: and Parisi \cite{BMP} attacked the problem using replica trick and
235: the Gaussian variational ansatz. The problem with their approach
236: was that it was not well suited to computing quantities such as
237: $P(\sigma)$ and $P_d(v)$.
238:
239:
240:
241:
242: In subsequent work \cite{GM} of the author with A. Migdal, it was
243: shown that at least the asymptotics of $P(\sigma)$ and $P_d(v)$ at
244: $\sigma, v \rightarrow +\infty$ can be derived carefully in the
245: saddle point approximation. It was shown that \bea \label{RHS}
246: P(\sigma) \propto \exp \left( - \frac{2 \sigma^3}{3 G} \right), \
247: \sigma \gg G^{\frac{1}{3}}\approx {\Urms \over L} \br P_d(v)
248: \propto \exp \left( - \frac{2 v^3}{3 r^3 G} \right), \ v \gg \Urms
249: \eea where $G=-\partial^2_x F(x)|_{x=0}$. The drawback of the
250: saddle point approach was that it allowed to determine the right
251: hand side asymptotics only. But that asymptotics is expected to
252: depend on whether the force $f$ of \rf{burgers} is Gaussian. The
253: asymptotics \rf{RHS} is true only if the random force of the
254: Burgers equation is Gaussian. But it does not have to be. For
255: example, one could take the random force whose strength is always
256: less than a certain value, $|f| < f_{\rm max}$. Then the
257: asymptotics \rf{RHS} would not be expected to hold. But if it is
258: so, then one could legitimately ask if at least some part of the
259: functions $P(v)$ and $P_d$ are universal, that is, they are
260: independent on the details of the random force $f$, except via the
261: force length scale $L$. On top of it, since any random Gaussian
262: force leads to \rfs{RHS}, one could ask if at least for the
263: Gaussian force, the entire functions $P(\sigma)$ and $P_d(v)$ are
264: universal (that is, independent of the force correlation function
265: $F(x)$).
266:
267: Some time later, an important paper by W. E, K. Khanin, A. Mazel
268: and Y. Sinai \cite{Sinai} came out. It was argued in that paper
269: that the asymptotics of $P(\sigma)$ at large negative $\sigma$ is
270: \be \label{LHS} P(\sigma) \propto {1 \over |\sigma|^{7 \over 2}},
271: \ \sigma \ll - \frac{\Urms}{L}. \ee The arguments leading to
272: Eq.~\rf{LHS} seemed to be well defined and valid independent of
273: the details of the random force setup. Those arguments were later
274: elaborated in a later paper by U. Frisch, J. Bec and B. Villone
275: \cite{Frisch}. In all of these papers, the asymptotics \rfs{RHS}
276: was not disputed.
277:
278: At the same time, these authors argued that the only universal
279: feature of the \PDF $P(\sigma)$ is its ``left tail'' \rfs{LHS}.
280: The rest of the function was proposed to be nonuniversal,
281: depending on the details of the random force.
282:
283: These claims were later checked numerically by R. Kraichnan and T.
284: Gotoh \cite{KG}, but they were not able to unambiguously confirm
285: or reject the power law \rf{LHS}. Some time later J. Bec
286: \cite{Bec} did a very thorough numerical analysis of the ``left
287: tail" and concluded that the power law Eq.~\rf{LHS} is indeed
288: correct.
289:
290: Here I would like to report a numerical confirmation of the tail
291: \rfs{LHS}. However, I would also like to suggest that the left
292: tail is not the only universal feature of the function
293: $P(\sigma)$.
294:
295:
296:
297: \section{Discrete Burgers Equation}
298:
299: To study the Burgers equation numerically, its space and time have
300: to be discretized. One way to do that would be to write
301: derivatives as differences. That is a very crude scheme, however.
302: Burgers equation allows a much more gentle way of discretizing its
303: time, which is based on the Hopf-Cole transformation. In this
304: section I am going to present the scheme as it is, and show how it
305: produces the Burgers equation in the continuum limit.
306:
307: Consider the {\sl energy function} $E_i(x)$ where $x$ changes from
308: $0$ to $2 \pi$ and $i$ is the integer index (which will later
309: become time). Now write down the following definition \be
310: \label{KPZ} E_i(x) = \min_{0\le y \le 2 \pi} \left[ 1-\cos
311: \left(x-y \right) +E_{i-1} (y) \right] + h_i(x). \ee Here $h_i$ is
312: a random function, which I take as \be \label{force} h_i(x) = A_i
313: \cos \left( x-\phi_i \right), \ee where $A_i$ is a random variable
314: (for example, random Gaussian with or taking values uniformly
315: distributed from $0$ to some maximum $A_{\rm max}$), and $\phi_i$
316: is a random phase uniformly distributed from $0$ to $2 \pi$. The
317: sign $\min$ in \rfs{KPZ} should be understood as a minimization
318: over $y$ which varies from $0$ to `$2 \pi$ of the expression in
319: the square brackets. \rfs{KPZ} plays the role of a definition of
320: $E_i$ in terms of $E_{i-1}$. As a starting point, one can take
321: $E_0(x)=0$. Finally, the functions $E_i(x)$ will automatically be
322: periodic in $x$.
323:
324: The equation \rfs{KPZ} is the discrete Burgers equation. To see
325: that, let us take the continuum limit. In the notations of
326: \rf{KPZ}, the continuum limit is achieved when $h_i \ll 1$, $E_i
327: \ll 1$. The minimization condition reads \be \sin(x-y)- \partial_y
328: E_{i-1} (y) = 0. \ee Since $E_i$ is small, one can expand \be
329: \label{lagr} x = y + \partial_y E_{i-1}. \ee Substituting it back
330: to \rf{KPZ} and expanding $E_{i-1}(x+y-x)$ in powers of $y-x$, one
331: finds \be E_i(x) - E_{i-1} (x) + \oh \left( \partial_x E_{i-1}(x)
332: \right)^2 = h_i(x). \ee The final step is replacing the discrete
333: index $i$ with the continuous time $t$, to find \be
334: \partial_t E + \oh \left( \partial_x E \right)^2 = h,
335: \ee
336: which is the so-called KPZ equation. Introducing the velocity $u=\partial_x E$, one gets the
337: Burgers equation
338: \be
339: \label{bur}
340: \partial_t u + u~\partial_x u = \partial_x h,
341: \ee with $\partial_x h$ identified as the force $f$. In this
342: equation, the viscosity term $\nu u_{xx}$ is absent. But it is
343: easy to see that find a global minimum in \rf{KPZ} is equivalent
344: to the infinitesimal viscosity term in \rf{bur}, following, for
345: example, Feigelman \cite{Feigelman}. The derivation below is
346: somewhat technical and is given for completeness only, so it is
347: possible to skip the equations Eqs.~\rf{xx1}-\rf{deri} without
348: losing any essential information.
349:
350: Introduce an alternative way of writing down \rf{KPZ}
351: \begin{widetext}
352: \be
353: \label{xx1}
354: \exp \left( - E_i(x_i) \right) = \lim_{\nu \rightarrow 0}
355: \int \prod_{j<i} dx_j~\exp \left\{ - \frac{2}{ \nu} \sum_{j\le i} \left[ 1-\cos \left(x_j-x_{j-1}\right)
356: + h_{j}(x) \right] \right\}.
357: \ee
358: \end{widetext}
359: It is clear that the integral, in the limit of vanishing $\nu$
360: effectively minimizes the expression in the exponential, doing the
361: work which is accomplished by the $\min$ sign in \rf{KPZ}. Passing
362: to the continuum limit, one finds the functional integral
363: \begin{widetext}
364: \begin{equation}
365: \exp \left( -E(y,T) \right) = \lim_{\nu \rightarrow 0} \int {\cal
366: D} x(t) ~\exp \left\{ - \frac{2}{\nu} \int_0^T dt \left[ \oh \dot
367: x^2 + h(x,t) \right] \right\}, \ x(T)=y.
368: \end{equation}
369: \end{widetext}
370: But the right hand side is none other than the propagator of
371: the Schr\"odinger equation, which enables us to conclude that
372: $\exp \left(-E\right)$ satisfies the Schr\"odinger equation and as
373: a consequence, \be \label{deri}
374: \partial_t E + \oh \left(\partial_x E \right)^2 - \nu \partial^2_x E = h,
375: \ \partial_t u + u~\partial_x u - \nu \partial^2_x u = \partial_x
376: h, \ee which concludes the derivation. (In the last line of
377: \rf{deri} I replaced the arguments of the function $E(y,T)$ by $x$
378: and $t$.)
379:
380: All this suggests that the following steps must be taken to
381: simulate the Burgers equation numerically. First, generate the
382: random ``force'' $h_i(x)$ using \rf{force}. $A_i$ can be taken,
383: for example, uniformly distributed on the interval from $0$ to
384: $A_{\rm max}$. Then solve the relationship \rf{KPZ} recursively to
385: find $E_i(x)$. The actual value of $E_i$ is not important for the
386: numerics of the Burgers equation. What is important is the value
387: of $y$ which minimizes the expression in the square brackets of
388: \rf{KPZ} for each value of $x$. Given the relationship between $x$
389: and $y$, one finds for the Burgers velocity $u_i = \partial_x
390: E_i(x)$ \be \label{burd} u_i(x) = u_{i-1}(y) + \partial_x
391: h_{i}(x). \ee This is the discrete analog of the obvious equation
392: $\dd{t} u(x(t),t) = f(x,t)$ where $\dd{t} x(t)= u(x(t),t)$, the
393: so-called Lagrangian trajectory (see below). Finally, for the
394: velocity gradients $\sigma_i(x) = \partial^2_x E_i (x)$ one finds
395: \be \label{sid} \sigma_i (x) = \cos\left(x-y\right)
396: \frac{\sigma_{i-1}(y)} {\cos \left(x-y \right) + \sigma_{i-1}(y)}
397: + \partial^2_x h_{i}(x), \ee which is the discrete version of the
398: \rfs{Lan} considered below. The relations \rfs{KPZ}, \rfs{burd},
399: and \rfs{sid} are those one has to iterate to solve the Burgers
400: equation numerically. To stay close to the continuum limit in
401: time, one has to take $A_{\rm max} \ll 1$. And finally, to do the
402: minimization in \rf{KPZ} in practice one has to discretize space
403: as well. With 2000 spacial discretization point, reliable data can
404: be accumulated over about 10000 steps in time, which takes about
405: 10 minutes computer time on a Sun workstation. The result of one
406: such calculation, with $A_{\rm max} = 0.01$ is shown on Fig 1.
407:
408:
409: \begin{figure}[htbp]
410: \centerline{\epsfxsize=3in \epsfbox{fig1.eps}}
411: \caption{A typical plot of $u(x)$ at a fixed time $t$ taken from an actual computer simulation}
412: \end{figure}
413:
414: The random force chosen according to \rf{force} fixes the function
415: $F$ to be $F(x) \propto \cos (x)$ and the characteristic length
416: scale $L$ is then of the order of the size of the system. If one
417: wants to use a shorter range force, one could take \be
418: \label{force1} h_i(x) = A_i \sum_{n=-\infty}^{\infty}\exp \left\{
419: - \frac{\left(x-2 \pi n+\phi_i \right)^2}{L^2} \right\}, \ee or
420: any other similar function, as long as it is periodic in space
421: $x$.
422:
423: \section {Statistics of the Velocity Gradients}
424:
425:
426: The properties of the solution to Burgers equation has been discussed in many publication (see for example
427: \cite{Frisch}).
428: Here I list those which are relevant for the present discussion.
429:
430: In the Burgers equation, the viscosity $\nu$ is supposed to be infinitesimally small.
431: Yet it cannot be set to zero. The reason for that is, the solution to this equation develop
432: shock waves, or sharp drops in the value of the velocity field. The width of this drops is
433: of the order of $1/\nu$ and they become discontinuities in $u$ when $\nu$ is taken to zero.
434: In the middle of the drop $\nu \partial^2_x u$ has a finite limit as $\nu \rightarrow 0$.
435:
436:
437:
438: To study the gradients of the velocity it is advantageous to
439: introduce the notion of the lagrangian coordinate $x(t)$. It is
440: the point which moves together with the fluid \be \dbyd{x(t)}{t} =
441: u(x(t)). \ee A discrete version of the lagrangian trajectory was
442: given by \rfs{lagr}. Once $x(t)$ reaches a shock wave, it becomes
443: ``trapped'' in it. For the purposes of this paper, it is natural
444: to assume that $x(t)$ ends as soon as it reaches the shock wave
445: (this is precisely what \rfs{lagr} does). Introduce the notion of
446: the gradient of the velocity on a trajectory defined by $x(t)$,
447: \be \label{grad1} \sigma (t) = \partial_x u(x(t),t). \ee As a
448: consequence of the Burgers equation, \be \label{Lan} \dot \sigma +
449: \sigma^2 =
450: \partial_x f(x(t),t). \ee Notice that the $\nu$ term can be
451: neglected since, by construction, $x(t)$ avoids shock waves and
452: terminates once it reaches them.
453:
454: The equation \rfs{Lan} can be interpreted as a Langevin equation
455: with the random force $g(t)=
456: \partial_x f(x(t),t)$.
457: Before proceeding further, one needs to establish the statistics of $g(t)$. $x(t)$ is correlated
458: with $f$ in a complicated way, and it is not {\sl a priori clear} if those correlations can be neglected.
459:
460: Fortunately, the discrete Burgers equation discussed in the
461: previous section provides an answer to this question. $g(t)$ is
462: none other than $\partial^2_x h_i(x)$. But $x$ is related to $y$
463: via \rfs{lagr}, and therefore, $x$ knows nothing about the random
464: phase $\phi_i$ included in the definition of $h_i$ in \rfs{force}.
465: Therefore, $g(t)$ is indeed just a white noise random force \be
466: \VEV{g(t) g(t')} = G~\delta(t-t') \ee where, as before,
467: $G=-\partial^2_x F(x)|_{x=0}\approx(\Urms/L)^3$. Given a Langevin
468: equation \rfs{Lan}, a Fokker-Planck equation can be written down
469: for the propability $P(\sigma)$ \be \pp{\sigma} \left( {G \over 2}
470: \pp{\sigma} + \sigma^2 \right) P = \pbyp{P}{t}. \ee In the present
471: context this equation was first written down in \cite{Bouchaud},
472: but in that paper it was given an interpretation different from
473: the one here. In fact, a slightly more general form of this
474: equation will be useful below. It has the form \be \label{FP}
475: \pp{\sigma} \left( {G \over 2} \pp{\sigma} + \sigma^2 \right) P +
476: \beta P = \pbyp{P}{t} \ee where $\beta$ is an arbitrary parameter.
477: At this stage, $\beta=0$. However, later a version of \rfs{FP}
478: with nonzero $\beta$ will be useful. Notice that the large
479: $\sigma$ asymptotics of $P$ obtained from the \rfs{FP} is
480: \rfs{RHS}, irrespective of the value of $\beta$.
481:
482: \rfs{Lan} is unstable. As soon as $\sigma$ becomes large negative,
483: it quickly reaches infinity. This process is none other than the
484: formation of a shock wave. A quick estimate of the probability
485: $P(\sigma)$ in this regime is easy to do. Once $\sigma$ is large
486: enough and negative, one can neglect $g(t)$ in \rfs{Lan}
487: completely and solve the equation to find \be \sigma(t) =
488: \frac{1}{t-t_0}, \ t < t_0 \ee The probability can be estimated to
489: be \be \label{naive} P(\sigma) \propto \int
490: dt~\delta(\sigma-\sigma(t)) \propto \frac{1}{\sigma^2}, \ \sigma
491: \ll -\frac{\Urms}{L}. \ee This estimate was used in some
492: publications to conclude that this should be the correct power law
493: in the \PDF of the Burgers equation \rfs{LHS}. However, to say so
494: would be premature.
495:
496: It is instructive to reproduce the power law \rfs{naive} with the
497: help of the Fokker-Planck equation. To do that, observe that one
498: has to solve this equation with an absorbing boundary conditions
499: at $\sigma \rightarrow -\infty$, corresponding to the trajectories
500: disappearing when shock waves are formed. The standard
501: substitution \be P=\exp\left( - \frac{\sigma^3}{3 G} \right)
502: \Psi(\sigma). \ee leads to the Schr\"odnger equation of a quantum
503: mechanical particle moving in a potential $U(\sigma) = \sigma^4/(2
504: G) - (\beta+1)\sigma$ \be \label{eq1} -\frac{G}{2}
505: \frac{\partial^2 \Psi}{\partial \sigma^2} +
506: \left(\frac{\sigma^4}{2 G} - \left( \beta + 1 \right) \sigma
507: \right) \Psi = - \pbyp{\Psi}{t}. \ee The solution to this equation
508: reads \be \Psi(\sigma) = \sum_{n=0}^{\infty} C_n \exp \left( - E_n
509: t \right) \Psi_n(\sigma), \ee where $C_n$ are arbitrary
510: coefficients, $E_n$ are the eigenstates, and $\Psi_n$ are the
511: eigenfunctions of the Schr\"odinger equation satisfying \be
512: -\frac{G}{2} \frac{\partial^2 \Psi_n}{\partial \sigma^2} +
513: \left(\frac{\sigma^4}{2 G} - \left( \beta+1 \right) \sigma \right)
514: \Psi_n = E_n \Psi_n. \ee At large times $t \gg (E_1-E_0)^{-1}$,
515: only the ground state survives
516: \begin{eqnarray}
517: \label{eq2}& P(\sigma,t) \propto \exp \left( - E_0 t \right)
518: P_0(\sigma), \ t \gg \left(E_1-E_0\right)^{-1}, \cr & P_0(\sigma)
519: = \exp\left( - \frac{\sigma^3}{3 G} \right) \Psi_0 (\sigma).
520: \end{eqnarray} A detailed analysis of Eqs.~\rf{eq1}-\rf{eq2} at arbitrary
521: $\beta$ can be found in the paper by S. Boldyrev \cite{Boldyrev}.
522: In the case of interest, $\beta=0$. Ref. \cite{Boldyrev} shows
523: that in this case $E_0>0$, and it confirms the asymptotics
524: \rfs{naive} for $P_0(\sigma)$. The asymptotics at large positive
525: $\sigma$ coincides with \rfs{RHS}. The fact that $E_0>0$ is not
526: surprising. It means that the total probability $\int d
527: \sigma~P(\sigma)$ dicreases with time. This is due to elimination
528: of trajectories which ended on shock waves.
529:
530: However, this cannot be the correct probability of observing a velocity gradient $\sigma$ at a given
531: point in space. Indeed, introduce the density of trajectories $\rho(x,t)$. It satisfies
532: the continuity equation
533: \be
534: \partial_t \rho + \partial_x \left(u \rho \right) = 0.
535: \ee In particular, the density around the given Lagrangian
536: trajectory $x(t)$ is given by $\rho_L(t) = \rho(x(t),t)$ and it
537: satisfies \be \dd{t} \rho_L + \sigma \rho_L = 0 \ee It is clear
538: that the probability that a particular trajectory happens to go
539: through an observation point where a velocity gradient is measured
540: is inversely proportional to the density of trajectories at this
541: point. Thus a more physically relevant probability distribution
542: function is defined as \be P(\sigma) = \VEV{ \frac{1}{\rho_L}
543: ~\delta (\sigma-\sigma(t))}. \ee This function satisfies a
544: Fokker-Planck equation \rfs{FP}, but with $\beta=1$. In this case,
545: it is possible to show that $E_0<0$ and \be P(\sigma) \propto
546: \frac{1}{|\sigma|^3}, \ \sigma \ll -\frac{\Urms}{L}. \ee This
547: asymptotics was proposed in the paper \cite{KG}. But it cannot be
548: the correct asymptotics either. Indeed, while the above analysis
549: correctly removes the trajectories which end at the points where a
550: new shock wave is formed, it fails to take into account those
551: which fall into ``mature'' shock waves, the ones which formed in
552: the past. That is why $E_0<0$ and the probability grows with time.
553: That comes from overcounting the trajectories.
554:
555:
556: Unfortunately, the \rfs{Lan} does not seem to contain enough
557: information to determine if there is a shock wave nearby. If that
558: is indeed the case, then the full Burgers equation \rfs{burgers}
559: has to be solved. That is the core of the problem of Burgers
560: turbulence, and the reason the solution to it seems so close and
561: yet so hard to get. The full Burgers equation \rfs{burgers} leads
562: to a full fledged 2D quantum field theory (via Martin-Siggia-Rose
563: formalism, for example). That theory is presumably very hard, if
564: not impossible, to solve. The ``reduced'' equation \rfs{Lan}, on
565: the other hand, leads to a quantum mechanics encoded in \rfs{eq1}.
566: The solution to quantum mechanics problems are clearly within
567: reach. The question is, therefore, if it is possible to extract
568: all the relevant information about the Burgers equation from the
569: Lagrangian equation \rfs{Lan}. At this stage of the present
570: analysis, it does not seem possible.
571:
572: To move further, I will make the following leap of faith. The
573: equation \rfs{FP} was studied in Ref.~\cite{Boldyrev} for
574: arbitrary $\beta$. It was found that while the large positive
575: $\sigma$ asymptotics of $P$ is always given by \rfs{RHS}, the
576: other asymptotics is given by \be P(\sigma) \propto \frac{1}
577: {\sigma^{\beta+2}}, \ \sigma \ll -\frac{\Urms}{L}. \ee On the
578: other hand, it is known from the analysis of Ref.~\cite{Sinai} and
579: \cite{Frisch} that the correct power of this asymptotics is $7/2$.
580: To match this asymptotics, I choose $\beta=3/2$. Then I propose to
581: solve the equation \rfs{FP} and compare what it gives with the
582: numerics. The results of the comparison is shown in next section.
583:
584: The equation \rfs{FP} with an arbitrary $\beta$ was proposed in
585: \cite{Polyakov} (but for $P_d(v)$ as opposed to $P(\sigma)$) where
586: $\beta$ was referred to as $b$-anomaly. But whether the present
587: discussion has anything to do with the methods of
588: Ref.~\cite{Polyakov} is unclear. To derive \rfs{FP} with
589: $\beta=3/2$, one needs to show that the rate at which the
590: trajectories fall into mature shock waves is proportional to the
591: extra $\oh \sigma P$ term generated in the Fokker-Planck equation
592: \rfs{FP}. How to do that is not known to me. Next, the \rfs{FP}
593: cannot be exact. After all, it clearly breaks down even if the
594: force is not Gaussian. However, it may be correct in some mean
595: field sense, for moderate values of $\sigma$. Finally, I'd like to
596: note that for this value of $\beta$, $E_0<0$ and the probability
597: still grows with time. That must be because this description gives
598: the relative probability of observing one value of $\sigma$ over
599: another, not the absolute probability, and the overall
600: normalization must be enforced by $P \rightarrow P/\Pi$ where
601: $\Pi=\int d\sigma~P(\sigma)$.
602:
603:
604: \section{Numerical Simulations}
605: With the setup descibed in the preceeding sections, it is very
606: easy to calculate $P(\sigma)$ numerically. Unlike some of the
607: previous work on this subject, I did not try to compute
608: $P(\sigma)$ from the data. Instead, I computed the integrated
609: probability density \be N(\sigma) = \int_{-\infty}^{\sigma}
610: d\mu~P(\mu). \ee Obviously its left tail is expected to go as
611: $N(\sigma) \propto 1/|\sigma|^{\frac{5}{2}}$. To plot it
612: numerically, it is enough to collect the data into an array, to
613: sort it in the increasing order, and plot the position of an array
614: entry as a function of its value. Fig 2. depicts $N(\sigma)$ in
615: the case when the force was chosen to be as in \rfs{force},
616: $A_{\rm max}=0.01$, and $\Urms \approx 0.02$.
617: \begin{figure}[htbp]
618: \centerline{\epsfxsize=3in \epsfbox{fig2a.eps}}
619: \caption{The integrated probability density $N(\sigma)$ as a function of $\sigma$}
620: \end{figure}
621:
622: Fig 3. shows the left tail in
623: the log-log format. For comparison, a straight line with the slope $-5/2$ is plotted.
624: \begin{figure}[htbp]
625: \centerline{\epsfxsize=3in \epsfbox{fig3a.eps}}
626: \caption{The plot of $\log N$ as a function of $\log(-\sigma)$ for negative $\sigma$. The straight
627: line has a slope of $-2.5$}
628: \end{figure}
629: The agreement is striking. One could object that the power law in
630: this graph does not extend over a sufficiently long range of
631: $\sigma$. To extend it further, one needs to do numerics at weaker
632: force for longer periods of time (to keep close to the continuum
633: limit even at large velocity fluctuations). It was not the task of
634: the present work to do large scale numerical simulations, and I
635: believe the agreement I have here is good enough (especially in
636: view of the work reported in Ref.~\cite{Bec}).
637:
638: Fig 4. shows the same probability function, but plotted together
639: with the (numerical) solution to the equation \rfs{FP} with
640: $\beta$ chosen to be $\beta=3/2$.
641: \begin{figure}[htbp]
642: \centerline{\epsfxsize=3in \epsfbox{fig4a.eps}}
643: \caption{$N(\sigma)$ taken from the numerics (upper curve) and
644: extracted from the solution to the \rfs{FP} with $\beta=3/2$.}
645: \end{figure}
646: To arrive at this picture, one parameter - the units of measure of
647: $\sigma$ - had to be adjusted to make the graphs overlap as much
648: as possible. The agreement is not perfect. Yet the curves are
649: close enough to suggest that perhaps it is not coincidential
650: either. One interesting number is the probability $p$ that the
651: velocity gradient is negative. It is especially interesting,
652: because it is independent of the adjustments of the units of
653: $\sigma$. Numerically $p\approx 0.21\pm 0.01$. The solution to the
654: \rfs{FP} gives about $0.18$. According to some claims in the
655: literature, this number is not universal. However, the value of
656: $p$ is reproducible whether I do the numerics with the long ranged
657: force \rfs{force} or shorter range force \rfs{force1} with various
658: values of $L$. So numerically $p$ appears to be universal, or
659: perhaps approximately universal.
660:
661: To conclude, the numerical data does not confirm that the \rfs{FP}
662: with the parameter $\beta=3/2$ is the {\sl exact} solution to the
663: problem. But it may be an {\sl approximate}, perhaps in some mean
664: field sense, solution to the problem, which reproduces the actual
665: probability density rather close. I believe this question deserves
666: further investigations.
667:
668: \section{Pinned Charge Density Waves}
669:
670: In the papers of the author with J.T. Chalker \cite{GC,GC1} a
671: remarkable correspondence was established between the randomly
672: driven Burgers equation and the physics of pinned charge density
673: waves. It would go beyond this paper to discuss charge density
674: waves, but mathematically the correspondence can be formulated in
675: the following way. Consider the functional
676: \begin{equation}
677: \label{PCDW} {\cal E}[x(t)] = \int_0^T dt \left[ \oh \dot x^2 +
678: h(x,t) \right],
679: \end{equation}
680: where $h(x,t)$ is the same random function as the one in
681: \rfs{xx1}. Let us find a function $x_0(t)$ which, when substituted
682: for $x(t)$ in \rfs{PCDW}, gives an absolute minimum for the
683: functional ${\cal E}$. Obviously it satisfies the minimization
684: equation
685: \begin{equation}
686: -{d^2 x_0\over dt^2} + \left.\pbyp{h(x,t)}{x}\right|_{x=x_0(t)} =
687: 0.
688: \end{equation}
689:
690: Now consider the energy cost of $x(t)$ deviating from $x_0(t)$.
691: Writing $x(t) \approx x_0(t) + \psi(t)$ we find the equation for
692: the normal modes of oscillations about the absolute minimum, given
693: by
694: \begin{equation}
695: \left[ -{d^2 \over dt^2}+ \left. {\partial^2 h(x,t) \over \d x^2}
696: \right|_{x=x_0(t)} \right]\psi = \omega^2 \psi.
697: \end{equation}
698: With the problem thus set up, one needs to calculate the average
699: number of modes with frequencies less than a given frequency
700: $\omega$. Such a function is denoted as $N(\omega)$.
701:
702: This problem, which was first formulated in the context of charge
703: density waves in Ref.~\cite{FL}, received a substantial amount of
704: attention in the literature (see
705: Refs.~\cite{Feigelman,Gia,AR,Fogler,GC,GC1}. After initial
706: disagreements, a consensus was built which gave for the function
707: $N(\omega)$ the value
708: \begin{equation}
709: \label{IDS} N(\omega) \propto \omega^5,
710: \end{equation}
711: for sufficiently small $\omega$. This was first proposed in
712: Ref.~\cite{AR} and then the derivation was improved in subsequent
713: publications until Ref.~\cite{GC} derived \rfs{IDS} in a
714: systematic way.
715:
716: One of the results of Refs.~\cite{GC,GC1} relates the calculation
717: of $N(\omega)$ to the following specific question in Burgers
718: turbulence. Consider a Burgers fluid \rfs{burgers} driven by a
719: random force $f(x,t)=\partial_x h(x,t)$. Consider a Lagrangian
720: trajectory $x_0(t)$ which moves with a fluid and never gets
721: absorbed by a shock wave for times $t<T$. Consider a velocity
722: gradient $\sigma(t)$ defined in \rfs{grad1}. Let us find time
723: intervals $t_1 < t < t_2$ such that $\sigma(t)<0$ within those
724: intervals. Now let us calculate the probability that
725: \begin{equation}
726: \label{cond1} \int_{t_1}^{t_2} dt~\sigma(t) < \log(\omega),
727: \end{equation}
728: for some small value of $\omega$. Then the function $N(\omega)$ of
729: the charge density wave problem coincides with this probability.
730:
731: Notice that knowing the tails of the probability distribution
732: $P(\sigma)$ would not help calculating $N(\omega)$. This is
733: because the typical functions $\sigma(t)$ which satisfy
734: \rfs{cond1} are not those for which $\sigma$ is very large
735: negative, but rather moderate negative $\sigma(t)$ which however
736: persist over long time intervals. Therefore, the results of W. E
737: {\sl at al} \cite{Sinai} are useless if we were to calculate
738: $N(\omega)$ with the help of \rfs{cond1}.
739:
740: Let us however use the methods of this paper, in particular
741: \rfs{eq1}, to calculate $N(\omega)$. To do so, we need to write
742: down a Feynman path integral formalism which corresponds to
743: \rfs{eq1} and then use the Lagrange multiplier method described in
744: Ref.~\cite{GC1}, section VI C, subsection 1, corrected however for
745: the presence of a parameter $\beta$ in \rfs{eq1}. Then we find
746: \begin{equation}
747: N(\omega) \propto \omega^{2 \beta+2}.
748: \end{equation}
749: Now we know the exact answer to the problem, \rfs{IDS}. If the
750: methods described in this paper were to reproduce the exact
751: answer, we have to choose $\beta=3/2$. However, {\sl this is the
752: same value of $\beta$ required to reproduce the asymptotics of W.
753: E {\sl et al}!}
754:
755: Two possible explanations of this are possible. First is, perhaps
756: the \rfs{eq1} with $\beta=3/2$ indeed reproduces the right
757: asymptotics of W. E {\sl et al} and for some reason also
758: reproduces the function $N(\omega)$ correctly, but this equation
759: is still generally not correct and the probability distribution
760: derived with its help and shown on Fig. 4 is not correct either.
761: Second is, the \rfs{eq1} is indeed correct and gives the correct
762: solution to the Burgers problem.
763:
764: The first explanation seems far fetched, as a remarkable
765: coincidence must be at play to give rise to it. The second
766: explanation sounds much more likely. And yet, no derivation of
767: \rfs{eq1} is known at this point.
768:
769:
770:
771:
772:
773:
774: \section {Conclusion}
775: In this paper, I demonstrate that it is easy to obtain reliable
776: numerical data on the behavior of the 1D Burgers equation. I show
777: that the data confirms the $7/2$ tail of the probability
778: distribution suggested in the literature. The data also suggests
779: that the entire probability distribution function could be
780: universal, determined by the solution to the equation \rfs{FP}
781: with the ``anomaly'' parameter $\beta=3/2$. I also discuss that it
782: is possible to reproduce the known solution to the pinned charge
783: density wave problem by applying the techniques discussed here and
784: choosing $\beta=3/2$, which gives extra weight to these methods.
785:
786: \section{Acknowledgements}
787: I am grateful to J.T. Chalker for many discussions. This paper is
788: based on the results obtained in part in the course of work on
789: paper \cite{GC}. The work presented here was supported in part by
790: EPSRC through its Advanced Research Fellowship programme.
791:
792:
793: \begin{thebibliography}{1}
794: \bibitem{Yakhot}
795: A. Chekhlov, V. Yakhot, Phys. Rev. E{\bf 51}, R2739 (1995); Phys. Rev. E{\bf 52}, 5681 (1995)
796:
797: \bibitem{Polyakov}
798: A. Polyakov, Phys. Rev. E{\bf 52}, 6183 (1995)
799:
800: \bibitem{BMP}
801: J.P. Bouchaud, M. Mezard, G. Parisi, Phys. Rev. E{\bf 54}, 3656 (1995)
802:
803: \bibitem{GM}
804: V. Gurarie, A. Migdal, Phys. Rev. E{\bf 54}, 4908 (1996)
805:
806: \bibitem{Sinai}
807: W. E, K. Khanin, A. Mazel, and Y. Sinai, Phys. Rev. Lett. {\bf
808: 78}, 1904 (1997)
809:
810: \bibitem{Frisch}
811: U. Frisch, J. Bec, B. Villone, Physica D vol. 152-153, 620 (2001)
812:
813: \bibitem{KG}
814: T. Gotoh, R. Kraichnan, Phys. Fluids {\bf 10}, 2859 (1998)
815:
816: \bibitem{Bec}
817: J. Bec, Phys. Rev. Lett. {\bf 87}, 104501, (2001)
818:
819: \bibitem{Feigelman}
820: M. Feigelman, Sov. Phys. JETP {\bf 52}, 555 (1980)
821:
822: \bibitem{Bouchaud}
823: J.P. Bouchaud, M. Mezard, Phys. Rev. E{\bf 54}, 5116 (1996)
824:
825: \bibitem{Boldyrev}
826: S. Boldyrev, hep-th/9707255
827:
828: \bibitem{GC}
829:
830: V. Gurarie, J.T. Chalker, Phys. Rev. Lett. {\bf 89}, 136801 (2002)
831:
832: \bibitem{GC1} V. Gurarie, J.T. Chalker, cond-mat/0305445,
833: submitted to Phys. Rev. B
834:
835: \bibitem{FL} H. Fukuyama and P.A. Lee, Phys. Rev. B {\bf 17}, 535 (1978)
836:
837:
838: \bibitem{Gia} T. Giamarchi and H.J. Shultz, Phys. Rev. B {\bf 37}, 325
839: (1988)
840:
841: \bibitem{AR} I. L. Aleiner and I. M. Ruzin, Phys. Rev. Lett. {\bf 72},
842: 1056 (1994)
843:
844: \bibitem{Fogler} M. Fogler, Phys. Rev. Lett. {\bf 88}, 186402
845: (2002)
846:
847: \end{thebibliography}
848: \end{document}
849: